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/9512088
1
9512
1995-12-13T16:13:45
Galaxy formation and evolution: what to expect from hierarchical clustering models
[ "astro-ph" ]
We give a brief review of current theoretical work in galaxy formation. Recent results from N-body and N-body/hydrodynamic simulations, and from semianalytic modelling are discussed. We present updated versions of some figures from Cole et al (1994). In particular, we show the effect of using the revised stellar population synthesis model of Bruzual and Charlot, which results in a much better match to the observed colour distribution of galaxies than before. We also compare the model output with recently available data on the galaxy luminosity function and the redshift distribution of galaxies in the B and K bands. The form of the Tully-Fisher relation at high redshift predicted by our semi-analytic scheme for galaxy formation is given.
astro-ph
astro-ph
Galaxy formation and evolution: what to expect from hierarchical clustering models C.S. Frenk, C.M. Baugh and S. Cole Physics Department, University of Durham, UK Abstract. We give a brief review of current theoretical work in galaxy formation. Recent results from N-body and N-body/hydrodynamic sim- ulations, and from semianalytic modelling are discussed. We present updated versions of some figures from Cole et al (1994). In particular, we show the effect of using the revised stellar population synthesis model of Bruzual and Charlot, which results in a much better match to the observed colour distribution of galaxies than before. We also compare the model output with recently available data on the galaxy luminosity function and the redshift distribution of galaxies in the B and K bands. The form of the Tully-Fisher relation at high redshift predicted by our semi-analytic scheme for galaxy formation is given. 1. Introduction In hierarchical clustering theories of galaxy formation, galaxies form by gas cool- ing and condensing into dark matter halos which, in turn, form by a hierarchy of mergers (White & Rees 1978). The context in which this process takes place is specified by a cosmological model that determines the spectrum of primordial density fluctuations and the rate at which they grow by gravitational instability. The best known example of such a model is the cold dark matter (CDM) model (see Frenk 1991 for a review), but a number of alternatives (mostly variants of CDM), have recently become popular in response to new data on large-scale structure and the COBE detection of anisotropies in the microwave background radiation. Regardless of the specific cosmological model that one wishes to con- sider, there are at least six distinct physical processes that need to be included in any theory of galaxy formation: • 1. The growth of dark matter halos by accretion and mergers. • 2. The dynamics of cooling gas. • 3. Star formation. • 4. Energy feedback into prestellar gas from the products of stellar evolu- tion. • 5. Evolution of the stellar populations that form. • 6. Galaxy mergers. 1 A number of theoretical tools have been developed over the years to inves- tigate these processes, both individually and collectively. N-body simulations have led to significant progress in understanding process (1.), while the recently developed N-body/hydrodynamic techniques are beginning to address processes (1-4) and (6). In addition, semianalytic modelling, a relatively new tool, can treat all six processes together and thus explore the effects of different assump- tions on the properties of the galaxy population as a whole. In this review, we will outline some of the areas where progress has been made and highlight some as yet unresolved issues. 2. Physical processes 2.1. Evolution of dark matter halos The main features of the formation of dark matter halos by hierarchical cluster- ing were already established in N-body simulations carried out a decade ago (eg. Frenk et al. 1985, 1988; Efstathiou et al. 1988). A protohalo perturbation, ini- tially expanding at a reduced rate, collapses, often into filamentary or sheet-like structures, which subsequently break up into roughly spherical lumps. These merge together producing a centrally concentrated and essentially smooth dark halo. This process is illustrated in Figure 1 which shows the development of a galactic halo in a flat 'low'-Ω CDM model. One of the main early results from N-body simulations was the realisation that the rotation curves of dark galactic halos in the standard CDM model are approximately flat, suggesting an explanation for the inferred structure of the halos of spiral galaxies (Frenk et al. 1985, Quinn et al. 1986). These simulations, however, were limited in particle number and did not resolve the inner regions of the halos where the visible galaxy actually forms. This issue has recently been addressed in a series of high-resolution simulations by Navarro, Frenk & White (1995). The density profiles of galactic halos in the CDM model show noticeable departures from an r−2 law, gently sloping from r−1 near the centre to r−3 near the virial radius. When the gravitational effect of a disk is included, the resulting rotation curves agree well with observations of galaxies, from dwarfs to bright galaxies, provided the disks fulfill two conditions: (i) their stellar mass-to- light ratio increases roughly as L1/2 and (ii) the baryon fraction increases with >200kms−1, luminosity such that for galaxies with observed circular velocity, Vc∼ there is only a weak of correlation between this velocity and total halo mass. It is unclear whether the observed disks of spirals satisfy these conditions. A second important early result concerns the angular momentum of galactic halos. This is acquired through tidal torques and, in the linear regime grows lin- early with time (eg White 1995). Tidal effects during merging events efficiently transfer the angular momentum invested in the orbits of the merging subclumps into the outer halo and, as a result, the inner parts of merger remnants end up rotating slowly (Frenk et al. 1985, Barnes & Efstathiou 1987, Cole & Lacey 1995). This non-linear feature has often been invoked as a possible explanation for the low rotation speeds of elliptical galaxies. Figure 1. The formation of a galactic dark matter halo in an N- body simulation. The left-hand column shows the projected particle distribution, in comoving coordinates, of a cubical region of present comoving length 27 Mpc. The right hand column shows, now in phys- ical coordinates, the growth of the large clump seen in the bottom left of the region. Each panel on the right hand row has length 3.8 Mpc. From top to bottom the epochs shown correspond to z = 5, 0.5 and 0. The parameters of the simulation are: mean density, Ω = 0.3; cosmo- logical constant, Λ = 0.7; Hubble constant, H0 = 100hkm s−1 Mpc−1, with h = 0.7; and spectrum normalisation, σ8 = 1.14 (as inferred from the COBE data). The simulation followed 262144 particles and was performed the T3D parallel supercomputer at Edinburgh. 2.2. The dynamics of cooling gas The main ideas here were put forward nearly twenty years ago by Rees & Ostriker (1977), Silk (1977) and Binney (1977). When a dark matter halo collapses, any gas admixed with it will also collapse, but whereas the dark matter free streams, the gas is shock heated. These early papers assumed that shocks would heat up the gas to the virial temperature of the halo, an assumption verified -- in the non-radiative limit -- in the N-body/hydrodynamic simulations of Evrard (1990). These and subsequent simulations (eg. Katz & White 1993, Navarro, Frenk & White 1995) also showed that, in this limit, the gas acquires a density profile that closely parallels that of the dark matter. Rees & Ostriker argued that if the cooling time of virialised gas was shorter than its dynamical time, the gas would collapse to make a galaxy. White & Rees (1978) recognized, however, that in a hierarchical model this simple scheme would lead to a cooling catastrophe since at early times the density is so high that all the gas would cool into subgalactic lumps where it would presumably turn into stars. This patently did not happen in the universe - there is still plenty of gas around today. White & Rees solved this problem by introducing the idea of feedback, whereby the energy released by supernovae associated with an early generation of stars reheats some of the gas before it has had a chance to condense into halos at high redshift. (Efstathiou 1992 has argued that photoionisation by a UV background at high redshift would have a similar effect.) Testing these simple physical arguments in numerical simulations is difficult because the propensity of the gas to cool at high density implies that the be- haviour of the gas is always determined by the resolution limit of the simulation. This numerical artefact, however, can be turned to advantage if it is loosely in- terpreted as an effective source of feedback. N-body/hydrodynamic simulations of representative cosmological volumes in which the gas is allowed to cool are still at an early stage (eg Katz et al. 1992, Cen & Ostriker 1992, Frenk et al. 1995) and show that the behaviour of the gas is more complex than expected in the simple analytic picture. Simulations of the formation of individual galaxies produce disks, often with beautiful spiral arms (e.g. Steinmetz & Muller 1995), but these disks rotate much too slowly. This is because merger events transfer angular momentum from gas fragments to the outer dark matter halo in much the same way as the mergers of collionsless particles do (Navarro, Frenk & White 1995). This angular momentum problem for disks remains a major unresolved issue in studies of galaxy formation. One possible solution may be, again, to invoke some form of feedback which might keep the gas hot and allow it to cool slowly rather to be collected in subclumps. 2.3. Star formation and feedback Current understanding of star formation and the attendant feedback, in the context of galaxy formation, is laughably poor. All that can be done at present is to try and model these processes in a heuristic fashion. For example, in an N-body/hydrodynamic simulation one can stipulate a number of conditions for gas to turn into stars, eg, that it be cool and dense (ie above the Jeans mass) and that it be inflowing into a halo. Systematic tests of such algorithms are just beginning (e.g. Navarro & White 1993). 2.4. Galaxy mergers Simulations of the merging of individual galaxy pairs or small groups have a long and distinguished history (see for example Barnes (1996)). Such simulations ad- dress issues such as the structure and rotation properties of merger remnants, or the gas flows triggered by mergers. From the point of view of galaxy formation in general, a key issue is the relative timescale for the merging of dark matter halos and the galaxies they harbour. As a consequence of their higher binding energy, galaxies take longer to merge than their halos. Furthermore, the simu- lations show that galaxies (or at any rate the clumps of cool gas identified with galaxies in the models) merge on a dynamical friction timescale, provided that the mass that is input into Chandrasekhar's classic formula is the total, gas plus dark matter, mass of the merging satellite (Navarro, Frenk & White 1995). 3. Semianalytic models Our understanding of the full range of complex phenomena listed in the Intro- duction can be approximated by a set of simple rules. These rules can then form the basis of a semianalytic model for galaxy formation, that follows the collapse and mergers of dark matter halos and the star formation histories of the galaxies. Such models (Kauffman et al. 1993, Lacey et al. 1993, Cole et al. 1994) have been successful in accounting for the general properties of the observed galaxy distribution, such as the shape of the luminosity function, faint number counts and colours. However, a number of fundamental problems remain that appear to suggest that the modelling of the processes (1 - 6) needs to be improved, rather than altering the choice of cosmology (Heyl et al. 1994). In the original model of Cole et al. (1994), the model galaxies were not as red as many observed ellipticals. A study of several stellar population codes (Charlot et al. 1995) has led to a revision of the Bruzual and Charlot (1993) models. This has resulted in the model galaxies being typically 0.2 mag redder in B − K. An updated comparison of the colour distribution of the model galaxies from Cole et al. with observed colours is given in Figure 2. The luminosity function of Cole et al. was flatter than that achieved by other semianalytic models, because of the strong feedback adopted, which severely restricts star formation in halos of low circular velocity. However, this model still predicts more faint galaxies than are observed (Loveday et al 1992), though recent results indicate that the faint end of the luminosity function is still uncertain (McGaugh 1994, Marzke et al. 1994). A comparison of the luminosity function predicted by our model with the data of Loveday et al. and Marzke et al. is given in Figure 3. The lower panel shows the comparison with the K-band data of Mobasher et al. (1986) and Glazebrook et al. (1994). Following Glazebrook et al. , we have corrected the Mobasher et al. magnitudes by +0.22 mag., to compensate for the different k corrections used, and we have applied a -0.3 correction to Glazebrook et al. 's magnitudes, so that they correspond to the 40h−1Kpc aperature used by Mobasher et al. . The Tully-Fisher relation recovered by Cole et al gives a good match to the observed scatter and slope at zero redshift (see Figure 11 of Cole et al). However, there is an offset between the observed relation and the prediction of the model, which suggests that the model galaxies are either too faint by about Figure 2. Histograms of B-K colour distributions for various ranges of B absolute magnitude. The broken lines are data from Mobasher et al. (1986) and show the observed number of galaxies in the data set. The model output is shown by the solid lines which have been normalised to enclose the same area as the data. The luminosity function of our model (solid line) compared Figure 3. with the B-band data of Loveday et al. (1992) and Marzke et al. (1994).We have normalised the model to match the knee of the B-band luminosity function. The lower panel shows the comparison with the K-band data of Mobasher et al. (1986) and Glazebrook et al. (1994). Figure 4. The Tully-Fisher relation predicted by our model as a function of redshift. The solid line shows the median magnitude in bins of circular velocity at z = 0; the dashed lines show the 20 and 80th percentiles. The points show the median and the errorbars show the location of the percentiles at z = 0.5. Figure 5. The redshift distribution of model galaxies compared with new data for K and B-band selected samples of Glazebrook et al. 1995a,b. The histograms show the observed distributions and the curves show the model galaxy redshifts. The K band data consists of 124 redshifts for galaxies with K < 17.3 and is weighted for incom- pleteness. The B band sample is 70% complete and contains 70 galaxies with 22.5 < B < 24. Figure 6. The median redshift as a function of limiting apparent magnitude in the B and K bands. The open points show the predictions of our semi-analytic model. 1.5 mag or are in halos that have a circular velocity that is ∼ 60% too high. This can be traced back to an overproduction in CDM-like cosmologies of halos typical of those that contain luminous galaxies. Both the Tully-Fisher and luminosity function problems could be related to surface brightness effects. We plan to incorporate a scale length into the models, allowing us to select only those galaxies above some surface brightness threshold. Figure 4 shows the evolution of the Tully-Fisher relation predicted by our model. The solid line shows the median magnitude in bins of circular velocity at z = 0. The dashed lines show the location of the 20th and 80th percentiles. The points show the position of the median magnitude as a function of circular velocity at redshift z = 0.5; the errorbars here indicate the location of the 20 and 80th percentiles. We also present an updated version of the redshift distributions predicted by the model of Cole et al. Figure 5 compares the model predictions with the recent redshift survey data of Glazebrook et al. (1995a,b). The model and observed distributions are in very good agreement. We plot the median redshift as a function of limiting apparent magnitude in Figure 6. The predictions of our model agree well with the faint redshift data currently available. We have extended the model to split the light of each galaxy up into a bulge and a disk component. Stars are formed in a disk when gas is accreted from the dark matter halo. Bulges are formed in violent merger events, which destroy the disks of the progenitors and are accompanied by a burst of star formation. This allows us to make a broad morphological classification of our galaxies and make predictions of galaxy properties as a function of bulge to disk luminosity ratio. We set the parameters that define the strength of a merger event and the bulge to disk ratios that distinguish between different morphological types by requiring that our model reproduces the local morphological mix. We are then able predict the mean colour and scatter in colour for different morphological classes in different environments, the mix of types in different environments as a function of redshift and the faint counts for the various galaxy types. We find good agreement with the faint HST counts of Glazebrook et al. (1995c) and recover the type of evolution in cluster membership reported by Butcher & Oemler (1978) (Baugh et al. 1995). References Barnes, J., 1996, to appear in the proceedings of IAU 171, "New Light on Galaxy Evolution" eds., Bender, R., Davies, R.L., (Kluwer) Barnes, J., Efstathiou, G., 1987, Ap.J, 319, 575 Baugh, C.M., Cole, S., Frenk, C.S., 1995, M.N.R.A.S, submitted Binney, J.J., 1977, Ap.J, 215, 483 Bruzual, G., Charlot, S., 1993, Ap.J, 405, 538 Butcher, H., Oemler, A., 1978, Ap.J, 219, 18 Cen, R., Ostriker, J.P., 1992, Ap.J, 339, 331 Charlot, S., Worthey, G., Bressan, A., 1995 Ap.J, in press Cole, S., Lacey, C.G., 1995 M.N.R.A.S, submitted Cole, S., Aragon-Salamanca, A., Frenk, C.S., Navarro, J., Zepf, S., 1994, M.N.R.A.S, 271, 781 Colless, M.M., Ellis, R.S., Taylor, K., Hook, R.N., 1990, M.N.R.A.S, 244, 408 Colless, M.M., Ellis, R.S., Broadhurst T., Taylor, K., Peterson, B.A., 1993, M.N.R.A.S, 261, 19 Efstathiou, G. 1992, M.N.R.A.S, 256, 43p Efstathiou, G., Frenk, C.S., White, S.D.M., Davis, M., 1988, M.N.R.A.S, 235, 715. Evrard, A.E. 1990, Ap.J., 363, 349. Frenk, C.S. 1991, in The Birth and Early Evolution of our Universe, Nobel Symp. No 79, eds J.S. Nilsson, et al. , World Sci., Physica Scripta, T36, 70. Frenk, C.S., White, S.D.M., Efstathiou, G., Davis, M., 1985, Nature, 317, 595. Frenk, C.S., White, S.D.M., Efstathiou, G., Davis, M., 1988, Ap.J., 327, 507. Glazebrook, K., Peacock, J.A., Collins, C.A., Miller, L., 1994, M.N.R.A.S, 266, 65 Glazebrook, K., Peacock, J.A., Miller, L., Collins, C.A., 1995a, M.N.R.A.S, 275, 169 Glazebrook, K., Ellis, R., Colless, M., Broadhurst, T., Allington-Smith, J., Tanvir, N., 1995b, M.N.R.A.S, 273, 157 Glazebrook, K., Ellis, R., Santiago, B., Griffiths, R., 1995c, M.N.R.A.S, 275, L19 Heyl, J.S., Cole, S., Frenk, C.S., Navarro, J., 1995, M.N.R.A.S, 274, 755 Katz, N., Hernquist, L., Weinberg, D.H., 1992, Ap.J, 399, L109 Katz, N., White, S.D.M., 1993, Ap.J, 412, 455 Kauffmann, G., White, S.D.M., Guiderdoni, B., 1993, M.N.R.A.S, 264, 201 Lacey, C.G., Guiderdoni, B., Rocca-Volmerange, B., Silk, J., 1993, Ap.J, 402,15 Loveday, J., Peterson, B.A., Efstathiou, G., Maddox, S.J., 1992, Ap.J, 390, 338 McGaugh, S., 1994, Nature, 367, 538 Marzke, R.O., Geller, M.J., Huchra, J.P., Corwin, H.G., 1994, A.J., 108, 437 Mobasher, B., Sharples, R.M., Ellis, 1986, M.N.R.A.S, 223, 11 Navarro, J.F., White, S.D.M., 1993, M.N.R.A.S, 265, 271 Navarro, J.F., Frenk, C.S., White, S.D.M., 1995, M.N.R.A.S, in press Rees, M.J., Ostriker, J.P., 1977, M.N.R.A.S, 179, 541 Silk, J., 1977 Ap.J, 211, 638 Songaila, A., Cowie, L.L., Hu, E.M., Gardner, J.P., 1994, Ap. J. Suppl., 94, 461 Steinmetz, M., Muller, E., 1995, M.N.R.A.S, 276, 549 Quinn, P.J., Salmon, J.K. and Zurek, W. 1986, Nature, 322, 329. White, S.D.M., 1995, Les Houches Lecture Notes White, S.D.M., Rees, M.J. 1978, M.N.R.A.S, 183, 341. Young, P., Lucey, J.L., 1995, in preparation
astro-ph/0511739
1
0511
2005-11-25T22:29:06
Angular Momentum Transport by Gravity Waves in the Solar Interior
[ "astro-ph" ]
We present self-consistent numerical simulations of the sun's convection zone and radiative interior using a two-dimensional model of its equatorial plane. The background reference state is a one-dimensional solar structure model. Turbulent convection in the outer convection zone continually excites gravity waves which propagate throughout the stable radiative interior and deposit their angular momentum. We find that angular velocity variations in the tachocline are driven by angular momentum transported by overshooting convective plumes rather than the nonlinear interaction of waves. The mean flow in the tachocline is time dependent but not oscillatory in direction. Since the forcing in this shallow region can not be described by simple linear waves, it is unlikely that the interaction of such waves is responsible for the solar cycle or the 1.3 year oscillation. However, in the deep radiative interior, the interaction of low amplitude gravity waves, continually excited by the overshooting plumes, is responsible for the angular velocity deviations observed there. Near the center of the model sun the angular velocity deviation is about two orders of magnitude greater than that in the bulk of the radiative region and reverses its direction (prograde to retrograde or vice versa) in the opposite sense of the angular velocity deviations that occur in the tachocline. Our simulations thus demonstrate how angular velocity variations in the solar core are linked to those in the tachocline, which themselves are driven by convective overshooting.
astro-ph
astro-ph
Angular Momentum Transport by Gravity waves in the Solar Interior Tamara M. Rogers Astronomy and Astrophysics Department, University of California, Santa Cruz, CA 95064 [email protected] Gary A. Glatzmaier Earth Sciences Department, University of California, Santa Cruz, CA 95064 ABSTRACT We present self-consistent numerical simulations of the sun's convection zone and radiative interior using a two-dimensional model of its equatorial plane. The background reference state is a one-dimensional solar structure model. Turbulent convection in the outer convection zone continually excites gravity waves which propagate throughout the stable radiative interior and deposit their angular mo- mentum. We find that angular velocity variations in the tachocline are driven by angular momentum transported by overshooting convective plumes rather than the nonlinear interaction of waves. The mean flow in the tachocline is time de- pendent but not oscillatory in direction. Since the forcing in this shallow region can not be described by simple linear waves, it is unlikely that the interaction of such waves is responsible for the solar cycle or the 1.3 year oscillation. However, in the deep radiative interior, the interaction of low amplitude gravity waves, continually excited by the overshooting plumes, is responsible for the angular velocity deviations observed there. Near the center of the model sun the angular velocity deviation is about two orders of magnitude greater than that in the bulk of the radiative region and reverses its direction (prograde to retrograde or vice versa) in the opposite sense of the angular velocity deviations that occur in the tachocline. Our simulations thus demonstrate how angular velocity variations in the solar core are linked to those in the tachocline, which themselves are driven by convective overshooting. Subject headings: convection: overshoot,mixing, internal gravity waves, solar interior -- 2 -- 1. Introduction Internal gravity waves are ubiquitous in nature. Their influence can be observed in striated cloud structures in our own atmosphere many days of the year. In addition to the visual display of these waves in our atmosphere, they have several other more profound consequences. The interaction of gravity waves with angular velocity shear produces the quasi-biennial oscillation (QBO), which dominates variability in the equatorial stratosphere and affects ozone levels. Gravity waves have been invoked to explain several physical problems in the sun's in- terior such as: (1) providing the extra mixing required to solve the Li depletion problem (Garcia-Lopez & Spruit 1991) (2) increasing turbulent mixing and affecting the solar neu- trino production (Press 1981), (3) maintaining the solid body rotation of the sun's radiative interior (Schatzman 1993, Kumar & Quataert 1997) (4) and controlling the solar cycle (Ku- mar, Talon & Zahn 1999). They have also been used to explain the orbital properties of binary star systems (Terquem et al. 1998). These theories, however, are only as good as the poorly understood excitation and evolution of gravity waves in stellar interiors. What grav- ity wave spectra and amplitudes are generated? How effectively do gravity waves transport angular momentum and mix species? Some confusion has existed. For example, although gravity waves were postulated to extract angular momentum from the solar radiative interior and so enforce solid body rotation (Schatzman 1993, Kumar & Quataert 1997), it was quickly pointed out (Gough & McIntyre 1998, Ringot 1998) that gravity waves tend to enhance local shear rather than smooth it. Recognizing this anti-diffusive nature of gravity waves in shear flows, several authors published papers postulating an angular velocity oscillation at the base of the solar convection zone analogous to the quasi-biennial oscillation (QBO) in the Earth's stratosphere (Baldwin et al. 2001). In one paper (Kumar, Talon & Zahn 1999), a spectrum of gravity waves generated by the overlying convection is prescribed (Goldreich et al. 1994). This spectrum is then integrated to give a flux of angular momentum that is transferred from the waves to the mean flow, resulting in a periodic oscillation of angular velocity at the base of the convection zone with a timescale of about 20 years. In another paper (Kim & MacGregor 2001), a two wave model is assumed: one prograde propagating wave, with a prescribed angular momentum flux, and one retrograde wave with another (negative) prescribed flux. In addition, these authors include viscous dissipation in the evolution equation for the mean flow and find that the nature of the resulting angular velocity oscillation depends sensitively on the value of the assumed viscous diffusivity. For large viscous diffusivity, a steady solution is found; whereas for small values a chaotic oscillation is found. A periodic solution is recovered only for intermediate values of the viscous diffusivity. -- 3 -- More recently, the Kumar, Talon & Zahn (1999) model has been extended to show that the oscillating shear layer at the base of the convection zone acts as a filter on the low fre- quency, short wavelength waves (Talon, Kumar & Zahn 2002, Talon & Charbonnel 2005). In their theory, the filter preferentially damps prograde waves, allowing predominantly ret- rograde waves, which carry negative angular momentum, into the deep interior. When these waves dissipate they transfer their negative angular momentum to the flow and therefore effectively extract prograde angular momentum from the low-latitude deep solar interior, leading to solid body rotation in the solar radiative zone. The major shortcoming of both of these models is twofold. First, neither model self- consistently calculates the generation of the gravity waves by the solar convection zone. In Kumar, Talon & Zahn (1999) a spectrum and amplitude of gravity waves is assumed, a spectrum that is unfortunately, untestable with observations and has not been reproduced in numerical simulations. In addition, this spectrum neglects the main production of gravity waves due to overshooting plumes. Second, and perhaps more importantly, neither model calculates the nonlinear wave-wave interactions which provide this "flux" of angular mo- mentum. Rather than parameterizing the nonlinear interaction of waves as a prescribed flux (Kumar, Talon & Zahn 1999 and subsequent papers, Kim & MacGregor 2001), this interaction should be self-consistently calculated. Here we present numerical simulations that address both of these issues. Our model solves the fully nonlinear Navier-Stokes equations in both the convective and radiative re- gions. Therefore, the gravity waves are self-consistently generated by an overlying convection zone and the nonlinear terms are retained in the radiation zone to account for the nonlinear wave interactions which affect the mean flow. 2. Wave - shear flow interactions In this section we briefly review the hydrodynamic process by which waves can transport their angular momentum to the flow in which they travel. The natural example of this process is the QBO. For a more complete discussion of the QBO see, for example, Baldwin et al. 2001, Lindzen (1990) and Holton (1994). For simplicity consider the Boussinesq equations for a viscous fluid in two dimensional cartesian (x-z) coordinates, x being the horizontal coordinate and z the vertical coordinate. After decomposing the horizontal velocity into mean flow (U , a function of z) and fluctu- ating (u') components and taking the proper horizontal average, the equation for the mean horizontal velocity becomes: -- 4 -- ∂U ∂t = − ∂u′w′ ∂z + ν ∂2U ∂2z (1) In (1), u′ represents the fluctuating horizontal velocity and w′ represents the fluctuating vertical velocity. This equation represents how small scale wave-wave interactions (Reynolds stresses) influence the mean flow (U). The vertically propagating waves carry horizontal momentum vertically. Eliassen & Palm (1961) showed that (in this simplified model with no viscosity) the first term on the right hand side of (1) is zero unless there is some wave attenuation (such as radiative damping or critical layers). Therefore, in the absence of wave attenuation, there is no transfer of angular momentum between the mean flow, U , and the Reynolds stresses. In the sun the proposed mechanism for wave attenuation is via radiative diffusion, while in the Earth the process is likely wave breaking. In the presence of differential rotation the picture becomes more complicated. In a rotating fluid, waves generated at a particular frequency are doppler shifted away from that frequency according to the equation: ω(r) = ωgen + m(Ωgen − Ω(r)) (2) where m is the horizontal wave number, ωgen is the frequency at which the wave is generated in the frame rotating at Ωgen and ω(r) is the frequency measured relative to the local rotation rate, Ω(r). Where Ω(r) > Ωgen prograde waves (m > 0) are shifted to lower frequencies and retro- grade waves (m < 0) are shifted to higher frequencies. Since radiative damping is strongly frequency dependent (damping length ∝ ω4), prograde (lower frequency) waves are damped in a shorter distance than retrograde (higher frequency) waves. This differential damping causes prograde waves to deposit their (positive) angular momentum closer to the genera- tion site than retrograde waves deposit their (negative) angular momentum, thus leading to a shear layer. As prograde (retrograde) waves continually deposit their positive (negative) angular momentum, the angular velocity amplitude and gradient continually increase. This increased angular velocity causes prograde waves to be shifted to ever smaller frequencies which are damped even closer to the generation site. In this way the peak in the prograde layer moves toward the source of the waves. When the prograde shear becomes sufficiently steep, it is broken down by viscous diffusion, leaving behind the retrograde layer. This pro- cess repeats, with the period of the process inversely proportional to the wave forcing. The prominent features of this physical process are then twofold: (1) prograde flow lies above retrograde motion (or vice versa) and (2) this pattern propagates toward the generation site. This physical picture was proposed initially by Lindzen & Holton (1968) and Holton & Lindzen (1972) to explain the QBO observed in the Earth's atmosphere. Later, the -- 5 -- physical theory was tested in the remarkable experiment by Plumb & McEwan 1978 and the basic physical mechanism was recovered. It has also been suggested that an oscillation similar to the QBO occurs in Jupiter's atmosphere (Leovy et al. 1991) and is coined the quasi-quadrennial oscillation (QQO) because of its four year period. The robustness of this mechanism led astronomers to hypothesize that the same physical mechanism could be operating in the solar tachocline (Kumar, Talon & Zahn 1999, Kim & MacGregor 2001, Talon & Charbonnel 2005). This could then provide a handsome explana- tion for oscillations at the base of the convection zone, whether on 20 year timescales (as in the dynamo, Kumar, Talon & Zahn 1999) or 1 year timescales (as in the 1.3 year oscillation, Kim & MacGregor 2001). We have reviewed this mechanism here so comparisons between this process and the zonal flow oscillations seen in the radiative region of our model can be clearly made. 3. Numerical Model The numerical technique and model setup are identical to those in Rogers & Glatzmaier 2005, except here we impose the equatorial rotation profile as a function of radius as inferred from helioseismology (Ω(r) is specified to be 465nHz in the convection zone, 435nHz in the stable region and the tachocline is fit to an error function). We solve the Navier-Stokes equations with rotation in the anelastic approximation in 2D cylindrical geometry (r,φ). The curl of the momentum equation, i.e., vorticity equation, is: ∂ω ∂t + (~v · ~∇)ω = (2Ω(r) + ω)hρvr − 2vr ∂Ω(r) ∂r − g T r ∂T ∂θ − 1 ρT r ∂T ∂r ∂p ∂θ + ν∇2ω (3) The heat equation is: ∂T ∂t + (~v · ∇)T = −vr( ∂T ∂r ∂T ∂r − (γ − 1)T hρ) + (γ − 1)T hρvr+ γκ[∇2T + (hρ + hκ) ] + γκ[∇2T + (hρ + hκ) ∂T ∂r ] + Q cv (4) In these equations, ~v is the velocity, with radial, vr, and longitudinal, vθ, components. The vorticity is ~ω = ~∇ × ~v and is normal to the equatorial plane in this 2D geometry. The functions hρ = dlnρ/dr, hκ = dlnκ/dr, g (gravity), T (temperature), ρ (density) and γ (ratio of specific heats, cp/cv) are radially dependent and taken from the solar model. T is the temperature perturbation and p is the pressure perturbation, which like ω, are functions of r, θ and time (t). -- 6 -- In our model we specify the thermal diffusivity as that given by the solar model, mul- tiplied by a constant for numerical stability: κ = kapmult ∗ 3 16σT 3ρ2kcp (5) here kapmult is generally set to 105, σ is the Stefan-Boltzman constant and k is the opacity. The viscous diffusivity, ν = µ/ρ, is set so that the dynamic viscosity (µ) is constant. We calculate the pressure term in (3) using the longitudinal component of the momen- tum equation: 1 ρr ∂p ∂θ = − ∂vθ ∂t − (~v · ~∇~v)θ + ν[(∇2~v)θ − hρ 3r ∂vr ∂θ ]. (6) These equations are supplemented by the continuity equation in the anelastic approximation ∇ · ρ~v = 0 (7) which is satisfied by expressing ρ~v as the curl of a streamfunction. The model extends from .001R⊙ to 0.93R⊙. The subadiabaticity in the radiative region is given by the solar model and we specify the superadiabaticity in the convection zone to be 10−7. These equations are solved using a Fourier spectral transform method in the longitudinal (θ) direction and a finite difference scheme on a non-uniform grid in the radial (r) direction. Time advancing is done using the explicit Adams-Bashforth method for the nonlinear terms and an implicit Crank-Nicolson scheme for the linear terms. The boundaries are impermeable and stress-free. The inner boundary is isothermal and the outer boundary is held at a constant heat flux. This code is parallelized using message passing interface (MPI) and the resolution is 2048 longitudinal zones x 1500 radial zones, with 620 radial zones dedicated to the radiative region. In the region just below the convection zone the radial resolution is 170km. This model was evolved for 1 simulated year, requiring nearly six million 5 second timesteps. 4. Convection zone and Tachocline 4.1. Angular Velocity Variations Helioseismic observations indicate that, in the equatorial plane, the convection zone spins faster than the radiative interior. The maintenance of differential rotation within the equatorial plane of the convection zone is a 3D process involving the transport of angular momentum in both radius and latitude and the Coriolis forces resulting from axisymmetric -- 7 -- meridional circulation. Since our 2D geometry captures only the transport in radius, we can not expect to achieve a realistic differential rotation profile with the 2D model. Therefore, we impose the observed equatorial angular velocity as a function of radius in our model. As mentioned above, the background angular velocity is set to a constant 465nHz in the convection zone and to a constant 435nHz in the radiation zone; we prescribe a smooth fit between these two values through the thin tachocline shear layer. The resulting gravity wave spectrum and angular momentum transport is then investigated. The time series of angular velocity, Ω′(r, t), relative to the prescribed background an- gular velocity, Ω(r), for this model is shown in Figure 1 over one simulated year. As seen in this figure, angular velocity in the convection zone is initially prograde, relative to the prescribed Ω(r), in the lower part of the convection zone and retrograde motion in the upper part. However, after the initial period, angular velocity becomes very time dependent and is dominantly retrograde in the lower part of the convection zone and prograde in the upper part. This profile is expected, given the density stratification and rotation of the solar inte- rior (Glatzmaier et al. 2005). Intermittently prograde motion from the top of the convection zone will extend to the base of the convection zone and overshoot into the tachocline. The angular velocity variations produced by these motions vary in amplitude between +/- 15nHz, slightly larger than the amplitudes of the observed 1.3 year oscillation (Howe et al. 2001). Figure 1a clearly shows that the angular velocity of the tachocline mimics the behavior in the lower part of the convection zone; when the lower part is prograde, the tachocline is prograde and vice versa. This indicates that the behavior of the tachocline is dictated by convection zone dynamics, rather than by gravity waves (see below). In the overshoot region fluid motions are strongly nonlinear (Rogers & Glatzmaier 2005b). Figure 2 shows the ratio of the horizontal fluid velocity to the horizontal phase speed (the Froude number) for a typical frequency and wavenumber (20µHz, l=10); this provides a measure of the nonlinearity of waves. For linear waves ux/cx << 1. However, as is clearly seen in the figure this criterion does not hold just below the convection zone. While only one ratio is shown as a function of radius, this can vary by an order of magnitude depending on the choice of frequency and wavenumber.1 However, using reasonable values for frequency and wavenumber, the smallest value of this ratio just below the convection zone is 0.1, still far from meeting the linearity criterion above. Furthermore, it is clear in this figure that linearization may not be justified in the solar core either, as suggested by Press (1981). 1For larger (smaller) values of horizontal wave mode number this ratio is larger (smaller). Similarly, for larger (smaller) frequencies this ratio is smaller (larger). -- 8 -- 4.2. Reynolds and Viscous Stresses The mean zonal flow is determined by a balance between viscous and Reynolds stresses (as seen in equation (1)). In order to better understand the angular velocity profile in time (as shown in Figure 1) we examine the sources of that balance. Figure 3 shows the horizontally averaged convergence of Reynolds stress (first term on the right hand side of (1), but now in cylindrical coordinates), the viscous stress (second term on the right hand side of (1), in cylindrical coordinates) and the sum of these (the time rate of change of the mean zonal flow U ) in both the convection zone and overshoot region over five days. In the model's convection zone, the Reynolds stresses are much larger than the viscous stresses. The pattern of Reynolds stress convergence has a semi-periodic behavior with disturbances propagating up and down from a radius near 1/3dcz (dcz is the convection zone depth). When positive (i.e., prograde) Reynolds stress convergence moves upward from this radius a negative value moves downward and vice versa. This process reverses with a frequency similar to the convective turnover frequency, roughly 10 times the inertial wave frequency. This oscillatory behavior is observed in our simulations only when we include a radiative region beneath the convection zone; we do not see it when we impose an impermeable lower boundary on the convection zone. Note also that this radius of 1/3dcz is approximately where vθ for the large eddies changes sign and below which smaller, counter- rotating cells exist. Figure 4 shows a zoomed-in region of the convection zone, displaying the vorticity within a few convective cells. Beneath the large cells spanning most of the convection zone is counter-rotating fluid that is driven by the plumes penetrating slightly into the stable overshoot region. As a plume rebounds it is diverted laterally and counter- rotating eddies are generated. It remains to be seen how these counter-rotating cells affect tachocline dynamics in a 3D simulation. Now consider the overshoot region in our model. Regions with positive (negative) Reynolds stress convergence in the lower convection zone correspond to the same signed disturbances in the overshoot region. These disturbances have a slight tendency to prop- agate downward, i.e., away from their source in the convection zone. When viscous terms are added (which in the overshoot region have magnitudes similar to the Reynolds stress terms) disturbances clearly propagate away from the convection zone (Fig. 3). This motion away from the convection zone 2 is also seen in Figure 1. The semi-periodic oscillations observed in ∂U /∂t affects only the amplitude of U, but not the direction. That is, the semi-periodic oscillation seen ∂U /∂t is not realized in U because of a slight asymmetry in the driving amplitude of prograde and retrograde motions. If, as has been assumed in the 2However, even if diffusion were neglected the disturbances do not move upward as in the QBO. -- 9 -- past (Kumar, Talon & Zahn 1999, Kim & MacGregor), these motions were produced with equivalent amplitudes, the semi-periodic oscillation observed in ∂U /∂t would be reproduced in U, however, that is not the case here. Given the above results it is clear that overshooting plumes themselves transport angular momentum in a way which is distinct from the nonlinear interaction of low amplitude waves. The main properties of a wave driven oscillation (motion in time toward the source, timescales different than the forcing timescales, and a double peaked layer) are not seen here. Therefore, given the link between Reynolds stresses in the convection zone and those in the tachocline, it is clear that the overshooting plumes provide the main transfer of angular momentum to the tachocline. These plumes are buoyantly braked, radiatively damped and the nonlinearity allows for mode-mode transfer which affects the damping rate. Furthermore, the momentum transport by these plumes is time dependent and complex. Their effects need to be included when considering the dynamics of the tachocline and their influence should not be treated simply as an increased flux in linear waves. In fact, the linear approximation of waves is not justified in the overshoot region and tachocline. 5. The Deep Interior and Core A banded radial differential rotation profile is observed deep within the radiative interior (Figure 5); however, the amplitude of the angular velocity variation is extremely low (0.02 nHz) relative to what it is in the convection zone. This radial shear grows in time (Figure 6) due to wave-wave interactions which enforce shear. However, there is some indication that the growth is slowing and possibly reversing (Figure 6) due to increased viscous diffusion as angular velocity gradients increase. The banded differential rotation is due to the broad spectrum of frequencies generated and their continual deposition of angular momentum due to radiative diffusion as they propagate. Banded differential rotation in the sun's radiative interior has not been inferred from helioseismology, possibly because the amplitude is too low, or because these small amplitude disturbances are easily neutralized in 3D. In the core of our model the amplitude of the angular velocity variation is 2-3 orders of magnitude higher than in the bulk of the radiative interior. This indicates that angular momentum transport from waves to zonal flow is more efficient in the core, which could be due to a number of factors, including: wave breaking, critical layers (where the phase speed of the wave approaches the mean rotation rate) or inefficient reflection as the wave frequency approaches the Brunt-Vaisala frequency. Discerning which of these processes is dominant in this region is extremely difficult when a large spectrum of waves is continually being excited and is beyond the scope of this paper. -- 10 -- Maxima in the convergence of Reynolds stress deep within the radiative interior (but a significant distance below the overshoot region) propagate upward (Figure 7) as expected from the nonlinear interaction of low amplitude waves. In addition, significant interference is seen in the lower radiation zone, which likely comes from the interaction of inward prop- agating and outward reflected waves. While we have not observed a complete reversal of the angular velocity in the bulk of the radiative region, there is some evidence that one is underway (Figure 6)3. In the core the mean angular velocity switches from retrograde to prograde at nearly the same time as the tachocline switches from prograde to retrograde. This suggests that the angular velocity of the core is linked to the angular velocity of the tachocline and hence, of the convection zone. The link between tachocline angular velocity and that of the core due to selective filtering has been elucidated previously in Talon, Kumar & Zahn 2002. Here we make the link between the tachocline and the convection zone and propose that a self-consistent study of differential rotation in the sun must not treat the radiative and convective regions separately. 6. Discussion Despite some similarity between the convective-radiative interface in the sun and the Earth's tropopause, there are several obvious differences. In the sun, convection is constantly driving gravity waves everywhere below the overshoot region; whereas the generation of large- amplitude gravity waves in the Earth's atmosphere is intermittent in time and space. Solar gravity waves travel down into a converging region of increasing density (and therefore, have very little chance of breaking); whereas in the Earth waves travel up into an expanding region of decreasing density. Furthermore, previous numerical simulations of convective penetration in a stratified atmosphere demonstrate a remarkable difference between penetration into an overlying stable region and penetration into an underlying stable region (Hurlburt, Toomre & Massaguer 1986). Penetration into an underlying stable region is characterized by thin localized downflows; whereas penetration into an overlying stable region is characterized by larger scale broad upflows. This asymmetry allows descending plumes to travel farther into an underlying stable region than ascending motions travel into an overlying stable region. Therefore, it is likely that penetrative convection plays a more crucial role at the base of the solar convection zone than it does at the Earth's tropopause. These differences can have profound effects on the role of overshoot and, hence, on the scale, frequency and amplitude of the waves generated and, so, on the angular momentum transport by these waves. 3It appears that in the deep interior a QBO-like oscillation is possible -- 11 -- Numerical simulations (Wedi & Smolarkiewicz 2005) of the Plumb-McEwan laboratory experiment attempting to reproduce the QBO, have shown that the type and period of an oscillation in the differential rotation profile depend sensitively on the forcing. In particular, it is found that random forcing rarely produces a periodic oscillation. Given the turbulent nature of the sun, it is likely that the forcing is fairly random. For reasons stated above, it is unlikely that there is a QBO-like oscillation associated with the solar tachcocline. However, QBO-like oscillations may occur in stars with radiative envelopes because of the inefficiency of overshoot into an overlying stable region, and because of more similar geometry. 7. Conclusions We have presented self-consistent numerical simulations of convective overshoot and gravity wave generation and the angular momentum transport by these processes in a 2D model of the dynamics in the solar equatorial plane. We find that angular velocity variations in the tachocline are driven by angular momentum transported by overshooting plumes rather than nonlinear interaction of low amplitude waves. These overshooting plumes are strongly nonlinear disturbances, which can not be accurately represented as an increased flux of linear waves. We observe a semi-periodic oscillation in amplitude, but not in direction, of the mean flow in the tachocline because of an asymmetry in the driving of prograde and retrograde motions. Since we find that linear gravity waves are not dominant in the tachocline it is unlikely that they are responsible for the 1.3 year oscillation or the 11 year solar cycle. It is no surprise that overshooting motions play a dominant role in the tachocline and we expect these results will persist in three-dimensions. In the deep radiative interior the continual deposition of angular momentum by the nonlinear interaction of gravity waves produces a radially banded differential rotation. How- ever, it remains to be seen whether this pattern persists in 3D, considering it's very low amplitude. In the model's core, the amplitude of the differential rotation (i.e., angular ve- locity) is larger, about two orders of magnitude larger than that in the bulk of the radiative region and similar to the magnitude within the convection zone. We observe retrograde motion in the core reversing to prograde motion in step with the counter-reversal (prograde to retrograde) at the tachocline. When there is predominantly prograde flow at the base of the convection zone it selectively filters out prograde propagating gravity waves, allowing predominantly retrograde waves to propagate to the core where they deposit their (negative) angular momentum, and vice versa (Talon, Kumar & Zahn 2002). Like previous results our simulations therefore suggest that the angular velocity variations in the solar core are linked -- 12 -- to those in the tachocline. However, unlike previous results we show here that the variations in the tachocline are driven by convective overshooting, therefore, linking core rotation to convective motions. We thank K. MacGregor, J. Christensen-Dalsgaard, D. Gough, C. Jones, P. Garaud and M. Metchnik for helpful discussions and guidance. T.R. would like to thank the NPSC for a graduate student fellowship and the Institute of Astronomy, Cambridge University for a travel grant. Support has also been provided by the DOE SciDAC program DE-FC02- 01ER41176, the NASA Solar and Heliospheric Program SHP04-0022-0123 and the UCSC Institute of Geophysics and Planetary Physics. Computing resources were provided by NAS at NASA Ames and by the NSF MRI grant AST-0079757. Baldwin, M.P., Gray, L.J., Dunkerton, K.Hamilton, Haynes, P.H., Randel, W.J., Holton, REFERENCES J.R. et al. 2001, Rev. Geophysics, 39, 179 Eliassen, A. and Palm, E. 1961 Geofysiske Publ., 22, 1 Garcia-Lopez, R.J., Spruit, H. 1991, ApJ, 377, 268 latzmaier, G.A., Evonuk, M. and Rogers, T.M. 2005, Nature, under review Gough, D.O., McIntyre, M.E. 1998, Nature, 394, 755 Howe, R., Christensen-Dalsgaard, J., Hill, F., Komm, R.W., Larsen, R.M., Schou, J., Thompson, M.J., and Toomre, J. 2001, Science, 287, 2456 Holton, J.R., An Introduction to Dynamic Meteorology, 4th edition, Elsevier Academic Press, 1994 Hurlburt, N.E., Toomre, J. and Massaguer, J.M. 1986, ApJ, 311, 563 Kim, E., MacGregor, K.B. 2001, ApJ556, L117 Kumar, P., Quataert, E. 1997, ApJ, 475, L143 Kumar, P., Talon, S., Zahn, J.P. 1999, ApJ, 520, 859 Leovy,C.B., Friedson, A.J., Orton, G.S. 1991 Nature, 354, 380 Lindzen, R.S., Dynamics in Atmospheric Physics, Cambridge University Press, 1990 -- 13 -- Plumb, R.A., McEwan, A.D. 1978, J. Atmos. Sci, 35, 1827 Press, W.H. 1981, ApJ, 245, 286 Ringot, O. 1998, A&A, 335, L89 Rogers, T.M., Glatzmaier, G.A. 2005, MNRAS, in press Schatzman, E. 1993 A&A, 279, 431 Talon, S., Kumar, P., Zahn, J.P. 2002, ApJ, 574, 175 Talon, S., Charbonnel, C. 2005, A&A, 440, 981 Wedi, N.P., Smolarkiewicz, P.K., 2005, JAS, submitted This preprint was prepared with the AAS LATEX macros v5.2. -- 14 -- Fig. 1. -- Top: Angular velocity variations, relative to the prescribed solar profile, as a func- tion of time and radius. Red represents prograde motion, while blue represents retrograde motion. Black line represents the convective-radiative interface. Variations are shown in nHz. Bottom: Angular velocity variation as a function of time at the convective-radiative interface. Dotted line represents zero fluctuation about the mean. -- 15 -- Fig. 2. -- The ratio of fluid velocities to the phase velocity (for a typical frequency and wavenumber), also known as the Froude number, a measure of the nonlinearity of the waves. This figure shows that motions are clearly nonlinear, both in the tachocline and the core. -- 16 -- Fig. 3. -- Convergence of the Reynolds stress, Viscous stress and the sum, ∂U /∂t, over a five-day period in both the convection zone (top) and overshoot region (bottom). A semi- periodic oscillation is seen both in the convection zone and overshoot region. This oscillation causes an oscillation in the amplitude (but not direction) of the mean zonal flow seen in Fig. -- 17 -- Fig. 4. -- Zoom in of vorticity in the convection zone. Large cells span the bulk of the con- vection zone with counter-rotating cells beneath due to the deflection of descending plumes as they encounter the stiff radiative region. -- 18 -- Fig. 5. -- Angular Velocity variations in the deep interior. Banded radial differential rotation is seen. The deviations in the core are 2-3 orders of magnitude larger than those in the bulk of the radiative interior. There appears to be a link between deviations in the core and those in the tachocline (see Figure 1). -- 19 -- Fig. 6. -- Time evolution of angular velocity variations. Some indication of reversal is seen in the bulk of the radiative interior between 223 days and 348 days. In the core there was clearly a reversal, which Figure 5 tells us was around 130 days. -- 20 -- Fig. 7. -- Convergence of the Reynolds stress in the bulk of the radiative interior (blue represents negative, red positive values). Disturbances move upward in time as expected from the nonlinear interaction of low amplitude waves. Significant interference is seen at smaller radii, probably due to the interaction of inward propagating and reflected waves.
astro-ph/0311281
1
0311
2003-11-12T15:33:26
Timing properties and spectral states in Aquila X-1
[ "astro-ph" ]
We have analyzed five X-ray outbursts of the neutron-star soft X-ray transient Aql X-1 and investigated the timing properties of the source in correlation with its spectral states as defined by different positions in the color-color and hardness-intensity diagrams. The hard color and the source count rate serve as the distinguishing parameters giving rise to three spectral states: a low-intensity hard state, an intermediate state and a high-intensity soft state. These states are respectively identified with the extreme island, island and banana states that characterize the atoll sources. The large amount of data analyzed allowed us to perform for the first time a detailed timing analysis of the extreme island state. Differences in the aperiodic variability between the rise and the decay of the X-ray outbursts are found in this state: at the same place in the color-color diagram, during the rise the source exhibits more power at low frequencies (< 1 Hz), whereas during the decay the source is more variable at high frequencies (> 100 Hz). The very-low frequency noise that characterizes the banana-state power spectra below 1 Hz cannot be described in terms of a single power law but a two-component model is required. In two outbursts a new 6-10 Hz QPO has been discovered and tentatively identified with the normal/flaring branch-like oscillation observed only at the highest inferred mass accretion rates. We have compared the spectral and timing properties of Aql X-1 with those of other atoll and Z sources. Our results argue against a unification scheme for these two types of neutron-star X-ray binaries.
astro-ph
astro-ph
Timing properties and spectral states in Aquila X–1 P. Reig1 , S. van Straaten2 , M. van der Klis2 3 0 0 2 v o N 2 1 1 v 1 8 2 1 1 3 0 / h p - o r t s a : v i X r a ABSTRACT We have analyzed five X-ray outbursts of the neutron-star soft X-ray transient Aquila X – 1 and investigated the timing properties of the source in correlation with its spectral states as defined by different positions in the color-color and hardness-intensity diagrams. The hard color and the source count rate serve as the distinguishing parame- ters giving rise to three spectral states: a low-intensity hard state, an intermediate state and a high-intensity soft state. These states are respectively identified with the extreme island, island and banana states that characterize the atoll sources. The large amount of data analyzed allowed us to perform for the first time a detailed timing analysis of the extreme island state. Differences in the aperiodic variability between the rise and the decay of the X-ray outbursts are found in this state: at the same place in the color-color diagram, during the rise the source exhibits more power at low frequencies (< ∼ 1 Hz), whereas during the decay the source is more variable at high frequencies (>∼ 100 Hz). The very-low frequency noise that characterizes the banana-state power spectra below 1 Hz cannot be described in terms of a single power law but a two-component model is required. In two outbursts a new 6-10 Hz QPO has been discovered and tentatively identified with the normal/flaring branch-like oscillation observed only at the highest inferred mass accretion rates. We have compared the spectral and timing properties of Aquila X – 1 with those of other atoll and Z sources. Our results argue against a unification scheme for these two types of neutron-star X-ray binaries. Subject headings: accretion, accretion disks — stars: neutron — stars: (Aquila X–1) — X-rays: stars individual 1G.A.C.E, Instituto de Ciencias de los Materiales, University of Valencia, P.O. Box 22085, E - 46071 Paterna-Valencia, Spain 2Astronomical Institute “Anton Pannekoek”, University of Amsterdam and Center for High-Energy Astrophysics, Kruislaan 403, NL- 1098 SJ Amsterdam, the Netherlands 1 1. Introduction Aquila X – 1 belongs to the general group of sys- tems known as low-mass X-ray binaries (LMXB). These systems consist of a neutron star or a black hole orbiting around a late-type star (later than A). The X rays are the result of the accretion of material from the companion onto the compact star. Mass transfer is thought to occur via Roche lobe over- flow, and hence proceeds via an accretion disk. In Aquila X – 1, the compact star is a neutron star and the mass-donating companion is a V=21.6 K7V star, located at an estimated distance of 2.5 kpc (Cheva- lier et al. 1999). An orbital period of 18.95 hours has been suggested (Chevalier & Ilovaisky 1991; Welsh, Robinson & Young 2000). The high-energy radiation is characterized by 1.5–2 month long transient X-ray outbursts during which the X-ray luminosity can in- crease by more than three orders of magnitude. They are thought to be due to thermal instabilities in the accretion disk (e.g. van Paradijs 1996). Their re- currence time and duration vary but typical values are ∼ 200 days and 40–60 days, respectively (Simon 2002). The source also displays Type I bursts (Zhang, Yu & Zhang 1998) that last a few tens of seconds and are interpreted as runaway thermonuclear burning of matter on the surface of the neutron star. Various schemes have been proposed to categorize the LMXBs: X-ray spectral behavior as a function of intensity and the requirement of a low-energy black- body component in the X-ray spectra (Parsignault & Grindlay 1978; Naylor & Podsiadlowski 1993), cluster analysis of a large number of source characteristics (Ponman 1982), detailed X-ray spectral fits (White & Mason 1985), X-ray hardness-intensity and color- color diagrams (Schulz, Hasinger & Trumper 1989), aperiodic variability at very low frequencies (Reig, Papadakis & Kylafis 2003). Most relevant for the purpose of this paper is the classification scheme in terms of the rapid aperiodic variability and the pat- terns that these sources display in color-color dia- grams (Hasinger & van der Klis 1989). In this scheme LMXBs are divided into two different subclasses, Z and atoll sources. The three spectral branches that make up the Z shape are called horizontal branch, normal branch and flaring branch and the two structures that occur in atoll sources are known as the island and the ba- nana (van der Klis 1989). At the lowest count rates an extension of the island state is sometimes seen with a harder spectrum and stronger band-limited noise than the ”canonical” island state (van Straaten, van der Klis & M´endez 2003 and references therein). The term extreme island state has been used to designate such state (Prins & van der Klis 1997; Reig et al. 2000). The spectral and timing properties of atoll sources in this state are reminiscent of those seen in the low/hard state of black hole systems (van der Klis 1994a, 1994b; Berger & van der Klis 1998; Olive et al. 1998; Belloni, Psaltis & van der Klis 2002). Both Z and atoll sources move through these patterns contin- uously, that is, without jumping from one branch to another, although the source motion along the spec- tral tracks is much slower in atoll sources when they are in the island state than in the banana state and in Z sources in all states. The classification of Aquila X–1 in this scheme was investigated by Reig et al. (2000) — see also Cui et al. (1998) —, who studied the correlated X-ray timing and spectral variations of Aquila X – 1 and presented evidence for its classifi- cation as an atoll source, exhibiting all classic atoll source states. This scheme has been revisited by Muno, Remil- lard & Chakrabarty (2002) and Gierli´nski & Done (2002). They reported that three transient atoll sources which display a wide dynamic range in inten- sity (Fmax /Fmin >∼ 100; 4U 1608–52, 4U 1705–44 and Aquila X – 1 trace out three-branch patterns in the color-color diagram similar to those of Z sources, and suggested this may be a general feature. Their study was based on the spectral properties only. However, as pointed out by Hasinger & van der Klis (1989) it is difficult to make a clear distinction between Z and atoll sources on the basis of the motion through the color-color diagram alone. Analysis of the rapid aperiodic variability to study the noise components in different regions of the color-color diagram and the actual time scales of the motion of the source through the diagram are important to identify source type and state. In this work we have studied the spectral and tim- ing properties of Aquila X – 1 in order to investigate whether atoll and Z sources can be unified into one classification scheme and provide new insights into the poorly known extreme island states of atoll sources. We find that although the color-color diagram shows a branch whose topology is similar to the normal branch of Z sources, neither the timing properties nor the mo- tion in the color-color diagram of Aquila X – 1 agree with those of Z sources. 2 2. Observations The data were retrieved from the RXTE archive and comprise all observations of Aquila X – 1 available from 1997 February to 2002 May. Data taken during satellite slews and Earth occultation were removed. Likewise, all Type I bursts were excluded from our analysis. The observations contain five X-ray out- bursts. Although the duration and maximum inten- sity differ, the profile of the outbursts is very similar and is characterized by a fast rise and a slower decay. During intensity maximum the light curve is complex and multi-peaked. The fourth outburst showed two minor outbursts (the peak intensity was one order of magnitude lower than the main outburst) 70 days and 110 days after the main peak, respectively. The long-term light curve of the observations is shown in Fig. 1. 3. The color-color diagram Background subtracted light curves corresponding to the energy ranges 2.0–3.5 keV, 3.5–6.0 keV, 6.0– 9.7 keV and 9.7–16.0 keV were used to define the soft and hard colors as SC=3.5–6.0/2.0–3.5 and HC=9.7– 16.0/6.0–9.7, respectively. The color-color diagram (CD) of Aquila X – 1 was then constructed by plotting the hard color as function of the soft color (Fig. 2). During the time spanned by the observations the re- sponse of the detectors varied due to ageing. In ad- dition, gain changes are applied occasionally, mak- ing the channel boundaries for a given energy range change with time, and also slightly affecting the effec- tive areas of the detectors. Each gain change is the start of a new ”gain epoch”. We reduced the effect of the color shifts that results from these gain changes by linearly interpolating between the count rates in the two energy channels straddling the energy boundaries in each epoch. Each data point of Aquila X – 1 was normalized with the closest corresponding Crab point within the same gain epoch in order to mitigate the response change effects on the colors. The final CD was obtained by averaging the five (one for each de- tector) normalized CDs and rebinning into 256-s data points. We excluded data for which the resulting rel- ative errors were larger than 5%. The total number of data points excluded from the analysis was <∼ 10%. In order to recover the true values of the colors of Aquila X – 1 the soft color should be multiplied by 2.35 and the hard color by 0.56 (quoted values are av- erages of the Crab colors during the observations and 3 for all five PCUs). The variation of the Crab colors computed as the root-mean-square, i.e. (the standard deviation over the mean color) was 5.3% and 1.7% for the soft and hard colors, respectively — see also Fig. 1 in van Straaten et al. (2003). 3.1. Spectral states Aquila X – 1 can be found in two main states: a low-intensity hard state and a high-intensity soft state. In addition, a short-lived intermediate or tran- sition state is found displaying values of the hard color in between the two main states. The aperiodic vari- ability as a function of spectral hardness, i.e. position in the CD of Aquila X – 1 was previously studied by Reig et al. (2000). With only two of the five out- bursts analysed in that work, the CD consisted of a soft branch, the banana branch, and two isolated groups of points which were called extreme island and island states. The large amount of data now analyzed reveals a more structured CD, in which the extreme island state appears as a more stable and longer-lived branch than the island state. The high/soft state cor- responds to the classical banana state and the island state represents the transition between the two main states. Outbursts 1, 4 and 5 (O1, O4 and O5) contain points in the two main states while outbursts 2 and 3 (O2 and O3) provide points to the banana state only. Observations of O2 and O3 began and finished when the source was still at a relatively high level of emis- sion (> 1200 c/s) and did not experience large ampli- tude changes (Imax /Imin < 5). In contrast, outburst 1, 4 and 5 extended over a larger dynamic range in intensity (Imax /Imin > 50). None of the points of the two minor outbursts which followed O4 (which will be termed here as O4′) contributed to the banana state. Outbursts 1, 4, 4′ and 5 include points in the ex- treme island state. While O4 and O5 include points both during the rise and during the decay, O1 gives points during the decay only. The island state oc- curred only during the decay of O1 and O5. All five outbursts (except for O4′) have points in the banana state but only O3 provides banana state points during the rise. 3.2. Motion in the color-color diagram Figure 3 displays the motion of Aquila X – 1 in the CD for outbursts 4 and 5. Data points are ∼1-day averages. Figure 4 shows the evolution of the colors with time for those outbursts that include spectral transitions (O1, O4, O5). The 2–16 keV intensity just before the transitions is also given. All quoted intensities in this section are background subtracted and correspond to 5 PCUs in the energy range 2–16 keV. There is a strong correlation between the po- sition of the source in the CD and its intensity. At the onset of the outbursts the data points distribute in the softest part of the extreme island state. The count rate is <∼ 100 c/s. As the intensity increases the source moves toward the right along the extreme island branch. The hard color (HC) slightly decreases on average. In about 6 days the count rate increases by one order of magnitude and the soft color (SC) increases by ≈+0.3. Then the source seems to jump to the banana state, recovering the initial values of the soft color. The hard color decreases by ≈ −0.5. The 2–16 keV intensity is ∼ 3 or 4 times that of the extreme island state prior to the spectral transition, namely 4000–6000 c/s. No intermediate points be- tween the two states are observed during the rise of the outbursts. This can be attributed to the fast rise and the lack of good sampling of the data. Indeed, the observational gaps between the last point of the extreme island branch and the first one of the banana branch were 2.6 and 3.1 days for O4 and O5, respec- tively. As the intensity continues to increase the source becomes harder, i.e., it moves to the right along the banana, with approximately constant hard color. The peak of the outburst is characterized by flaring activ- ity with erratic changes in count rate. The count rate at the peak is 50–100 times the minimum detected count rate. Despite this irregular behavior the corre- lation between the intensity and the colors is main- tained in the sense that the lower intensity points in the flares display a lower soft color as it is illustrated in Fig. 5. In other words, during the flare maxima the source lies at the right end of the banana branch, and in between flares it moves to the left. Thus the flaring variability in the light curve translates into the CD in a back and forth motion which approximately extends over the right half part of the banana state. This motion is quite fast. As an example, in one of the flares in O5 the soft color changed by 0.08 in about 2.4 hr. In another flare by 0.06 in 0.55 hr. This flar- ing behavior is seen in all outbursts but O2 shows it exclusively (Fig. 5). Such behavior stops once the source gets to half way the decay of the outburst, at which point the source shifts to lower soft colors (to the left along the banana branch). As the outburst declines the source moves back along the banana branch, abandoning it when the count rate becomes lower than ∼ 800 c/s. The transi- tion to the hard state occurs at much lower soft color than when it entered the banana, although not nece- sarily at the lowest soft color. The banana state cov- ers values of the soft color between 0.87 (O5) and 1.16 (O3). The transition to the hard state takes place at SC ≈ 0.92–0.93 (Fig. 3). The count rate at which the transition between the two main states takes place is higher when the source is in the rise of the outburst. Hard to soft transitions occur when the count rate is well above ∼ 1000 c/s. Soft to hard transitions when the count rate is well below ∼ 1000 c/s as it can be seen in Fig. 4. Rather than jumping directly to the hard branch the source remains for a short time (≤ 0.1 days) in an intermediate state, the island state. Note that such a short island state episode would usually have been missed due to data gaps in the rise. The count rate in the island state is ∼ 200–400 c/s. At even lower count rates the source finds itself back in the extreme island branch, moving toward the left as the intensity decreases, to eventually become too weak to measure the colors sufficiently well. The source re-enters this state during the decay when the count rate goes below ∼ 200 c/s. The re-entry point occurs at a lower soft color (SC ≈ 1.05–1.10) then when the source left the extreme island state (SC ≈ 1.3). As the speed of mo- tion along the extreme island state is approximately constant, the time during which the source can be found in this state is shorter during the decay of the outburst than during its rise. A difference in the po- sition of the source in the CD depending on whether the source intensity increases or decreases was already recognized by Reig et al. (2000). The main result that should be stressed is that transitions between states do not occur at the same point of the spectral branches. During the rise of the X-ray outburst the source occupies the hardest parts of the extreme island state before the spectral transition to the banana state. During the decay it tends to occupy the softest part of the banana branch before moving into the island state. However, the points of departure and arrival are different. In the CD, this translates into some sort of rectangular track (Fig. 3) along which Aquila X – 1 moves clockwise as the count rate first increases and then decreases. In addition to Fig. 4, Tables 1 and 2 also illustrate 4 the time scale for the motion through the diagram. Table 1 and Table 2 give upper limits on the dura- tion of the source in each state and the time scales of the spectral transitions, respectively. Note that the observational gap for the spectral transitions from the extreme island state to the banana state is ap- proximately two times longer than viceversa. Given the speed of motion of the soft color along the ex- treme island branch (roughly 0.05 day−1 ) a three-day gap might imply a change in soft color of about 0.15. Thus, it is possible that the actual entry point into the banana branch is at lower values of the soft color. In this case the motion of Aquila X – 1 in the CD would be similar to that of 4U 1705-44 (Barret & Olive 2002), in which both the transition from (to) the extreme island state to (from) the banana state occur at the same point of the banana branch. The motion in the CD then would then resemble an in- verted triangle rather than a rectangle. 4. Timing analysis In order to investigate the aperiodic variability of Aquila X – 1 we obtained power spectra by dividing the 2–60 keV PCA data of each observation into 256-s segments and calculated the Fourier power spectrum of each segment up to a Nyquist frequency of 2048 Hz. The high-frequency end (1500–2048 Hz) of the power spectra was used to determine the underlying Pois- son noise (approximated by a constant power level), which was subtracted before performing the spectral fitting. Our aim is to investigate the aperiodic vari- ability in correlation with the spectral states, which in turn, correlate with the source count rate. Thus we divided the color-color plane into various regions and obtained a mean power spectrum for each region. The mean power spectrum was the result of averag- ing all the power spectra corresponding to the points enclosed in each color region. See Table 3 to find the average values of the colors, boundaries and intensity of each region. Prior to this, we investigated whether power spectra from the same CD region taken at dif- ferent outbursts were consistent with having the same properties. We found that while this was true for the banana state, some differences existed in the extreme island state depending on whether the source was in the rise or the decay of the X-ray outburst. The extreme island state is populated with points from the rising part of O4 and O5, the decaying part of O1, O4 and O5 and from the extended extreme is- land state of O4′ . Consequently, the extreme island state was first separated into points pertaining to the rise, to the decay or to O4′ . Then each one of these groups was further divided into three subgroups ac- cording to the value of the soft color. Note that in the case of the rise and decay regions this division implies a division in count rate as well as in hard color. On average, the extreme island sates EISrise1 and EISde- cay2, (see Table 3) have lower intensity and are softer than EISrise2 and EISdecay1, respectively. In turn, EISrise2 contains softer points than EISrise3. The is- land state points (intermediate state) defined another region. The timing properties of the banana state in atoll sources have been seen to vary smoothly as the soft color, or equivalently, the count rate increases, that is, from the lower (left) to the upper (right) ba- nana branch. Therefore, the banana branch was split into five regions (BS1 to BS5) in increasing order of the soft color. Each region approximately contains the same amount of data. There is no unanimity in the literature about the names of the power spectral noise components nor about the mathematical functions used to fit those components. In this work we have followed the ter- minology of Belloni et al. (2002) and fitted the power spectra using Lorentzian functions only. Also, for dis- playing our power spectra we have used the power times frequency representation ν ×P (ν ) vs ν , in which the product of the power density in the rms normal- ization and the Fourier frequency is plotted as a func- tion of the Fourier frequency (Belloni et al. 1997). As fit parameters we use the frequency νmax at which the Lorentzian attains its maximum value in such repre- sentation, and the Q value, which gives a measure of the coherence of the Lorentzian. In terms of the half width at half maximum or ∆, these quanties are re- lated by νmax = pν 2 0 + ∆2 and Q = ν0/2∆, where ν0 is the central frequency of the Lorentzian. Between three and five Lorentzians were needed in order to get a good fit. If during the fitting procedure a Lorentzian component resulted with negative Q value then we set it to zero. It should be noted that the naming scheme of Bel- loni et al. (2002) was applied, in the case of low- luminosity neutron stars, to systems displaying the island state only. Thus, a complete description of some power spectra, especially those in the banana state, in terms of the noise components given in that work is not always possible and extra or new compo- nents are needed in order to obtain acceptable fits. 5 The extensions to the Belloni et al. naming scheme we use are those of van Straaten et al. (2003). In the terminology of Belloni et al. (2002), the band-limited noise of atoll sources in the (extreme) island state consists of three Lorentzians namely, Lb that accounts for the power spectrum at low frequen- cies (<∼1 Hz), Lℓ and Lu that fit its high-frequency end. Lℓ may show up as a broad bump at frequencies 10–25 Hz (normally in the island state) or as a nar- row QPO at frequencies >∼ 100 Hz (normally in the banana state). In fact, at frequencies above >∼ 100 Hz Lℓ and Lu represent the kHz QPOs seen in low-mass X-ray binaries. However, note that this identifica- tion of Lℓ with both a broad peak and the lower kHz QPO is only tentative and the broad bump at 10-25 Hz might alternatively represent a separate compo- nent (van Straaten et al. 2002). The intermediate frequency range (1–100 Hz) is described by one broad Lorentzian or ”hump”, Lh and one narrow Lorentzian LLF referred to as the low-frequency QPO. Not de- tected by Belloni et al. (2002) but also present in some atoll sources (van Straaten et al. 2002) is the so-called hectoHerz QPO, LhH z , a broad feature with characteristic frequency at ∼ 100 Hz. 4.1. Spectral states: aperiodic variability Table 4 lists the results of the fits and Fig. 6 de- picts the power spectra and fit functions for different regions of the CD (see Table 3 to identify the position in the CD). 4.1.1. The extreme island state (EIS) The extreme island state power spectrum contains, in order of increasing frequency, the low-frequency band-limited noise, Lb , the broad peak low-frequency noise Lh , the broad version of Lℓ and one more com- ponent which could be either the hHz QPO or a kHz QPO. Based on the rms/frequency correlations of the noise components seen in other atoll sources, van Straaten et al. (2002, 2003) favored the upper kHz interpretation. If Lℓ is also interpreted as the lower kHz QPO then the intervals EISrise2, EISrise3 and ISO4′ 2 (see Fig.6) would represent the first detection of the two kHz QPO in Aquila X – 1. Alternatively, the fourth component could be the hHz QPO. Note that the narrow ∼1160 Hz peak in EISO4′ 2 is not sta- tistically significant. An F-test reveals that the prob- ability that the improvement of the fit (in terms of a lower χ2 ) by the addition of an extra Lorentzian 6 happen by chance is larger than 30%. The character- istic frequencies of the noise components increase and their rms values decrease slightly as the count rate increases, i.e., the source moves toward the right in the extreme island state. Figure 7 displays the characteristic frequency of the Lh and Lℓ components as a function of the band- limited noise component Lb for Aquila X – 1, the atoll sources 4U 1608–52 (van Straaten et al. 2003), 4U 1728–34 and 4U 0614+09 (van Straaten et al. 2002) and the Z source GX5–1 (Jonker et al. 2002). In the case of GX5–1 te horizontal branch oscillation is plotted instead of Lh . Note also that the broad-band noise in the Z source was not fitted with a Lorentzian but with a cutoff power law, P (ν ) ∝ ν−α e−ν /νcut . By differentiating νP (ν ) and equating the resulting func- tion to zero the maximum frequency of the Z band- limited noise, νmax = (1 − α)νcut , was derived. Fig- ure 7 seems to confirm that i) the 1–20 Hz broad fea- ture Lh observed in atoll sources at low count rates (island states) has the same origin as the horizon- tal branch oscillations (HBO) observed in Z sources (Psaltis, Belloni & van der Klis 1999, van Straaten et al. 2003), ii) the broad bump at frequencies 10–25 Hz is associated with the lower kHz QPO. The ex- treme island state data in Aquila X – 1 sample the lowest characteristic frequencies so far detected. One interesting result from the timing analysis of the extreme island state is the differences in the ape- riodic variability of the source during the rise and decay of the outbursts. Figure 8 shows three power spectra of Aquila X – 1 and an enlargement of the CD displaying the extreme island state. All three power spectra are associated with the same value of the colors, i.e, they occupy the same position in the CD (marked in Fig. 8). However, they correspond to different parts of the X-ray outburst: rise (diamonds), decay (stars) and when the source found itself in the extended island state of O4′ (crosses). During the rise the power is concentrated at low frequencies whereas during the decay, the source exhibits more power at high frequencies During the extended island state of outburst O4′ there is roughly equal power at all fre- quencies. The differences in the characteristic fre- quencies of the noise components are much less pro- nounced and consistent with one onother whenever the same component appears. 4.1.2. The banana state The low-frequency end (<∼ 1 Hz) of the banana state power spectra is dominated by the very-low fre- quency noise (VLFN). One single power law is nor- mally used to fit this component. However, the power spectrum of Aquila X – 1 below 1 Hz cannot be de- scribed by one component only. In order to fit the VLFN we used two zero-centered Lorentzians that were called low VLFN (LLV LF N ) and high VLFN (LHV LF N ) (Schnerr et al. 2003). The characteris- tic frequencies of these components are found to be independent of the position of the source in the CD: νmax ∼ 0.01 Hz and ∼0.03 Hz, respectively. However, while the fractional rms variability of lower frequency VLFN roughly remains constant at ∼3.4%, the higher frequency VLFN becomes stronger — increasing from ∼0.9% in the lower banana to ∼2% in the upper ba- nana — as the count rate increases. Above 1 Hz the banana state power spectra contain one broad Lorentzian, describing the band-limited noise and one QPO (two in BS1). The character- istic frequency and width of the broad band-limited noise component decrease as the count rate increases (i.e. as the source moves toward the right). This component has also been seen in 4U 1608–52 (van Straaten et al. 2003), 4U 0614+09 and 4U 1728– 34 (van Straaten et al. 2002). However, the lack of data precludes a more detailed comparison. Only 4U 1728–34, with just two measurements, seems to have the same behavior as Aquila X – 1 i.e., the fre- quency of this noise component decreases along the banana. Following van Straaten et al. (2003) we will refer to this component as Lb2 . Next to, some- times on top of, the band-limited noise in the banana state there is a relatively narrow peak, whose char- acteristic frequency and width do not correlate with any other parameter. This narrow QPO has a fre- quency in between that of the hHz Lorentzian and Lh . van Straaten et al. (2003) — see also di Salvo et al. (2001) and van Straaten et al. (2002) — ar- gued that it is the band-limited noise component Lb which transforms into a narrow QPO above certain frequency (∼ 20 Hz in 4U 1608–52). The appearance of Lb2 and this transformation of Lb occurs coinciden- tally. In Aquila X – 1, although the narrow version of Lb appears systematically in all banana-state power spectra it is always below 3σ . The detection of a second kHz QPO in Aquila X – 1 is only marginal (van der Klis 2000). Based on the values of the fitting parameters and the relation between the QPO frequency and the X-ray colors, M´endez, van der Klis & Ford (2001) concluded that whenever a kHz QPO is detected in Aquila X – 1 it is always the lower kHz QPO. The fractional rms and characteristic frequency of the second QPO in BS1 are consistent with being the lower kHz QPO (Fig. 1 of M´endez, van der Klis & Ford 2001). Also, the char- acteristic frequencies of the Lℓ and Lb in BS1 agree with the correlation of Fig. 7 if an extrapolation to higher frequencies of the Aquila X – 1 points is done. 4.1.3. The island state (IS) The analysis of the transition state is hampered by poor statistics given the relatively low number of points. One single Lorentzian with peak at ≈ 20 Hz and Q value of 0.15 accounts for the entire 0.01– 100 Hz frequency band. This noise component is identified as Lb . A second narrower Lorentzian with νmax ≈ 710 Hz and Q ≈ 3 fits the noise at higher frequencies. Its fractional rms, ∼ 17%, is too high to agree with the lower kHz QPO. In fact, the value of the frequency and rms are similar to what it is seen in 4U 1608–52 and 4U 1728–34 for the upper kHz QPO (Fig. 1 of M´endez, van der Klis & Ford 2001). Nev- ertheless, given the relatively noisy spectrum at high frequencies such identification should be taken with care. 4.2. The normal/flaring branch-like oscilla- tions A QPO with frequency 7–14 Hz has been found in at least two atoll sources, XTE J1806–246 (Wi- jnands & van der Klis 1999; Revnivtsev, Borozdin & Emelyanov 1999) and 4U 1820–30 (Wijnands et al. 1999). These QPOs are detected at the high- est inferred mass accretion rates only, are short-lived (a few hundreds of seconds) and are localized in a very narrow region of the CD, namely the tip of the upper banana. They are reminiscent of those ob- served in Z sources in the normal and flaring branches, when the source is accreting near the Eddington limit. Thus the detection of similar QPO in atoll sources at significantly lower accretion rates questions the current models that explain NBO in terms of near- Eddington accretion (Fortner, Lamb & Miller 1989). We searched for NBO-like QPO in Aquila X – 1 and tentatively found two of them at ∼ 3σ signifi- cance, at frequencies 10.3±0.5 Hz in O3 and 5.8±0.2 7 Hz in O5 (Fig. 9). They have similar rms ampli- tude (≈ 0.7%) and Q-values (≈ 3). The average values of the colors and intensity associated with these QPOs are SC=1.11, HC=0.52, IX=5045 count s−1 and SC=1.06, HC=0.51, IX=6870 count s−1 , re- spectively. For the sake of comparison with other sources we also give the value of the X-ray luminosity: 8.5 × 1036 erg s−1 for outburst 3 and 8.1 × 1036 erg s−1 for outburst 5. As a confirmation that they occur at the highest count rates, typical luminosities in the softest parts of the extreme island and banana states are 3 × 1034 and 2 × 1036 erg s−1 , respectively. The values of the luminosity are for a distance of 2.5 kpc and the energy range 2–16 keV. In the other three outbursts the detection is only marginal. The low statistical significance of these QPOs (<∼2.5) can be ascribed to their extremely short life. 5. Discussion We have analysed all the available RXTE data of Aquila X – 1 between 1997 February and 2002 May and investigated its spectral and timing properties in correlation. This work has been motivated by recent reports that show that if large data sets are used, the atoll sources displaying the largest ampitude varia- tions in intensity show three-branch patterns in the color-color diagram (CD), reminiscent of Z sources (Muno, Remillard & Chakrabarty 2002; Gierli´nski & Done 2002). Consequently, they conclude that the distinction between atoll and Z sources might be no longer justified. The extended island state, island state and banana state of atoll sources would then correspond to the horizontal branch (HB), normal branch (NB) and flaring branch (FB), respectively. The results of these studies were based on the spectral properties (i.e. color-color diagram) only. In order to be able to distinguish between the different types of low-mass X-ray binaries the rapid aperiodic variabil- ity associated with the various spectral states needs to be addressed as well (Hasinger & van der Klis 1989), and that is what we have attempted in this paper. We first make a comparison of the timing properties of Aquila X – 1 with those of other atoll sources and then discuss whether ot not it can be considered as a Z source. 5.1. Comparison with atoll sources 4U 1608–52 and Aquila X – 1 are the atoll sources that display the largest dynamic range in X-ray in- tensity (Imax /Iquiesence > 1000). Thus a comparison between these two sources is of particular interest. 4U 1608–52 has been extensively studied by van Straaten et al. (2003). By comparing the shape of the power spectra (see Fig. 7 in van Straaten et al. 2003 and our Fig. 6) and the results of the power spectral fits (their Table 2 and 3 and our Table 4) one can find dis- tinct similarities between the two sources: the power spectra of intervals A, B and C in 4U 1608–52 resem- ble those of the extreme island state of Aquila X – 1, while interval D of 4U 1608–52 is reminiscent of the island state in Aquila X – 1. Likewise, the shape and number of Lorentzians appearing in the power spectra of the extreme island state of Aquila X – 1 are simi- lar to those of the three low-luminosity X-ray bursters studied by Belloni et al. (2002). It is worth noting the relevant role of the hard color in connection with the timing properties of the extreme island state. In Aquila X – 1 the charac- teristic frequencies of the timing features in the ex- treme island state increase from left to right. This is opposite to what it is seen in 4U 1608-52, where frequencies increase from right to left. However, in both sources the increase in frequencies occurs in cor- relation with an overall decrease in hard color. We therefore suggest that in the extreme island state it is the hard color, rather than the intensity, which is the main determining factor of the timing properties. We note, that black hole candidates are likewise char- acterized by power spectral states whose occurrence seems to primarily depend on spectral hardness and which are relatively insensitive to intensity, and which are associated with two-dimensional motion through the color-intensity plane rather than motion along a one-dimensional track (Homan et al. 2001). It may be that in the island and extreme island states, the states in which atoll sources are most similar to black holes (e.g., van der Klis 1994a; Belloni, Psaltis & van der Klis 2002), possibly because in these states the inner disk radius is furthest from the neutron star surface, this 2-D picture better describes neutron star behavior than the 1-D description appropriate to the banana state and to Z sources. Leaving aside the kHz QPOs, the overall shape of the banana-state power spectra of Aquila X – 1 present similarities with those of other atoll sources, namely, a strong red-noise component, the very-low frequency noise (VLFN) below ∼ 1 Hz and a broad and a narrow components describing the band-limited noise (BLN) in the frequency range 1–100 Hz. Com- 8 pared to 4U 0614+09, 4U 1608-52 and 4U 1728– 34, Aquila X – 1 exbibits the strongest VLFN but the weakest BLN. While the VLFN is normally de- scribed by either a power law with index between 1.5-2.0 or a zero centered Lorentzian in the other atoll sources, it requires the use of two model com- ponents in Aquila X – 1, similar to the case of GX 13+1 (Schnerr et al. 2003). The broad band-limited noise component Lb2 in Aquila X – 1 and 4U 1608- 52, or equivalently the zero centred Lorentzian in 4U 1728–34 and 4U 0614+09 have similar rms values. In contrast, Lb , which is clearly detected in these other atoll sources, is not significant in Aquila X – 1 in the banana state. 5.2. Is Aquila X – 1 a Z source? Gierli´nski & Done (2002) suggested that the name atoll source is no longer appropiate as these type of low-mass X-ray binaries also display a three-branch pattern in the color-color diagram when large data sets of observations are used, reminiscent of the Z pattern for which Z sources are named. This ef- fect is more pronounced in systems exhibiting large flux variations (Muno et al. 2002), like Aquila X – 1. The CD of Aquila X – 1 (Fig. 2) does indeed show a branch (the island state) that can be inter- preted to connect the two main branches like in a Z shape. In addition, the correlation between the char- acteristic frequency of the 1–20 Hz broad Lorentzian Lh and the band-limited noise component (Fig. 7) at low count rates, suggesting a common origin with the horizontal branch oscillations and LFN in Z sources and the discovery of 7–14 Hz QPO at the highest X-ray flux (Fig. 9), reminiscent of the Z-source nor- mal/flaring branch oscillations, strengthen the simi- larities between Aquila X – 1 and Z sources. However, there are a number of systematic differ- ences that clearly puts Aquila X – 1 outside the group of Z sources. First, there is the motion in the CD. In the extreme island state as the count rate increases the soft color increases and the hard color decreases. Then the source makes a rapid transition to the ba- nana state and proceeds to harder colors (right) as the count rate increases further. However, after the peak of the outburst, when the intensity decreases the source does not follow exactly the reverse path. As shown in Fig. 3, Aquila X – 1 enters the banana state (extreme island state) at a higher (lower) value of the soft color than when it leaves it. The source moves clokcwise following a rectangle track. It should be 9 pointed out that the entrance into the banana state at such high soft color is not caused by the time reso- lution used (∼ 1 day) in Fig. 3. A 256-s bin shows the same effect. However, we cannot rule out an observa- tional origin since the time gap between the transition from the extreme island state to the banana state is ≈ 5 days for outburst 5 and ≈ 2.6 for outburst 4 (Table 2). If this is the case and the source entered the banana state from the left then the pattern that it would trace out in the CD would be that of an in- verted triangle, much like 4U 1705–44 (Barret & Olive 2002). In contrast, the re-entrance into the extreme island state at a low soft color is probably not cause by the lack of data given the relatively short gap between the island state and the extreme island state (just 1.6 days in outburst 5). In either case this behavior is not that of Z sources. In Z sources, the NB is not a fast transition between the other two branches, and the source follows the same path whether it is going up or down along the NB. In this respect, Aquila X – 1 is like the other two atoll sources that display high- amplitude flux variations. More detailed looks into the motion of 4U 1608–52 and 4U 1705–44 have re- vealed different topologies from the classical Z-shaped track. The motion of 4U 1608–52 resembles the greek letter ǫ (van Straaten et al. 2003) and 4U 1705–44 describes an inverted triangle (Barret & Olive 2002). As noted above, these various topologies may sim- ply be different aspects of 2-D motion through the color-intensity plane in which it is hard color that de- termines the timing properties. Second, the amplitude of the X-ray luminosity change over the Z track is also different. In Aquila X – 1 the X-ray luminosity throughout the CD, i.e., from the left of the banana state to the right of the extreme island state is about three orders of magnitude. In Z sources, luminosity changes are typically less than a factor of 2 (di Salvo et al. 2000, 2001, 2002). Third, another difference is the velocity and time spent by the source in each spectral state. Jonker et al. (2002) calculated that the Z source GX 5–1 spent most of the time in the normal branch. In con- trast, Aquila X – 1 spends the smallest percentage of the the total observing time in the transition state (the analogue to the normal branch would Aquila X – 1 be a Z source). Wijnands et al. (1997) found that Cyg X–2 moves through the Z most slowly on the HB, faster on the NB and fastest on the FB. During a typical X-ray outburst, i.e., when Aquila X – 1 is X-ray active, it spends most of its time in the banana five X-ray outbursts covering a period of more than five years. Three spectral states show up in the color- color diagram of Aquila X – 1: a low/hard state, identified with an extreme island state, a high/soft state corresponding to the classic banana state and an intermediate state associated with the island state of atoll sources. We have found that the hard color plays a crucial role in determining the timing proper- ties of the extreme island state. Although the overall distribution of these states in the color-color diagram may resemble a Z, neither the motion in the CD, nor the typical time scales through the different branches of the CD, nor the range of X-ray luminosities, nor the timing properties at very low count rates (the ex- istence of the extreme island state) are compatible with the classical Z-source behavior. PR is a researcher of the programme Ram´on y Cajal funded by the University of Valencia and the Spanish Ministery of Science and Technology. PR also wants to thank the University of Crete for pro- viding in part the resources needed to carry out this work. This research has made use of data ob- tained through the High Energy Astrophysics Science Archive Research Center Online Service, provided by the NASA/Goddard Space Flight Center. branch (the analogue to the flaring branch). Finally, the properties of the aperiodic variability of Aquila X – 1 also differ from those in Z sources, especially at very low count rates. We can establish the following differences: – The very existence of the extreme island state. Although the characteristic frequencies in Aquila X – 1 increase along the extreme island state as the count rate increases as Z sources do in the HB, the latter never reach the low frequencies that are seen in atoll sources in the extreme island state, even when at the left end of the HB. Like- wise, the rms amplitude of the noise compo- nents in the extreme island state is significantly higher than in a Z source HB. The total rms amplitude of those components in Aquila X – 1 amounts to ∼ 30%, while typically in Z sources, it does not go above ∼ 10%. The extreme island-state power spectra are typical of atoll sources. – The peaked noise components, such as Lh and the upper kHz QPO are narrow QPOs on the HB in Z sources, with typical values of the Q parameter above 3 for the upper kHz QPO, whereas they are considerably broader peaked noise components in the extreme island state in Aquila X – 1(Q < 0.5). – In Aquila X – 1, the normal/flaring branch- like oscillation does not occur during the ”nor- mal/flaring branch”, the island or the i.e. lower banana branches but in the upper banana branch, just like in other atoll sources. – The strength of the kHz QPOs decreases as the Z source moves along the Z track from HB to NB. Typically, the kHz QPOs become unde- tectable by the time the source reaches the mid- dle of the NB. In contrast, Aquila X – 1 presents kHz QPOs in the island and banana states. – The time scales of the aperiodic variability tend to decrease (frequencies increase) as the Z source moves along the FB, which is the opposite be- havior to what is is observed in Aquila X – 1 in the banana branch. 6. Conclusions We have investigated the timing and spectral prop- erties of the soft X-ray transient Aquila X – 1 during 10 REFERENCES Nowak, M.A., 2000, MNRAS, 318, 361 Barret, D, & Olive, J.F., 2002, ApJ, 576, 391 Belloni, T., van der Klis, M., Lewin, W. H. G., van Paradijs, J., Dotani, T., Mitsuda, K., & Miyamoto, S., 1997, A&A, 322, 857 Belloni, T., Psaltis, D., van der Klis, M., 2002, ApJ, 572, 392 Berger, M., & van der Klis, M.,1998, A&A, 340, 143 Chevalier, C. & Ilovaisky, S. A. 1991, A&A, 251, L11 Chevalier, C., Ilovaisky, S.A., Leisy, P., Patat, F., 1999, A&A, 347, L51 Cui, W., Barret, D., Zhang, S.N., Chen, W., Boirin, L., Swank, J., 1998, ApJ, 502, L49 Di Salvo, T., Stella, L., Robba, N. R., van der Klis, M., Burderi, L., Israel, G. L., Homan, J., Cam- pana, S., Frontera, F., Parmar, A. N., 2000, ApJ, 544, L119 di Salvo, T., M´endez, M., van der Klis, M., Ford, E., Robba, N. R., 2001, ApJ, 546, 1107 di Salvo, T., Farinelli, R., Burderi, L., Frontera, F., Kuulkers, E., Masetti, N., Robba, N. R., Stella, L., van der Klis, M., 2002, A&A, 386, 535 Fortner, B., Lamb, F. K., Miller, G. S., 1989, Nature, 342, 775 Gierlinski, M., Done, C., 2002, MNRAS, 331, L47 Hasinger, G. & van der Klis, M. 1989, A&A, 225, 79 Homan, J., Wijnands, R, van der Klis, M. Belloni, T., van Paradijs, J., Klein-Wolt, M. Fender, R. M´endez, M., 2001, ApJS, 132, 377 Jonker, P. G., van der Klis, M., Homan, J., M´endez, M., Lewin, W.H.G., Wijnands, R., Zhang, W., 2002, MNRAS, 333, 665 M´endez, M. van der Klis, M., Ford, E.C., 2001, ApJ, 561, 1016 Muno, M.P., Remillard, R.A., Chakrabarty, D., 2002, ApJ, 568, L35 Naylor, T., Podsiadlowski, Ph, 1993, MNRAS, 262, 929 Olive, J. F., Barret, D., Boirin, L., Grindlay, J. E., Swank, J. H., & Smale, A. P. 1998, A&A, 333, 942 Parsignault, D.R., Grindlay, J.E., 1978, ApJ, 225, 970 Ponman, T, 1982, MNRAS, 200, 351 Prins, S., van der KLis, M., 1997, A&A, 319, 498 Psaltis, D., Belloni, T., van der Klis, M., 1999, ApJ, 520, 262 Reig, P., M´endez, M., van der Klis, M., Ford, E. C., 2000, ApJ, 530, 916 Reig, P., Papadakis, I., Kylafis, N., 2003, A&A, 398, 1103 Revnivtsev, M., Borozdin, K., Emelyanov, A., 1999, A&A, 344, L25 Schnerr, R.S., Reerink, T., van der Klis, M., Homan, J., M´endez, M., Fender, R.P., & Kuulkers, E., astro-ph/0305161 Schulz, N. S., Hasinger, G. & Trumper, J. 1989, A&A, 225, 48 Simon, V., 2002, A&A, 381, 151 van der Klis, M., 1989, ARA&A, 27, 517 van der Klis, M, 1994a, A&A, 283, 469 van der Klis, M, 1994b, ApJS, 92, 511 van der Klis, M., 2000, ARA&A, 38, 717 van Paradijs, J., 1996, ApJ, 464, L139 van Straaten, S., van der Klis, M., di Salvo, T., Bel- loni, T., 2002, ApJ, 568, 912 van Straaten, S., van der Klis, M., M´endez, M., 2003, ApJ, in press, astro-ph/0307041 Welsh, W.F., Robinson, E.L., Young, P., 2000, AJ, 120, 943 White, N.E., Mason, K.O., 1985, SSRv, 40, 167 Wijnands, R., van der Klis, M., 1997, ApJ, 482, L65 Wijnands, R., van der Klis, M., 1999, ApJ, 514, 939 Wijnands, R., van der Klis, M., Rijkhorst, E. 1999, ApJ, 512, L39 11 Zhang, S. N., Yu, W. & Zhang, W., 1998, ApJ, 494, L71 This 2-column preprint was prepared with the AAS LATEX macros v4.0. 12 Table 1 Duration and intensity (2–16 keV, 5 PCUs) of the spectral states for each of the five X-ray outbursts States Duration days EIS BS IS EIS BS IS EIS BS IS EIS BS IS EIS BS IS EIS BS IS –/0.14 15.1 0.10 – 30.96 – – 38.95 – 6.02/2.03 18.05 – 63 – – 6.54/1.33 41.53 0.01 Imin c/s Outburst 1 –/56 583 200 Outburst 2 – 1213 – Outburst 3 – 1227 – Outburst 4 110/70 368 – Outburst 4′ 77 – – Outburst 5 87/128 810 458 Imax c/s –/63 3057 220 – 2694 – – 6175 – Imean c/s –/60 2164 209 – 1778 – – 4656 – 1295/156 4835 – 650/127 2330 – 530 – – 340 – – 1444/255 8555 463 542/203 4925 460 EIS: extreme island, BS: banana, IS: island The left (right) values correspond to the rise (decay) of the outburst Table 2 Duration in days of the spectral transitions. These values should be considered as upper limits as they include observational gaps Transition Outburst 1 Outburst 4 Outburst 5 4.94 2.63 – EIS−→BS 2.64 1.06 4.83 BS−→EIS 0.98 – 1.89 BS−→IS IS−→EIS 2.84 – 1.64 13 Table 3 Boundaries and properties of the color regions on which the timing analysis was performed States (SC, HC )1 (SC, HC )2 (SC, HC )3 EISrise1 EISrise2 EISrise3 EISdecay1 EISdecay2 EISO4′ 1 EISO4′ 2 IS BS1 BS2 BS3 BS4 BS5 (0.89,1.08) (1.14,1.06) (1.19,1.04) (1.01,1.02) (0.92,1.11) (0.96,1.04) (1.08,1.04) (0.97,0.71) (0.83,0.47) (0.93,0.41) (0.99,0.40) (1.04,0.41) (1.01,0.42) (1.12,1.08) (1.21,1.07) (1.34,1.05) (1.11,1.03) (1.11,1.14) (1.24,1.06) (1.07,1.06) (1.11,0.89) (0.93,0.41) (0.99,0.40) (1.04,0.41) (1.10,0.42) (1.18,0.50) (0.89,1.25) (1.14,1.15) (1.23,1.10) (1.00,1.12) (0.88,1.43) (0.96,1.26) (0.96,1.26) (0.82,0.97) (0.89,0.59) (0.95,0.57) (1.01,0.57) (1.02,0.57) (1.05,0.61) (SC, HC )4 Mean Mean Mean Count Ratea (c/s) SC HC 185 1.14 1.01 925 1.09 1.18 1063 1.07 1.25 183 1.07 1.05 76 1.24 0.98 1.05 1.15 188 356 1.14 1.14 255 0.90 0.98 1275 0.50 0.91 2086 0.49 0.97 3507 0.49 1.02 5248 0.50 1.07 1.12 0.51 5315 (1.11,1.26) (1.26,1.14) (1.29,1.11) (1.11,1.23) (1.06,1.45) (1.10,1.26) (1.10,1.26) (0.95,1.05) (0.95,0.57) (1.01,0.57) (1.02,0.57) (1.05,0.61) (1.09,0.68) Number of power spectra 54 34 23 37 40 152 367 28 313 368 368 381 250 χ2 (dof ) 122/92 105/84 95/87 87/92 106/92 106/87 114/84 99/91 193/87 139/89 100/88 89/88 99/88 aBackground subtracted count rate for the five PCU and bandwidth 2–16 keV 14 Table 4 Power spectral parameters. Errors are 1σ (∆χ2 = 1) Statea LLV LF N EISrise1 EISrise2 EISrise3 EISdecay1 EISdecay2 EISO4′ 1 EISO4′ 2 IS BS1 BS2 BS3 BS4 BS5 EISrise1 EISrise2 EISrise3 EISdecay1 EISdecay2 EISO4′ 1 EISO4′ 2 IS BS1 BS2 BS3 BS4 BS5 EISrise1 EISrise2 EISrise3 EISdecay1 EISdecay2 EISO4′ 1 EISO4′ 2 IS BS1 BS2 BS3 BS4 BS5 - - - - - - - - 0.0130±0.0007 0.0144±0.0009 0.0160±0.0009 0.0133±0.0007 0.0103±0.0010 - - - - - - - - 3.19±0.06 3.44±0.05 3.49±0.05 3.33±0.06 3.29±0.11 - - - - - - - - 0b 0b 0b 0b 0b - - - - - - - - 0.36±0.11 0.20±0.02 0.25±0.01 0.27±0.01 0.25±0.01 LHV LF N Lb2 Lb Characteristic frequency νmax (Hz) - 0.10±0.01 - 0.44±0.02 - 0.61±0.01 2.3±0.3 - - 0.47±0.05 - 0.130±0.006 - 0.284±0.009 - 20±2 42±2 62±5 73±5 45±5 24±3 76±3 16±1 73±4 65±6 13±1 Integrated fractional rmsc - 14.8±0.7 13.6±0.3 - - 11.9±0.4 - 9.5±0.6 - 11.0±0.4 - 16.0±0.4 - 14.9±0.2 15.7±0.5 - 5.0±0.2 <1.3 3.4±0.1 <1.2 2.3±0.1 <1.4 1.79±0.06 <1.1 <1.0 1.55±0.06 Q value 0b 0b 0b 0b 0b 0.14±0.04 0.03±0.02 0.11±0.09 0b 0.16±0.08 0.23±0.09 0.29±0.07 0.41±0.08 - - - - - - - - 5b 3.5b 2.9±1.2 2.4±1.1 2.6±1.7 - - - - - - - - 0.89±0.07 1.48±0.04 1.76±0.03 1.97±0.02 2.10±0.02 - - - - - - - - 0b 0b 0b 0b 0b Lh Lℓ LhH z ,Lu 0.59±0.07 2.62±0.06 5.4±0.2 17±3 - 1.11±0.07 2.17±0.04 - - - - - - 16.1±0.6 15.8±0.6 18.2±0.5 13.3±0.7 - 15.9±1.0 14.9±0.5 - - - - - - 0b 0.42±0.05 0.16±0.05 0b - 0.22±0.09 0.33±0.04 - - - - - - 12.1±1.0 21.6±0.9 32±2 - - 10.8±0.8 17.9±0.6 - 861±6 - - - - 19.2±0.5 16.1±0.6 10.0±1.0 - - 17.9±1.0 17.3±0.6 - 5.4±0.2 - - - - 0b 0.19±0.07 0.49±0.15 - - 0.15±0.13 0.16±0.06 - 7±1 - - - - - 233±23 273±24 1160b 880±250 180b 223±17 700±40 - - - - - - 11.4±0.4 12.2±0.8 19±2 38.5±0.5 17.8±1.9 14.0±0.8 17±2 - - - - - - 0.48±0.15 0.43±0.16 0b 0b 0b 0.25±0.13 1.9±0.9 - - - - - aRefers to the regions defined in the color-color diagram (see Table 3) bFixed cUpper limits are at 95% confidence level 15 Fig. 1.— Light curve of the entire set of observa- tions showing the five X-ray outbursts. Different sym- bols represent different spectral states: extreme island (circles), island (stars) and banana (dots). 16 Fig. 2.— Color-color diagram of Aquila X – 1. The soft and hard colors are relative to those of the Crab. Each point in the color-color diagram represents 256 s. No error bars are given but all the points shown in this plot have relative errors smaller than 5%. Dif- ferent symbols represent different outbursts as indi- cated. The lines separate the banana-state regions (from BS1 in the left to BS5 in the right). 17 Fig. 3.— Track followed by the source in the color- color diagram during outbursts 4 and 5. Each data point is a one day average. 18 Fig. 4.— Evolution of the hard color showing the transitions between the different spectral states: ex- treme island (circles), island (stars) and banana (dots). Time refers to the onset of each outburst. 19 Fig. 5.— Light curve of the second X-ray outburst (upper panel) and the corresponding color-color di- agram. The terms high and low refer to the high- intensity and low-intensity points of the flares. 20 Fig. 6.— Power spectra and best fits for different positions in the color-color diagram (see Table 3) The lines represent the different noise components as follows: for the island states, Lb (dashed), Lh (dotted), Lℓ (dot-dashed) and Lu (dot-dot-dot-dashed); for the banana state LLV LF N (dashed), LHV LF N (dotted), Lb (dot-dashed), Lb 2 (dot-dot-dot-dashed). The second dashed line in BS1 corresond to Lℓ . 21 Fig. 7.— Characteristic frequency of the Lh and Lℓ noise components of the atoll sources 4U 1608–52, 4U 1728–34 and 4U 0614+09 (dots) and the HBO of the Z sources GX 5-1 (stars) as a function of the characteristic frequency of the band-limited noise (Lb for the atoll sources and HBO for GX5–1). Aquila X – 1 observations have been represented by open circles. 22 Fig. 8.— Comparison of the power spectra for three different parts of the X-ray outburst. All three power spectra correspond to the same region of the CD. An enlarged view of the island state marking the color region is also plotted. 23 Fig. 9.— N/FBO-like feature in the power spectra of Aquila X – 1 during the peak of outbursts 3 and 5. 24
0709.3175
1
0709
2007-09-20T10:25:17
A deeply embedded young protoplanetary disk around L1489 IRS observed by the submillimeter array
[ "astro-ph" ]
Circumstellar disks are expected to form early in the process that leads to the formation of a young star, during the collapse of the dense molecular cloud core. It is currently not well understood at what stage of the collapse the disk is formed or how it subsequently evolves. We aim to identify whether an embedded Keplerian protoplanetary disk resides in the L1489 IRS system. Given the amount of envelope material still present, such a disk would respresent a very young example of a protoplanetary disk. Using the Submillimeter Array (SMA) we have observed the HCO$^+$ $J=$ 3--2 line with a resolution of about 1$''$. At this resolution a protoplanetary disk with a radius of a few hundred AUs should be detectable, if present. Radiative transfer tools are used to model the emission from both continuum and line data. We find that these data are consistent with theoretical models of a collapsing envelope and Keplerian circumstellar disk. Models reproducing both the SED and the interferometric continuum observations reveal that the disk is inclined by 40$^\circ$ which is significantly different to the surrounding envelope (74$^\circ$). This misalignment of the angular momentum axes may be caused by a gradient within the angular momentum in the parental cloud or if L1489 IRS is a binary system rather than just a single star. In the latter case, future observations looking for variability at sub-arcsecond scales may be able to constrain these dynamical variations directly. However, if stars form from turbulent cores, the accreting material will not have a constant angular momentum axis (although the average is well defined and conserved) in which case it is more likely to have a misalignment of the angular momentum axes of the disk and the envelope.
astro-ph
astro-ph
Astronomy & Astrophysics manuscript no. ms July 16, 2021 c(cid:13) ESO 2021 7 0 0 2 p e S 0 2 ] h p - o r t s a [ 1 v 5 7 1 3 . 9 0 7 0 : v i X r a A deeply embedded young protoplanetary disk around L1489 IRS observed by the Submillimeter Array C. Brinch1, A. Crapsi1, J. K. Jørgensen2,3, M. R. Hogerheijde1, and T. Hill1 1 Leiden Observatory, P.O. Box 9513, 2300 RA Leiden, The Netherlands e-mail: [email protected] 2 Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Mail Stop 42, Cambridge, MA 02138, USA 3 Argelander-Institut fur Astromomie, Universitat Bonn, Auf dem Hugel 71, 53121 Bonn, Germany ABSTRACT Context. Circumstellar disks are expected to form early in the process that leads to the formation of a young star, during the collapse of the dense molecular cloud core. It is currently not well understood at what stage of the collapse the disk is formed or how it subsequently evolves. Aims. We aim to identify whether an embedded Keplerian protoplanetary disk resides in the L1489 IRS system. Given the amount of envelope material still present, such a disk would respresent a very young example of a protoplanetary disk. Methods. Using the Submillimeter Array (SMA) we have observed the HCO+ J = 3 -- 2 line with a resolution of about 1′′. At this resolution a protoplanetary disk with a radius of a few hundred AUs should be detectable, if present. Radiative transfer tools are used to model the emission from both continuum and line data. Results. We find that these data are consistent with theoretical models of a collapsing envelope and Keplerian circumstellar disk. Models reproducing both the SED and the interferometric continuum observations reveal that the disk is inclined by 40◦ which is significantly different to the surrounding envelope (74◦). Conclusions. This misalignment of the angular momentum axes may be caused by a gradient within the angular momentum in the parental cloud or if L1489 IRS is a binary system rather than just a single star. In the latter case, future observations looking for variability at sub-arcsecond scales may be able to constrain these dynamical variations directly. However, if stars form from turbulent cores, the accreting material will not have a constant angular momentum axis (although the average is well defined and conserved) in which case it is more likely to have a misalignment of the angular momentum axes of the disk and the envelope. Key words. Stars: formation -- circumstellar matter -- ISM: individual objects: L1489 IRS -- Submillimeter 1. Introduction Circumstellar disks constitute an integral part in the for- mation of low-mass protostars, being a direct result of the collapse of rotating molecular cloud cores and the likely birthplace of future planetary systems. When a molecu- lar cloud core collapses to form a low-mass star, its initial angular momentum causes material to pile up on scales of ∼ 100 AU in a circumstellar disk (Cassen & Moosman 1981; Terebey et al. 1984; Adams et al. 1987, 1988). Due to accretion processes as the young stellar object (YSO) evolves, the disk grows in size while the envelope dissi- pates (e.g., Basu 1998; Yorke & Bodenheimer 1999). In the more evolved stages of young stellar objects such disks can be imaged directly at optical and near-infrared wave- lengths (e.g., Burrows et al. 1996; Padgett et al. 1999) and their chemical and dynamical properties can be de- rived from comparison to submillimeter spectroscopic ob- servations (e.g., Thi et al. 2001). Still, this latter method is often not unique, but requires a priori assumptions about the underlying disk structure. In the earlier em- bedded stages this method is even more complex, be- cause the disk is still deeply embedded in cloud ma- terial so that any signature of rotation in the disk it- self, for example, is smeared out by emission that orig- inates in the still collapsing envelope. By going to higher resolution using interferometric observations, as well as observing high density gas tracers, more reliable detec- tions of protoplanetary disks can be made as direct imag- ing is approached (Rodr´ıguez et al. 1998; Qi et al. 2004; Jørgensen et al. 2005b). In this paper we present such high-angular resolution submillimeter wavelength obser- vations from the Submillimeter Array (SMA) of the em- bedded YSO L1489 IRS. We show how these observations place strong constraints on its dynamical structure and evolutionary status through detailed modeling. 2 C. Brinch et al.: A deeply embedded young protoplanetary disk around L1489 IRS L1489 IRS (IRAS 04016+2610) is an intriguing pro- tostar in the Taurus star forming region (d = 140 pc) classified as a Class I YSO according to the classification scheme of Lada & Wilking (1984). It is still embedded in a large amount of envelope material in which a significant amount of both infall and rotation has been observed on scales ranging from a few tenths of an AU out to several thousands of AUs (Hogerheijde 2001; Boogert et al. 2002). This large degree of rotation makes this source an inter- esting case study for the evolution of angular momentum during the formation of low-mass stars, potentially linking the embedded (Class 0 and I) and revealed (Class II and III) stages. In a recent study by Brinch et al. (2007) (referred to as Paper I), a model of L1489 IRS was presented based on data from a large single-dish molecular line sur- vey (Jørgensen et al. 2004). This study was motivated by the intriguing peculiarities seen in L1489 IRS, such as the unusual large size and shape of the circumstellar material. The aim was to constrain the structure of its larger-scale infalling envelope and to place it in the right context of the canonical picture of low-mass star formation (see for example several reviews in Reipurth et al. 2007). The model presented in Paper I describes a flattened envelope with an inspiraling velocity field, parameterized by the stellar mass and the (constant) angle of the field lines with respect to the azimuthal direction. A spherical temperature profile was adopted and the model did not explicitly contain a disk. The mass of the envelope is 0.09 M⊙, adopted from Jørgensen et al. (2002). The use of such a "global" model is sufficient when working with single- dish data, where the emission is dominated by the emission from the large-scale envelope. With this description it was possible to accurately reproduce all the observed single- dish lines. What the study of Paper I could not address, is whether a rotationally dominated disk is present on scales of the order of hundreds of AUs, although the amount of rotation which is inferred by the single-dish observations certainly suggests that a disk should have formed. In ad- dition, two specific puzzles remain about the structure of L1489 IRS inferred on basis of the model when compared to other studies in the literature. First, the best fit was ob- tained with a central mass of 1.35 M⊙, which is a very high value given that the luminosity of the star has been deter- mined to be only 3.7 L⊙ (Kenyon et al. 1993). Second, it was found that the best fit inclination of the system was 74◦. This is in agreement with the result from Hogerheijde (2001) where they showed that in order to reproduce the observed aspect ratio the inclination cannot be less than 60◦. However, in a recent study, Eisner et al. (2005), mod- eled the spectral energy distribution (SED) of L1489 IRS and demonstrated that this required a significantly differ- ent systemic inclination of only 36◦. In this paper, we try to address these issues through arc second scale interferometric observations of the dense gas tracer HCO+ J = 3− 2 from the Submillimeter Array. These observations provide information on the gas dynam- ics on scales ∼100 AU and reveal an embedded proto- planetary disk. We show how careful modeling of the full SED from near-infrared through millimeter wavelengths can place strong constraints on the geometry of such a disk. The outline of this paper is as follows: Sect. 2 describes the details of the observations and data reduction while Sect. 3 presents the data. Sect. 4 introduces our model in which we have now included a disk and we also show how this model can fit all available observations including our new SMA data. Finally, in Sect. 5 and 6 we discuss the implications of our model and summarize our results. 2. Observations and data reduction Observations were carried out at the Submillimeter Array (SMA)1 located on Mauna Kea, Hawaii. Our data set consists of two separate tracks of measurements in dif- ferent array configurations. The first track was obtained on December 11, 2005. This track was done in the com- pact configuration resulting in a spatial resolution of ∼ 2.5′′ with projected baselines ranging between 10 and 62 kλ. A second track was obtained on November 28, 2006, in the extended configuration with the resolution of about 1′′ and baselines between 18 and 200 kλ. The resolution of the two configurations corresponds to linear sizes of 350 AU and 140 AU respectively. The synthesized beam size of the combined track using uniform visibility weighting is 0.9′′× 0.7′′ with a position angle of 78◦. In both tracks the receiver was tuned to HCO+ J = 3 -- 2 at 267.56 GHz. We used a correlator configuration with high spectral resolution across the line, providing a chan- nel width of 0.2 kms−1 over 0.104 GHz. The remainder of the 2 GHz bandwidth of the SMA correlator was used to measure the continuum. No other lines in this band are expected to contaminate the continuum. We did not en- counter any technical problems during observations and the weather conditions were excellent during both tracks. For the compact configuration track, τ225 was 0.06 and for the extended track τ225 was at 0.08. Mars was used as flux calibrator in the first track, and Uranus for the second track. The quasars 3c454.3, 3c273, and 3c279 were used for passband calibration for both tracks, while the complex gains were calibrated using the two quasars 3c111 and 3c84 located within 18 degrees from L1489 IRS. The calibrators were measured every 20 min- utes throughout both tracks. Their fluxes have been de- termined to be 3.1 and 2.6 Jy for 3c111 and 3c84 in the compact track, and 2.3 and 2.2 Jy respectively in the ex- tended track. For the gain calibration, a time smoothing scale of 0.7 hours was used which ensures that the large scale variations in the phase during the track are corrected for. The data do not show any significant small scale varia- tions (i.e., rapid fluctuations in the phases or amplitudes). 1 The Submillimeter Array is a joint project between the Smithsonian Astrophysical Observatory and the Academia Sinica Institute of Astronomy and Astrophysics and is funded by the Smithsonian Institution and the Academia Sinica C. Brinch et al.: A deeply embedded young protoplanetary disk around L1489 IRS 3 Table 1. Summary of the SMA observations. Source α, δ(2000)a Frequency Integrated line intensity Peak line intensity Noise level (rms) Continuum flux Noise level (rms) Synthesized beam size (Uniform weighting) L1489 IRS 04:04:42.85, +26:18:56.3 267.55762 GHz (1.12 mm) 41.85 Jy beam−1 km s−1 2.04 K 0.35 Jy beam−1 36.0±2.3 mJy 3.7 mJy beam−1 0.9′′ × 0.7′′ a Coordinates are given at the position where the continuum emission peaks. The quasars appear as point sources even at the longest baselines, which means that phase decorrelation due to atmospheric turbulence is negligible. The signal-to-noise of the calibrators is >50 per integration. The data were reduced using the MIR software pack- age (Qi 2005)2. Due to the excellent weather conditions, the data quality is very high and the data reduction proce- dure went smooth and unproblematic. All post-processing of the data, including combining the two visibility sets was done using the MIRIAD package (Sault et al. 1995)3. Relevant numbers are presented in Table 1. In this paper we also make use of continuum measure- ments of L1489 IRS found in the literature ranging from the near-infrared (Kenyon et al. 1993; Eisner et al. 2005; Padgett et al. 1999; Park & Kenyon 2002; Whitney et al. 1997; Myers et al. 2005) to (sub)millimeter wavelengths (Hogerheijde & Sandell 2000; Moriarty-Schieven et al. 1994; Motte & Andr´e 2001; Hogerheijde et al. 1997; Ohashi et al. 1996; Saito et al. 2001; Lucas et al. 2000). 1987; Kessler-Silacci et al. 3. Results The two tracks of observations provide us with two sets of (u, v) -- points which, when combined and Fourier trans- formed, samples the image (x, y) plane well from scales of 1 to 30′′ (140 to 4200 AU). Emission on spatial scales greater than this is filtered out by the interferometer due to its finite shortest spacing. This filtering is accounted for when comparing models to the observations. The measured continuum emission is shown as an im- age in Fig. 1. Previous attempts to measure the contin- uum at 1.1 mm with the BIMA interferometer were not successful (Hogerheijde 2001). The SMA however, reveals a complex and detailed, slightly elongated, structure. Two reconstructed images of the HCO+ emission are shown in Fig. 2 using the natural and uniform weight- ing schemes, optimizing the signal-to-noise and angular resolution, respectively. In this figure, the zero moment map is plotted as solid contour lines and the first moment 2 Available at http://cfa-www.harvard.edu/∼cqi/mircook.html. 3 Available at http://bima.astro.umd.edu/miriad/miriad.html Fig. 1. Reconstructed image of the continuum emission at the highest resolution (0.9′′× 0.7′′). The contours are linearly spaced with 1σ starting at 2σ and dashed lines are negative contour levels. is shown as shaded contours. It is clear that both im- ages reveal an elongated, flat structure. In the uniformly weighted image there is a large amount of asymmetry in the structure, which is less prominent in the naturally weighted image. The velocity contours in the lower panel of Fig. 2 are seen to be closed around a point which is offset with some 3′′ from the peak of the emission. There is also clearly a gradient in the velocity field along the major axis of the object. This feature was previously reported by Hogerheijde (2001) using interferometric observations of HCO+ J = 1 -- 0, and low S/N HCO+ J = 3 -- 2 obser- vations, from the BIMA and OVRO arrays. The gradient in the velocity field coincides almost perfectly with the long axis of the structure. When compared to the BIMA HCO+ J = 3 -- 2 observations, the SMA data give 5 -- 10 times better resolution. Furthermore, the BIMA data had to be self-calibrated using the HCO+ J = 1 -- 0 image as a model and thus the resulting image was somewhat depen- dent of the structure of the lower excitation emission. The image we obtain from the SMA data is of considerably better quality. In Fig. 3, the zero moment emission contours have been superposed on the near-infrared scattered light im- age taken by the Hubble Space Telescope (Padgett et al. 1999). The figure shows that many of the details in the SMA image coincide with features seen in the scattered light image. The spectral map in Fig. 4 shows single spectra at po- sitions offset by 2′′ from each other. Each position shown here is the spectrum contained within one synthesized beam. The center most spectrum coincides with the peak of the emission in Fig. 2. Here is seen a very broad, dou- ble peaked line. Moving outward from the center, the lines become single peaked and offset in velocity space with re- 4 C. Brinch et al.: A deeply embedded young protoplanetary disk around L1489 IRS Fig. 3. HCO+ emission (1σ) contours superposed on the near-infrared scattered light image (Padgett et al. 1999). ) c e s c r a ( δ ∆ 4 2 0 −2 −4 4 2 0 ∆α (arcsec) −2 −4 Fig. 4. Spectral map of L1489 IRS. The spectra are evenly spaced with 2′′. The bandwidth in each panel is 20 kms−1. The scale on the y-axis goes from 0 to 2 Jybm−1. where the protoplanetary disk is expected to be present (see Section 1). Furthermore, the HCO+ J = 3 -- 2 transi- tion traces H2 densities of 106 cm−3 or more (Schoier et al. 2005), which are expected in the inner envelope and disk. The model of Paper I was not explicitly made to mimic a disk, but rather to describe the morphology of the im- ages on large scales. Using this model, which works well on a global scale, we can explore how well it fits the SMA spectra when extrapolated to scales unconstrained by the single-dish observations. Fig. 2. Zero and first moment plots of the HCO+ J = 3 -- 2 emission toward L1489 IRS. Contours are linearly spaced by 2σ with a clip level of 2σ. σ equals 1.1 and 2.0 Jy bm−1 km s−1 for the natural and uniform weighting schemes respectively. Negative contours appear as dashed lines. spect to the systemic velocity. Perpendicular to the long axis of the structure, the line intensity falls off quickly. The lines shown in Fig. 4 are reconstructed using the nat- ural weighting scheme. For the remainder of this paper we will use the uniformly weighted maps where the resolution is optimal. Fig. 5 shows the emission in a position-velocity dia- gram, where the image cube has been sliced along the major axis to produce the intensity distribution along the velocity axis. The PV -- diagram shows almost no emission in the second and fourth quadrant which is a strong indica- tor of rotation. The small amount of low velocity emission seen close to the center in the fourth quadrant may be accounted for by infalling gas. 4. Analysis With the resolution provided by the SMA it is possible to probe the central parts of L1489 IRS on scales of ∼100 AU, C. Brinch et al.: A deeply embedded young protoplanetary disk around L1489 IRS 5 Fig. 6. This figure shows three spectra from the SMA observations with model spectra superposed. The two off- positions are chosen in the direction along the long axis of the object. The offsets are chosen to be a resolution element. Three models are also shown: The dashed line shows the unmodified model from Paper I, the full line is the model from Paper I, but inclined at 40◦, and the dotted line shows a model where HCO+ is absent on scales smaller than 200 AU. To check that this emission does in fact originate from scales less than 200 AU, a model where HCO+ is com- pletely absent within a radius of 200 AU is also shown in Fig. 6 with the dotted line. In this model the wide wings disappear and only a very narrow line is left. The two off- positions, which lie outside of the radius of 200 AU, are not affected minimally introducing this cavity. Fig. 6 also shows a spectrum which is made from the Paper I model, but inclined at 40◦ (full line). This is a considerably better fit than the other two models (again the two off-positions are little affected). While the obser- vations of the larger scale structure (Paper I, Hogerheijde (2001)) demonstrates that the inclination of the flattened, collapsing envelope is ∼74◦, the SMA observations sug- gest that a change in the inclination occurs on scales of 100 -- 200 AU, reflecting a dynamically different component there. On the other hand, since the model of Paper I was tuned to match the single-dish observation on scales of ∼1000 AU or more and did not explicitly take the disk into account it is not unexpected that it does not fully reproduce the SMA observations. 4.1. Introducing a disk model To reproduce also the interferometric observations we have modified the model from Paper I on scales corresponding to the innermost envelope and disk: The improvements consists largely of two things, namely a different parame- terization of the density distribution and the explicit in- clusion of a disk. The description used in Paper I has a discontinuity for small values of r and θ: again, this is not a problem when working with the large scale emission, but becomes a problem when interfaced with a disk where a contin- uous transition from envelope to disk is preferred. The Fig. 5. Position-Velocity (PV) diagram. This plot shows that the emission on the scales measured by the SMA is entirely dominated by rotation. The black curves show the Keplerian velocity, calculated from the dynamic mass (1.35 M⊙) obtained in Paper I. The curves have been in- clination corrected with 74◦ for the full lines and 40◦ for the dashed lines. Contours start at 2σ and increase with 1σ = 0.12 Jy beam−1. Fig. 6 compares the SMA observations to the predic- tions of this model (dashed line). For doing this compari- son, the model is imaged using the (u, v)-spacings from the observations, so that directly comparable spectra are ob- tained. While the spectra away from the center are fairly well reproduced in terms of line width, the center position is too wide, i.e., the model produces velocities that are too extreme compared to the measured velocities. 6 C. Brinch et al.: A deeply embedded young protoplanetary disk around L1489 IRS new parameterization we use for the envelope is adopted from Ulrich (1976). In this representation the envelope density is given by, Rrot(cid:19)−1.5(cid:18)1 + ρenv(r, θ) = ρ0(cid:18) r ×(cid:18) cos θ 2 cos θ0 cos θ cos θ0(cid:19)−1/2 cos 2θ0(cid:19)−1 , + Rrot r (1) 90 30 is the solution of where θ0 the parabolic motion of an infalling particle given by r/Rrot(cos θ0 − cos θ)/(cos θ0 sin 2θ0) = 1, Rrot = 150 AU is the centrifu- gal radius of the envelope, and ρ0 is the density on the equatorial plane and at the centrifugal radius. Since the outer radius, total mass, aspect ratio, and peak density are held fixed, the new parameterization is qualitatively similar to the Plummer-like profile used in Paper I. The numerical value of Rrot is chosen so that the maximum deviation in density in the (r, θ)-plane only reaches 20% in the most dense parts (outside of the disk cavity) and has an average deviation of less than 5%. 7 0 1 108 109 106 In addition to this envelope, a generic disk is also in- troduced in the model. The density of the disk is given by, Fig. 7. A slice through the model of L1489 IRS: The top panels show the temperature on different scales and the bottom panels show the corresponding density profiles. ρdisk(r, θ) = Σ0 (r/R0)−1 √2π H0 exp(− 1 2(cid:20) r cos θ H0 (cid:21)2) , (2) where θ is the angle from the axis of symmetry, R0 is the disk outer radius and H0 is the scale height of the disk. The model thus has four free parameters: disk radius, pressure scale height, mass (taken as a fraction of the total mass of the circumstellar material), and inclination. In order to maximize the mid-infrared flux, we use a flat disk, i.e., with no increase in scale height with radius and fix the scale height to 0.25. The outer radius is fixed to 200 AU, the distance from the center of the emission in Fig. 2 to the position of the closed velocity contour, thereby effec- tively reducing the number of free parameters to two. This outer disk radius is not a strongly constrained parameter since neither the SED nor the model spectra are influ- enced much by changes in this value. It is only possible to rule out the extreme cases, where the disk either gets so small that it is no longer influencing the SED or where it gets so big that it holds a significant fraction of the total mass. We find that the best fit is provided by a disk mass of Mdisk = 4 × 10−3 M⊙. A cross section of the model is shown in Fig. 7. The velocity field in the envelope is similar to the one used in Paper I, where the velocity field was parameterized in terms of a central mass and an angle between the ve- locity vector and the azimuthal direction, so the the ratio of infall to rotation could be controlled by adjusting this angle. The best fit parameter values obtained in Paper I of 1.35M⊙ for the central mass and flow lines inclined with 15◦ with respect to the azimuthal direction, are used. No radial motions are allowed in the disk: only full Keplerian motion is present in the region of the model occupied by the disk. In order to produce synthetic observations which can be directly compared to the data, we use numerical ra- diative transfer tools. The 3D continuum radiative trans- fer code RADMC (Dullemond & Dominik 2004) is used to calculate the scattering function and the temperature structure in the analytical axisymmetric density distribu- tion. In these calculations, the luminosity measurement by Kenyon et al. (1993) of 3.7 L⊙ is used for the central source. The solution is then "ray -- traced" using RADICAL (Dullemond & Turolla 2000) to produce the spectral en- ergy distribution and the continuum maps at the observed frequencies. In both RADMC and RADICAL we assume certain properties of the dust. We use the dust opacities that give the best fit to extinction measurements in dense cores (Pontoppidan et al., in prep). The same density and temperature structure are af- terward given as input to the excitation and line radia- tive transfer code RATRAN (Hogerheijde & van der Tak 2000) which is used to calculate the spatial and frequency dependent HCO+ J = 3 -- 2 emission. The HCO+ mod- els are post-processed with the MIRIAD tasks uvmodel, invert, clean, and restore to simulate the actual SMA ob- servations. 4.2. Modeling the continuum emission In Fig. 8, we compare our model to the continuum obser- vations. We calculate (u, v) -- amplitudes at 1.1 mm which we plot on top of the observed amplitudes. The result is in good agreement with the data, suggesting that the dust emission is well-described by our parameterization down C. Brinch et al.: A deeply embedded young protoplanetary disk around L1489 IRS 7 Fig. 8. The averaged (u, v) -- amplitudes of the contin- uum at 1 mm in both compact and extended config- urations. The full line show our model. The compact configuration visibilities cover the (u, v) -- distance up to 60 kλ. The black triangle marks the total flux at 1 mm (Moriarty-Schieven et al. 1994). The histogram indi- cates the zero-signal expectation values. Fig. 9. The model spectral energy distribution assuming inclinations of 40◦ (full line) and 50◦ and 74◦ (broken lines) plotted on top of flux measurements from the liter- ature (marked by crosses) and the Spitzer/IRS spectrum from 2 -- 40 microns (red line). to scales of ∼ 100 AU. This comparison is not very de- pendent on the inclination, since the continuum emission is quite compact. For this particular figure, an inclination of 40◦ is assumed. Recently, Eisner et al. (2005) modeled the SEDs of a number of Class I objects in Taurus including L1489 IRS. The model used in their work is parameterized differently from ours, but it essentially describes a similar structure. As pointed out above, Eisner et al. find a best fit inclina- tion of 36◦ in contrast to the inclination of 74◦ found in Paper I. To test the result of Eisner et al. (2005) we calculated the SED using the described model with only the system inclination as a free parameter (Fig. 9). The best fit is found for an inclination of about 40◦, in good agreement with the result of Eisner et al. (2005). The models with inclinations of 50◦ and 74◦, are also plotted in Fig. 9 in order to show how the SED depends on inclination. Note that the quality of our best fit to the SED is comparable to the fit presented by Eisner et al. (2005), whereas the fit becomes rapidly worse with increasing inclination larger than 40◦, thus bringing about dis- agreement with the results from (Hogerheijde 2001) and Paper I. The three different models plotted in Fig. 9 can- not be distinguished for wavelength above ∼60 µm, which corresponds to a temperature of about 40 K using Wiens displacement law. This temperature occurs on radial dis- tances of approximately 100 AU from the central object which means that outside of this radius, the SED is no longer sensitive to the inclination. We interpret this be- havior as a change in the angular momentum axis on disk scales, causing the disk to be inclined with respect to the envelope. Fig. 10. Three transitions of HCO+ observed by single- dish telescope (see Paper I for details) with our model superposed. The quality of the fit is comparable to the fit in Paper I although a disk model has now been included. 4.3. Modeling the HCO+ emission In Paper I it was found that the characteristic double peak feature of the HCO+ J = 4 -- 3 line could not be reproduced with an inclination below 70◦. On the other hand including a disk inclined by 40◦ into the model of Paper I does not alter the fit to the single dish lines (Fig 10): the geometry and the velocity field of the material at scales smaller than ∼300 AU does not influence the shape of these lines. This fit is not perfect though as it still overestimates the red- shifted wing in the J = 3 -- 2 line and the width of the J = 1 -- 0 line slightly, but the quality of the fit is similar to that presented in Paper I. We need to test how well this tilted disk model works with the line observations from the SMA. It was shown in Sect. 4 that the off-position spectra are weakly dependent on changes in the inclination and that the same is true when using the model where the disk is tilted with respect 8 C. Brinch et al.: A deeply embedded young protoplanetary disk around L1489 IRS Fig. 11. The SMA spectrum towards the center of L1489 IRS with our model, consisting of a disk which is inclined with respect to the envelope, superposed. As in Fig. 6, the model has been imaged with the (u, v)-spacings from the observations. Fig. 12. Similar to Fig. 8 but for the HCO+ J = 3 -- 2 measurements. The triangle marks the total flux obtained from the single dish observations. The amplitudes have been averaged over ±5 kms−1 relative to the systemic ve- locity. The histogram indicates the zero-signal expectation values. The insert is an enlargement of the longest base- lines. to the envelope. The central position however, is seen in Fig. 11 to be very well reproduced in terms of line width and wing shape by this model. The model spectrum is slightly more asymmetric than the data, which means that the infall to rotation ratio is not quite correct on disk scales. The magnitude of the velocity field projected onto our line of sight is correct though, since the width of the line is well fit. We also compared the averaged (u, v) -- amplitudes to evaluate the quality of our model similar to the pro- cedure for the continuum. Fitting the (u, v) -- amplitudes in this way tests whether the model produces the right amount of emission at every scale. The best fit is shown in Fig. 12. The zero-spacing flux has been shown in this plot as well, marked by the black triangle. This method has previously been used to study the abundance varia- tions of given molecules in protostellar envelopes, in par- ticular imaging directly where significant freeze-out occurs in protostellar envelopes (Jørgensen 2004; Jørgensen et al. 2005a). The model does a good job reproducing the ob- served HCO+ brightness distribution on almost all scales with a constant abundance, except around 20 -- 40 kλ where it overestimates the amount of observed flux. This corre- spond to a radius of about 1000 AU, that is well outside of the disk. On the other hand single-dish observations of low-mass protostellar envelopes suggest that these are scales where significant freeze-out in particular of CO may occur at temperatures . 20 − 30 K, which because of the gas-phase chemical relation between CO and HCO+ also reflects directly in the observed distribution of HCO+ (e.g., Jørgensen et al. 2004, 2005b). Despite this, the dif- ference between the model prediction and observations is small with this constant abundance, suggesting that the amount of freeze-out is small in L1489 IRS. This is in agreement with the single-dish studies of Jørgensen et al. (2002, 2004, 2005b), who generally found little depletion in the envelopes of L1489 IRS and other Class I sources in contrast to the more deeply embedded, Class 0, proto- stars. The model parameter values obtained in this section including values from Paper I are summarized in Table 2. The best fit values that are derived in this paper are tuned by hand rather than systematically optimized by a χ2 min- imization method. Therefore we can only give an estimate on the uncertainties. However, Fig. 9 shows that the SED fit gets rapidly worse when changing the inclination, and given the strong dependence of the envelope inclination on the single-dish lines, we estimate that both angles are accurate to within 10◦. For the disk mass and radius, the uncertainties on the parameter values are less well defined for reasons given in Section 4.1, but we estimate an accu- racy within a factor of 2 in each of these parameters. 5. Discussion It seems that depending on the physical size scales that we probe, the solution favors a different inclination. When taking all available observations into account we need to introduce a model where the angular momentum vector of the disk is misaligned with the angular momentum vec- tor of the (rotationally flattened) envelope. We adopt a two component model (illustrated in Fig. 13), disk and envelope, but in reality, the angular momentum axis may indeed be a continuous function of radius on the scales of the disk (. a few hundred AU). The origin of such a misalignment can actually be explained in a simple way by considering the initial conditions of the collapse. C. Brinch et al.: A deeply embedded young protoplanetary disk around L1489 IRS 9 Table 2. Model parameter values Envelope Outer radiusa Envelope massa Inclinationa,b HCO+ abundancea Value 2000 AU 0.093 M⊙ 74◦ 3.0×10−9 Disk Radiusb Disk massb Inclinationb Scale heightc Central massa Central luminosityc Distancec Value 200 AU 4×10−3 M⊙ 40◦ z=0.25R 1.35 M⊙ 3.7 L⊙ 140 pc a Values from Paper I b Fitted values c Values adopted from the literature ω envelope ω disk L.O.S. 4000 AU 500 AU Fig. 13. An (exaggerated) illustration of the proposed model where the angular momentum axis is changing with radius. The line of sight is illustrated by the dashed line. The state of a pre-stellar core before the gravitational collapse begins is typically described by a static sphere with a solid body rotation perturbation (Terebey et al. 1984). Due to strict angular momentum conservation, such a model would be perfectly aligned throughout the col- lapse and consequently disk formation models are often described numerically by axisymmetric representations (e.g., Yorke & Bodenheimer 1999). However, there is no a priori reason why a pre-stellar core should rotate as a solid body. Of course the cloud has an average angular momentum which on a global scale de- termines the axis of rotation. The cloud does not collapse instantly to form a disk though, but rather from the in- side and out as described by Shu (1977), and therefore a shell of material from deep within the cloud, which may very well have an average angular momentum that is dif- ferent from the global average, will collapse and form a disk before material from further out has had a change to accrete yet. Actually, if the parental cloud is turbulent, it is to be expected that the accreting material has ran- domly oriented angular momentum and a misalignment of the angular momentum on different scales is more likely than a perfect alignment. In this way, a system similar to the model proposed to describe L1489 IRS in this paper can be formed. To obtain a gradient in the angular momentum as ini- tial condition for the collapse, one may consider a cloud which is not spherically symmetric or one which does not collapse around its geometrical center but rather around an over dense clump offset from the center. The former sce- nario may indeed be true for the case of L1489 IRS which is seen to be dynamically connected to a neighboring cloud (which was also modeled in Paper I to explain the excess of cold emission seen in some of the low-J lines). Actually, the uniformly weighted moment map (Fig. 2) shows sig- nificant asymmetry with a secondary emission peak to the north-east. It is interesting to note that the HCO+ emission from the SMA observations agree very well with the near- infrared image (Fig. 3): the secondary peak, a few arc sec- onds northeast of the main continuum and HCO+ peaks, nicely coincides with a bright spot in the scattered light image and also the shape of the cavity towards the south follows the contours of the HCO+ emission. The same de- tails are not revealed in the naturally weighted SMA im- age, in which shorter baselines are given more weight. We thus conclude that by going to the extended SMA config- uration, it is possible to probe structure in YSOs on the same scales as can be resolved by large near-infrared tele- scopes such as the Hubble Space Telescope. The structure seen in both images, however, also emphasizes that for a fully self-consistent description of L1489 IRS, a global non-axisymmetric model has to be considered. Another (non-exclusive) explanation of a misaligned disk would be that L1489 IRS formed as a triple stellar system and, that due to gravitational interaction, one of the stars was ejected. L1489 IRS would thus be a binary system as suggested by Hogerheijde & Sandell (2000). The loss of angular momentum due to the ejection would result in a rearrangement of the remaining binary, which could "drag" the inner viscous disk along. Such a scenario has been investigated numerically by Larwood et al. (1996). In the case of L1489 IRS, it would have to be a very close bi- nary with a separation of no more than a few AUs since the near-infrared scattered light image (Padgett et al. 1999) with a resolution of ∼0.2′′ (30 AU), does not reveal mul- tiple sources. The ratio of binary separation to disk ra- dius is therefore significantly lower than the cases investi- gated by Larwood et al. (1996) and it is therefore not clear whether similar effects could be present in L1489 IRS. If indeed the system is binary, it would resolve the issue that we find a central mass of 1.35 M⊙ while the luminosity is estimated to be 3.7 L⊙ (Kenyon et al. 1993). A single young star that massive would require a much higher luminosity, but two stars of 0.6 -- 0.7 M⊙ would fit nicely with the estimated luminosity since luminosity is not a linear function of mass for YSOs. 10 C. Brinch et al.: A deeply embedded young protoplanetary disk around L1489 IRS To test whether L1489 IRS indeed is a binary and to constrain the innermost disk geometry on AU scales, mea- surements of temporal variability in super high resolution are needed. The timescale for variations of a possible bi- nary is of the order of years depending on the exact sepa- ration of the stars (assuming scales of 1 AU). Such obser- vations would, for example, be feasible with ALMA. 6. Conclusion In this paper, we have presented high angular resolution interferometric observations of the low-mass Class I YSO L1489 IRS. The observations reveal a rotationally dom- inated, very structured central region with a radius of about 200 -- 300 AU. We interpret this as a young Keplerian disk which is still deeply embedded in envelope material. This conclusion is supported by a convincing fit with a disk model to the SED. We conclude further that the inclination of the disk is not aligned with the inclination of the flattened envelope structure, due to the possibility that L1489 IRS is a binary system and/or that the average angular momentum axis of the cloud is not aligned with the angular momentum axis of the dense core that originally collapsed to form the star(s) plus the disk. We find that a disk with a mass of 4 × 10−3 M⊙, a radius of 200 AU, and a pressure scale height of z=0.25R is consistent with both the SED and the HCO+ observa- tions. Only a small amount of chemical depletion of HCO+ is allowed for, due to a slight over-estimate of the (u, v) -- amplitudes at 20 -- 40 kλ by our model, in agreement with the results from previous single-dish studies and the na- ture of L1489 IRS as a Class I YSO. The combination of a detailed modeling of the SED with spatially resolved line observations, which contains information on the gas kinematics, appears to be a very efficient way of determining the properties of disks, es- pecially embedded disks which are not directly observable. Acknowledgments The authors would like to thank Kees Dullemond for making his code available to us. CB is supported by the European Commission through the FP6 - Marie Curie Early Stage Researcher Training programme. JKJ acknowledges support from an SMA fellowship. AC was supported by a fellowship from the European Research Training Network "The Origin of Planetary Systems" (PLANETS, contract number HPRN-CT-2002-00308). The research of MRH and AC is supported through a VIDI grant from the Netherlands Organiztion for Scientific Research. References Basu, S. 1998, ApJ, 509, 229 Boogert, A. C. A., Hogerheijde, M. R., & Blake, G. A. 2002, ApJ, 568, 761 Brinch, C., Crapsi, A., Hogerheijde, M. R., & Jørgensen, J. K. 2007, A&A, 461, 1037 Burrows, C. J., Stapelfeldt, K. R., Watson, A. M., et al. 1996, ApJ, 473, 437 Cassen, P. & Moosman, A. 1981, Icarus, 48, 353 Dullemond, C. P. & Dominik, C. 2004, A&A, 417, 159 Dullemond, C. P. & Turolla, R. 2000, A&A, 360, 1187 Eisner, J. A., Hillenbrand, L. A., Carpenter, J. M., & Wolf, S. 2005, ApJ, 635, 396 Hogerheijde, M. R. 2001, ApJ, 553, 618 Hogerheijde, M. R. & Sandell, G. 2000, ApJ, 534, 880 Hogerheijde, M. R. & van der Tak, F. F. S. 2000, A&A, 362, 697 Hogerheijde, M. R., van Dishoeck, E. F., Blake, G. A., & van Langevelde, H. J. 1997, ApJ, 489, 293 Jørgensen, J. K. 2004, A&A, 424, 589 Jørgensen, J. K., Bourke, T. L., Myers, P. C., et al. 2005a, ApJ, 632, 973 Jørgensen, J. K., Schoier, F. L., & van Dishoeck, E. F. 2002, A&A, 389, 908 Jørgensen, J. K., Schoier, F. L., & van Dishoeck, E. F. 2004, A&A, 416, 603 Jørgensen, J. K., Schoier, F. L., & van Dishoeck, E. F. 2005b, A&A, 435, 177 Kenyon, S. J., Calvet, N., & Hartmann, L. 1993, ApJ, 414, 676 Kessler-Silacci, J. E., Hillenbrand, L. A., Blake, G. A., & Meyer, M. R. 2005, ApJ, 622, 404 Lada, C. J. & Wilking, B. A. 1984, ApJ, 287, 610 Larwood, J. D., Nelson, R. P., Papaloizou, J. C. B., & Terquem, C. 1996, MNRAS, 282, 597 Lucas, P. W., Blundell, K. M., & Roche, P. F. 2000, MNRAS, 318, 526 Moriarty-Schieven, G. H., Wannier, P. G., Keene, J., & Tamura, M. 1994, ApJ, 436, 800 Motte, F. & Andr´e, P. 2001, A&A, 365, 440 Myers, P. C., Fuller, G. A., Mathieu, R. D., et al. 1987, ApJ, 319, 340 Ohashi, N., Hayashi, M., Kawabe, R., & Ishiguro, M. 1996, ApJ, 466, 317 Padgett, D. L., Brandner, W., Stapelfeldt, K. R., et al. 1999, AJ, 117, 1490 Park, S. & Kenyon, S. J. 2002, AJ, 123, 3370 Qi, C. 2005, The MIR Cookbook, The Submillimeter Array/Harvard-Smithsonian Center for Astrophysics Qi, C., Ho, P. T. P., Wilner, D. J., et al. 2004, ApJ, 616, L11 Reipurth, B., Jewitt, D., & Keil, K., eds. 2007, Protostars and Planets V (University of Arizona Press) Rodr´ıguez, L. F., D'Alessio, P., Wilner, D. J., et al. 1998, Nature, 395, 355 Adams, F. C., Lada, C. J., & Shu, F. H. 1987, ApJ, 312, Saito, M., Kawabe, R., Kitamura, Y., & Sunada, K. 2001, 788 Adams, F. C., Shu, F. H., & Lada, C. J. 1988, ApJ, 326, 865 ApJ, 547, 840 Sault, R. J., Teuben, P. J., & Wright, M. C. H. 1995, Astronomical Data Analysis Software and Systems IV, C. Brinch et al.: A deeply embedded young protoplanetary disk around L1489 IRS 11 77, 433 Schoier, F. L., van der Tak, F. F. S., van Dishoeck, E. F., & Black, J. H. 2005, A&A, 432, 369 Shu, F. H. 1977, ApJ, 214, 488 Terebey, S., Shu, F. H., & Cassen, P. 1984, ApJ, 286, 529 Thi, W. F., van Dishoeck, E. F., Blake, G. A., et al. 2001, ApJ, 561, 1074 Ulrich, R. K. 1976, ApJ, 210, 377 Whitney, B. A., Kenyon, S. J., & Gomez, M. 1997, ApJ, 485, 703 Yorke, H. W. & Bodenheimer, P. 1999, ApJ, 525, 330
astro-ph/0010187
1
0010
2000-10-10T14:28:59
Lensing in the Hercules Supercluster
[ "astro-ph" ]
We report Keck LRIS observations of an arc-like background galaxy near the center of Abell 2152 (z=0.043), one of the three clusters comprising the Hercules supercluster. The background object has a redshift z=0.1423 and is situated 25 arcsec north of the primary component of the A2152 brightest cluster galaxy (BCG). The object is about 15 arcsec in total length and has a reddening-corrected R-band magnitude of $m_R = 18.55\pm0.03$. Its spectrum shows numerous strong emission lines, as well as absorption features. The strength of the H-alpha emission would imply a star formation rate $\SFR \approx 3h^{-2} \msun $yr$^{-1}$ in the absence of any lensing. However, the curved shaped of this object and its tangential orientation along the major axis of the BCG suggest lensing. We model the A2152 core mass distribution including the two BCG components and the cluster potential. We present velocity and velocity dispersion profile measurements for the two BCG components and use these to help constrain the potential. The lens modeling indicates a likely magnification factor of $\sim1.9$ for the lensed galaxy, making A2152 the nearest cluster in which such significant lensing of a background source has been observed. Finally, we see evidence for a concentration of early-type galaxies at $z=0.13$ near the centroid of the X-ray emission previously attributed to A2152. We suggest that emission from this background concentration is the cause of the offset of the X-ray center from the A2152 BCG. The background concentration and the dispersed mass of the Hercules supercluster could add further to the lensing strength of the A2152 cluster.
astro-ph
astro-ph
To appear in AJ, January 2001 Preprint typeset using LATEX style emulateapj 0 0 0 2 t c O 0 1 1 v 7 8 1 0 1 0 0 / h p - o r t s a : v i X r a LENSING IN THE HERCULES SUPERCLUSTER1 John P. Blakeslee2,3,4, Mark R. Metzger3, Harald Kuntschner2, and Patrick Cot´e3,5,6 To appear in AJ, January 2001 ABSTRACT We report Keck LRIS observations of an arc-like background galaxy near the center of Abell 2152 (z = 0.043), one of the three clusters comprising the Hercules supercluster. The background object has a redshift z = 0.1423 and is situated 25′′ north of the primary component of the A2152 brightest cluster galaxy (BCG). The object is about 15′′ in total length and has a reddening-corrected R-band magnitude of mR = 18.55 ± 0.03. Its spectrum shows numerous strong emission lines, as well as absorption features. The strength of the Hα emission would imply a star formation rate SFR ≈ 3 h−2 M⊙ yr−1 in the absence of any lensing. However, the curved shaped of this object and its tangential orientation along the major axis of the BCG suggest lensing. We model the A2152 core mass distribution including the two BCG components and the cluster potential. We present velocity and velocity dispersion profile measurements for the two BCG components and use these to help constrain the potential. The lens modeling indicates a likely magnification factor of ∼ 1.9 for the lensed galaxy, making A2152 the nearest cluster in which such significant lensing of a background source has been observed. Finally, we see evidence for a concentration of early-type galaxies at z = 0.13 near the centroid of the X-ray emission previously attributed to A2152. We suggest that emission from this background concentration is the cause of the offset of the X-ray center from the A2152 BCG. The background concentration and the dispersed mass of the Hercules supercluster could add further to the lensing strength of the A2152 cluster. Subject headings: galaxies: clusters: individual (Abell 2152) -- galaxies: elliptical and lenticular, cD -- gravitational lensing 1. INTRODUCTION Strong gravitational lensing of background sources by clusters of galaxies provides some of the most unambigu- ous evidence for the presence of large amounts of dark matter in clusters. Lensing studies suggest that the mass in clusters is greater and more centrally concentrated than implied, for instance, by the X-ray properties (e.g., Mel- lier, Fort, & Kneib 1993; LeF`evre et al. 1994; Waxman & Miralda-Escude 1995; Smail et al. 1995). Moreover, the mass appears to clump around the luminous galaxies, so that consideration of substructure and the effect of the massive central galaxy are often important for understand- ing the observations (e.g., Miralda-Escud´e 1995; see also Tyson, Kochanski, & Dell Antonio 1998). This is partic- ularly true in more moderate mass clusters with velocity dispersions σcl ∼< 1000 km s−1, for which the proportion- ally larger gravitational effects of a massive cD must be included to properly account for the lensing (Williams, Navarro, & Bartelmann 1999). Although there have been great advances in our knowl- edge of the form of the density distributions of massive dissipationless particles in simulations of various resolu- tion (e.g., Dubinski & Carlberg 1991; Navarro, Frenk, & White 1996, 1997; Moore et al. 1998, 1999b), our knowl- edge of the central mass distributions of actual galaxy clus- ters remains fragmentary. Low redshift clusters, because of their larger angular sizes, can in principle provide a better opportunity for studying these mass distributions. Their lensing properties are acutely sensitive to the shape of the inner cluster potential, although because such clus- ters tend to be of modest mass (very high-mass clusters being few and far between), the individual galaxies will produce relatively larger perturbations on this potential, complicating the interpretation. Until recently, moderate- to high-redshift clusters have been the exclusive purview of strong lensing studies, but lensing by low-redshift clusters has been gaining increased attention. For instance, Allen, Fabian, & Kneib (1996) discovered a redshift z = 0.43 lensed arc in the massive cooling flow cluster PKS 0745 -- 191 at z = 0.103. Campu- sano, Kneib, & Hardy (1998) explored lensing models for a z = 0.073 spiral galaxy located near the central ellip- tical in the richness class 0 cluster A3408 at z = 0.042; they obtained an upper limit to the magnification factor of ∼1.7 from their "maximum mass" model. This object turned out to be the only viable lensing candidate found in an imaging survey of 33 southern galaxy clusters with z ≤ 0.076 by Cypriano et al. (2000), who also presented an estimate of the expected number of arcs and arclets in low-redshift clusters. Blakeslee & Metzger (1999) discov- ered a lensed arc at z = 0.573 in the nearby cD cluster A2124 at z = 0.066 and found magnification factors near 1Based on observations obtained at the W.M. Keck Observatory, which is jointly operated by the California Institute of Technology, the University of California, and the National Aeronautics and Space Administration. 2Department of Physics, University of Durham, South Road, Durham, DH1 3LE, United Kingdom 3California Institute of Technology, Mail Stop 105-24, Pasadena, CA 91125 4Current address: Department of Physics and Astronomy, Johns Hopkins University, Baltimore, MD 21218 5Sherman M. Fairchild Fellow 6Current address: Department of Physics and Astronomy, Rutger University, 136 Frelinghuysen Road, Piscataway NJ, 08854-8019 1 2 Lensing in Hercules 10 from their best-fitting models of this system. Despite the recent progress, the lensing properties of low-redshift clusters remain unclear. For instance, A2124 had by far the lowest redshift in the sample studied by Williams et al. (1999) of 24 lensing clusters with measured velocity dispersions, the next nearest being at z = 0.171. In this paper, we present imaging and spectroscopic obser- vations and give a lensing analysis of the redshift z = 0.043 cluster Abell 2152 in the Hercules supercluster. 2. CLUSTER PROPERTIES The Hercules supercluster is a close grouping of three Abell clusters: the richness class 2 cluster A2151 (the clas- sical "Hercules cluster") and the two richness 1 clusters A2147 and A2152, all at z ∼ 0.04. The supercluster was first identified by Shapley (1934). Although A2151 domi- nates from the standpoint of the number of galaxies, A2152 is projected closest to the center of the grouping, which is < 4 h−1 Mpc) in diame- so tight that a circle of just 2◦ ( ∼ ter easily encompasses all three clusters. Two other Abell clusters, A2148 and A2107, lie at about the same redshift and are only ∼ 15 h−1 Mpc away. At lower density en- hancements, the Hercules supercluster forms part of an ex- tended, filamentary supercluster that includes A2162 and A2197/A2199 about 3000 km s−1 in the foreground and A2052/A2063 about 3000 km s−1 in the background (Abell 1961; Postman, Huchra, & Geller 1992). This extended ten-cluster supercluster includes more members than any other supercluster in the Postman et al. (1992) catalogue. All three members of the Hercules supercluster proper are classified as Bautz-Morgan type III, Rood-Sastry class F clusters (Abell et al. 1989; Struble & Rood 1987), in- dicating morphological irregularity and perhaps dynami- cal youth. All three have relatively high spiral fractions, with that of A2151 being near 50% (Tarenghi et al. 1980). A2151 is also famous for its high degree of internal sub- clustering (e.g., Bird et al. 1995). A2152, the focus of this paper, has two bright early-type galaxies at its center (see Figure 1) that together are designated UGC 10187. The brighter component (NED 02) was chosen by Post- man & Lauer (1995) as the A2152 brightest cluster galaxy (BCG) and has a redshift z = 0.0441. The other compo- nent (NED 01) is 47′′ to the northwest and about 0.5 mag fainter; Postman & Lauer classify it as the second ranked galaxy (SRG). The SRG has a relative velocity of +330 km s−1 with respect to the BCG (see §3.3). Because A2152 and A2147 are separated by just 1.8 h−1 Mpc, measurements of their velocity dispersions are problematic. Zabludoff et al. (1993) reported σcl = 1081 km s−1 for A2147 and σcl = 1346 km s−1 for A2152, the latter being the highest dispersion in their sample of 25 "dense peaks." Recently, Barmby & Huchra (1998) have carried out a detailed analysis of the Hercules supercluster kinematics using 414 galaxy velocities. These authors em- ployed a four-component model, including the three Abell clusters and a "dispersed supercluster" component. They derived much lower velocity dispersions of 821+68 −55 km s−1 and 715+81 −61 km s−1 for A2147 and A2152, respectively, from 93 and 56 galaxy velocities. Because 122 galaxies in the Barmby & Huchra sample are assigned to the dispersed component with a high ve- locity dispersion of 1407 km s−1, rather than to one of the individual clusters, this approach may tend to bias the cluster dispersions low. However, it is clearly more consistent in finding a mean redshift of hzi = 0.0432 for A2152, much closer to that of the BCG, whereas previous studies found means 10 -- 15% lower, near the mean values for A2147 and A2151. Moreover, the lower velocity disper- sions are more in line with the X-ray properties of these clusters, e.g., the X-ray gas temperature in A2152 has been estimated from Einstein data to be kTX = 2.1 keV (White, Jones, & Forman 1997). We note, however, that the latest tabulation by Struble & Rood (1999) lists hzi = 0.0398 and σcl = 1338 km s−1 for A2152 from 62 velocities. The centroid of the X-ray emission associated with A2152 is not coincident the BCG but is about 2.′1 east. Its position is within 40′′ of a pair of bright early-type galax- ies which the Barmby & Huchra redshift catalogue reveals to be at z = 0.134. Each of these background galaxies is intrinsically ∼ 70% more luminous than the A2152 BCG (using a relative K-correction of 0.11 mag for ellipticals). We return to this point in §5. 3. OBSERVATIONS A2152 was observed as part of the deep R-band Keck imaging study by Blakeslee (1999) of the BCG halo regions in six nearby rich clusters. Clusters were selected to have velocity dispersions similar to that of the Coma cluster and to lie in the redshift range cz = 10,000 -- 20,000 km s−1. These six were then observed because they were reachable during a single night in the northern spring. Based on this dataset, Blakeslee & Metzger (1999) reported follow- up spectroscopy of a magnitude mR = 20.86 lensed arc located 27′′ from the center of the cD galaxy in Abell 2124. We have visually searched the dataset for other bright lensing candidates, based on proximity to the BCG, tangential orientation, and arc-like appearance. The only obvious bright candidate was an arc-like object in A2152. We obtained follow-up Keck spectroscopy for this object and the A2152 BCG/SRG pair. 3.1. Imaging Data The imaging observations have already been described in detail by Blakeslee (1999). Here we summarize them and report new photometric results for the A2152 arclet. A2152 was imaged with the Low Resolution Imaging Spec- trograph (LRIS, Oke et al. 1995) on the Keck II tele- scope under photometric conditions. The image scale was 0.′′211 pix−1, and the total integration was 2000 s in the R band. Reductions were carried out using the Vista package. The seeing in the final stacked image was an exceptional 0.′′53. The photometry was calibrated to the Cousins R band using Landolt (1992) standard stars; the photometric zero point is accurate to better than 0.02 mag. Figure 1 displays the A2152 Keck LRIS image, a smaller portion of which was shown by Blakeslee (1999). The pos- sible lensed galaxy (galaxy "A") is located 25.′′2 from the center of the BCG at a position angle PA = 11.◦4. The ma- < 15′′, where the jor axis of the BCG is at 11◦ ± 2◦ at radii ∼ ellipticity (difference of the axis ratio from unity) of the light distribution is ǫℓ = 0.10. Beyond this radius the BCG isophotes have significant overlap with those of the SRG and they are sufficiently round that the angle is difficult to determine. The isophotes of the SRG show significant twisting; the PA of its major axis swings from 170◦ at Blakeslee, Metzger, Kuntschner, & Cot´e 3 r ∼< 5′′ to ∼ 136◦ by r ≈ 25′′, while the ellipticity stays fairly constant around ǫℓ = 0.32 ± 0.03. G1/G2 are in the background with redshifts of z = 0.1335 and z = 0.1353, respectively. During our spectroscopic ob- servation of the BCG/SRG (described in §3.3), the LRIS slit picked up the magnitude mR = 19.2 galaxy labeled G3, which turns out to have z = 0.1326. Galaxy "M" is an A2152 member with a velocity within ∼ 100 km s−1 of the BCG (Barmby & Huchra). The center of the X-ray emission in this field (Jones & Forman 1999) is 20′′ from G2 and 50′′ from G1 and is also labeled in Figure 1. Figure 2 shows an enlarged view of galaxy A along with an isophotal contour map. The impression is that of a disk galaxy whose disk has been warped into an arc-like shape (arguably through lensing, as we discuss in §4). The object is visible in our image for a length of about 15′′. The measured full-width at half maximum along the mi- nor axis is 0.′′67, which would be about FWHM ∼ 0.′′4 in the absence of atmospheric blurring. This translates to a < 0.7 h−1 kpc at the redshift of this object (see metric size ∼ below); apparently it has a very compact nucleus. The to- tal magnitude of A, corrected for 0.105 mag of R-band Galactic extinction (Schlegel, Finkbeiner, & Davis 1998), is mR(arc) = 18.55 ± 0.03. Our main measurement results are collected in Table 1. 3.2. Spectroscopy of Galaxy A The simplest and most important test of gravitational lensing is whether or not the supposed lensed object is actually in the background. Thus, we obtained one 900 s spectrum of galaxy A with LRIS on the Keck II telescope on the night of 25 March 1999 (UT). We used a 300 line grating and a 1′′ wide slit oriented at PA = 113◦, approx- imately along the major axis of galaxy A. The seeing was poor, about 1.′′8, so the slit missed a good fraction of the light. The dispersion was 2.44 A pix−1, and the effective resolution was about 10 A (FWHM). Halogen flats and arc-lamp exposures were taken for calibration. The LRIS configuration and IRAF reductions were identical to those of Blakeslee & Metzger (1999), except that no flux stan- dard spectrum was obtained. Fig. 2. -- Two views of the possibly lensed galaxy in A2152 (galaxy A). North is up and east is to the left. The top view shows a 1′ × 1′ region of the R-band LRIS image centered on the galaxy. The stretch is linear, and a model of the BCG/SRG halo light has been subtracted. The cross marks the center of the subtracted BCG, which is 25′′ from A. The lower panel zooms in by a factor of two, showing an isophotal contour map of the 30′′ square region around A. The outermost con- tour is at an extinction-corrected surface brightness of 24.8 mag arcsec−2, corresponding 2.0% of the sky level in the image, and the other contours are in steps of 1 mag. The curvature of the object is apparent in both panels. All galaxies with known redshifts (either from Barmby & Huchra or the present study) are labeled in Figure 1. It is somewhat curious that of 7 galaxies (including galaxy A) in the central ∼ 6′ field of a rich cluster, only 3 turn out to be cluster members. As mentioned above, the galaxy pair Figure 3 shows the reduced 1-d spectrum of galaxy A with the strongest emissions lines labeled at top and some absorption lines labeled at the bottom. The emission ap- peared to be confined to the central ∼ 4′′ of the galaxy, so this is all that was extracted. The spectrum is of very high signal-to-noise (it was not a priori obvious, in the absence of spiral structure or any spectral information, that there whould be such strong features), and we determine the redshift to be z = 0.1423 ± 0.0001. The object is therefore about 3.3 times more distant than the intervening A2152 cluster, and the lensing hypothesis remains viable. If un- lensed, the absolute magnitude of A, with a K-correction of 0.05 mag, would be MR = −20.57 (for h = 0.7 and Ω=0; MR = −21.22 for h = 0.5 and Ω=1), or a luminosity of about 0.48 L∗ R (e.g., Lin et al. 1996). The Hα emission line has an equivalent width of EWHα = 117 ± 6 A, corrected for [N II] emission. The Hβ equivalent width is EWHβ = 28.2 ± 2.0 A. The [O II] λ3727 and [O III] λ5007 lines have equivalent widths of 83.3 ± 2.3 Aand 54.1 ± 2.0 A, respectively. Thus, the emis- sion is quite strong, but not too anomalous in compari- son to many late-type spirals or irregulars (e.g., Kennicutt 1992). 4 Lensing in Hercules Fig. 3. -- Reduced 1-d spectrum of galaxy A. The redshift is z = 0.1423. Various emission lines are identified above the spectrum and some absorption features are labeled at bottom, including the Ca II H and K lines, G band, Mg b, and Na D. The atmospheric A and B absorption bands are also labeled; A-band absorption affects the [Sr II] lines, but not Hα. Although we do not have a good absolute flux calibra- tion of the spectrum, we can roughly estimate the star for- mation rate (SFR) from the Hα equivalent width and the R-band luminosity. We use the spectrophotometric syn- thesis results of Fukugita, Shimasaku, & Ichikawa (1995) to calibrate the absolute flux in Cousins R, which has an effective wavelength of 6588 A. About 75% of the R-band light comes from the region of emission within the galaxy, and from the measured Hα equivalent width we obtain a flux fHα ≈ 7.38 × 10−15 ergs s−1 cm−2 and a luminosity of LHα ≈ 1.84 × 1041 h−2 ergs s−1 (Ω = 0). To estimate the internal extinction, we perform a makeshift relative flux calibration using the Feige 110 spectrum from Blakeslee & Metzger (1999) and find an Hα-to-Hβ flux ratio of 4.25; assuming an unextincted value of 3.0 and the extinction curve of Cardelli, Clayton & Mathis (1989) gives a cor- rection at Hα of a factor of 2.15. This is close to the "canonical" correction of ∼1 mag for spirals (e.g., Kenni- cutt 1992) but will yield a SFR 16% lower. Finally, using the updated calibration from Kennicutt (1998), we find SFR ≈ (3 ± 1) h−2 M⊙ yr−1, or (6 ± 2) M⊙ yr−1 for a typi- cal H0. Though not qualifying as a "starburst galaxy," the SFR is quite substantial for a 0.5 L∗ galaxy. The quoted errorbar is approximate and reflects mainly the ∼ 30% scatter quoted by Kennicutt. The calibration based on [O II] equivalent width yields nearly the same SFR for this galaxy. 3.3. Kinematics of the Central Galaxy With the intent of improving the constraints on the A2152 potential for the lensing models, we obtained three 1100 s spectra of the BCG/SRG pair with LRIS on 16 June 1999. We again used a 1′′ wide slit but the 900/5500 grat- ing, yielding a dispersion of 0.84 A pix−1. The slit was oriented at PA = 122◦ in order to pass through the cen- ters of the two galaxies. Again, halogen flats and arc lamps were taken for calibration, and we also obtained spectra of the K4 III velocity standard HD 213947. The spectra were processed, wavelength calibrated, and rectified using standard IRAF routines. Figure 4 displays the central 1.′′1 extracted from the re- duced 2-d spectra. Although the spectra are clearly very similar, the BCG exhibits a small amount of Hβ emission that is absent from the SRG. Such optical line emission is quite common among central cluster galaxies, and may be associated with cluster "cooling flows" (e.g., Crawford et al. 1999). This emission may therefore be evidence that the BCG is indeed at the center of the A2152 cluster po- tential and that the offset of the Jones & Forman (1999) X-ray position is due to contamination from a background X-ray source. The kinematical analysis of the data was performed within the MIDAS package. First, the spectra were re- binned along the slit (spatial direction) in order to achieve a minimum signal-to-noise (S/N) of 30 per A at all radii. The S/N of the central pixels is higher. Then the contin- uum was removed by a fifth order polynomial fit. Finally we determined the velocity dispersion and rotational ve- locity as a function of radius with the Fourier Correlation Quotient (FCQ) method (Bender 1990) allowing only for a Gaussian stellar velocity distribution. When we allowed for non-Gaussian distributions and fitted for the higher- order moments, the results for the rotational velocity and Blakeslee, Metzger, Kuntschner, & Cot´e 5 dispersion were unchanged within the errors but showed slightly more scatter at the largest radii. The detailed FCQ setup, such as continuum fit order, wavelength ex- traction and error estimation was optimized with Monte Carlo simulations. Fig. 4. -- The central 1.′′1 extracted from the 2-d spectra of the BCG and SRG (observed simultaneously through the slit) are illustrated. The spectra have been deredshifted for pur- poses of comparison. Some emission is visible within the Hβ absorption at λ = 4861 A in the spectrum of the BCG, but not in the SRG. Figure 5 shows the measured velocity and velocity dis- persion profiles for the BCG and SRG. Note that the 122◦ position angle is ∼ 70◦ away from the major axis of the BCG, and so the profile is more appropriate to the minor axis. For the SRG, the PA is closer to the major axis. Both galaxies show significant rotation, and both rotate in the sense opposite to that of the combined BCG/SRG system. Although the velocity dispersions decline outward from the centers of the galaxies, the dispersion profile for the SRG appears to flatten at a value of σ ∼ 230 km s−1 outside the central 3′′. Likewise, the BCG's velocity dispersion > 7′′. As profile does not show a significant decrease for r ∼ seen in the top panel of Figure 5, the center of rotation in the BCG is offset by about 0.′′2 (1 LRIS pixel) from the luminosity center; this is likely due to a slight offset of the galaxy center within the slit. The mean velocities, rotational velocities, and central dispersions are tabulated along with our other results in Table 1. We have estimated the total (1-sigma) errors on the mean velocities to be ±25 km s−1, but this is dominated by zero-point calibration errors (flexure, illumination of slit, etc.). However, the relative velocity of the BCG/SRG pair is better determined at ∆v = 331±7 km s−1. This accords well with previous measurements of ∆v = 327±100 km s−1 from Tarenghi et al. (1979) and ∆v = 341±50 km s−1 from Wegner et al. (1999); less well with ∆v = 502 ± 21 km s−1 from Davoust & Consid`ere (1995). (The velocities used by Barmby & Huchra for these two galaxies come from Tarenghi et al.). Fig. 5. -- Velocity and dispersion profiles for the BCG and SRG (labeled). Note that the radial coordinate defined here increases to the southeast along the slit, i.e., the BCG is in the direction of positive radius with respect to the SRG. Thus, the point in the SRG's dispersion profile at (r, σ) = (8.′′6, 291) may be beginning to show the effects of the BCG halo light. Central velocity dispersions have been measured previ- ously for the BCG by Oegerle & Hoessel (1991), who found σ0 = 295±25 km s−1, and by Wegner et al. (1999), who re- ported σ0 = 280±26 km s−1; both results are in agreement with our much more precise value of σ0 = 295 ± 7 km s−1. For the SRG, we find a central dispersion of σ0 = 271 ± 8 km s−1, while Wegner et al. found σ0 = 313 ± 34 km s−1. Their significantly higher dispersion is likely due to the strong rotation in this galaxy, which these authors did not account for when extracting the spectrum of the central several arcseconds of the galaxy. Finally, our BCG mean velocity is ∼ 2.5 σ discrepant with the Wegner et al. value, which is higher by 95 km s−1, but it agrees within the er- rors with the values from Tarenghi et al. (1979), Oegerle & Hoessel (1991), and Davoust & Consid`ere (1995). No pre- vious measurements of the rotation or dispersion profile in either galaxy could be found. Finally, as noted in §3.1, the LRIS slit picked up the magnitude mR = 19.24 galaxy labeled G3 in Figure 1. We find it to be in the background with a velocity cz = 39,738 km s−1. This is just ∼ 290 km s−1 less than the velocity of G1 and ∼ 830 km s−1 less than that of G2. 6 Lensing in Hercules 4. IS IT LENSED? To model A2152 as a lens, we use a potential of the form (Blandford & Kochanek 1987) mary lens (cluster dark matter) can reproduce the char- acteristics of the arc. These models are degenerate in σ1D and rc for determining the mass enclosed within the arc radius. Model 1 fits within this family of models. Fig- ure 6 shows the predicted magnifications from model 1 with a 10′′ core as a function of observed source position. The position of galaxy A implies a magnification of ∼ 1.9, dropping to ∼ 1.8 for a model with a 40′′ soft core. Given this model, we can ask what the intrinsic source shape of galaxy A would be; Figure 7 shows contours of the observed galaxy surface brightness along with the same contours mapped back to the source plane via the lens equation. The implied shape is that of a fairly symmetric galaxy with a disk and dominant bulge, with no indication of a warp. While we cannot make quantitative estimates of cluster mass parameters with a single lensed galaxy, we find that a simple model constructed from a low-dispersion cluster centered at the BCG can produce the shear needed to ex- plain the curved shape of the background galaxy from a more symmetric source. Indeed, galaxy A must be dis- torted in some fashion by the foreground cluster, and we can rule out two of our other three plausible models based on the implied lensing geometry. We cannot, however, rule out a low-dispersion cluster mass centered on the X- ray center from lensing alone, although this would mean galaxy A must be intrinsically warped. Another galaxy with an unusual shape, elongated radially from the BCG center, is visible about 11′′ to the east of the BCG (see the upper panel of Fig. 2) and lies near some critical radii of model 1. If in the background, this galaxy may pro- vide much better constraints on the shape of the inner potential, however the redshift is unknown and we do not attempt to model it here. Fig. 6. -- Magnification of a background source at z = 0.142 produced by the simple lens model. The model shown has σcl = 715 km s−1 with a 10 arcsec soft core, σBCG = 250 km s−1, and σSRG = 240 km s−1. Starting from the outside, contours indicate magnifications of 1.5, 1.73, 2, 4, and 16. The critical curve is near the magnification=16 contour, and a sec- ond critical curve is present close to the center. ψ(r′) = 4π(cid:16) σ1D DS  1 + c (cid:17)2 DLS  1 2 r′2 r2 c ! , − 1  where σ1D is the line-of-sight velocity dispersion in the limit r′ ≫ rc for the spherical case, DS and DLS are an- gular size distances to the source and from the lens to the source, respectively, and rc defines a softening radius. An ellipticity parameter ǫp can be introduced through r′2 = (1−ǫp)x2 + (1+ǫp)y2, where (x, y) are coordinates aligned with the major and minor axes of the potential. This represents a softened isothermal sphere for ǫp = 0 and provides a fairly good representation of the dark halo potentials in the cores of clusters (e.g., Mellier et al. 1993; Tyson et al. 1998). We use the more limited relation with a fixed power law because there are too few lensing con- straints in this system to differentiate various potential profiles, and we have used only circularly symmetric po- tentials as there is little constraint on the ellipticity. We explored four different simple models for the lensing potential. We included three terms in each of these mod- els, corresponding to the smooth halo mass in the cluster and the masses associated with the BCG and SRG. For the latter two components, we use analytic potentials to ap- proximate the velocity dispersions measured in §3.3 with a small core. In our model 1, the center of the cluster potential was fixed at the cD center with an asymptotic dispersion of σcl = 715 km s−1 as measured by Barmby & Huchra (1998) and a softened core radius of 10′′. Model 2 also fixes the cluster potential at the BCG, but uses the higher dispersion estimate of 1340 km s−1 (Struble & Rood 1999). Models 3 and 4 are the same as models 1 and 2, but fix the cluster potential center at the X-ray center (while of course keeping the BCG and SRG model components fixed). Within each model, we also explored a range of core softening radii. Our primary constraints for the lens models are the ra- dius, size, and orientation of galaxy A. The modest dis- tortion and lack of a counterarc suggests a source position that lies outside the tangential and radial caustics (e.g., Grossman & Narayan 1988) and whose image lies out- side the critical curve. These constraints quickly rule out model 2 in most forms: with a small core radius, the criti- cal curve for z = 0.142 lies at around 35 arcsec, well outside the galaxy radius. This would imply multiple images and large distortions for galaxy A. Even with a large core ra- dius, the effective radial magnification becomes large and can be ruled out by the lens appearance. Model 3 does not produce significant magnification or distortion for the galaxy; thus under such a model, the arc-like shape would need to be intrinsic. Model 4 produces shear and magni- fication at the position of galaxy A, but the shear direc- tion is not directed tangential to the BCG, as observed in galaxy A. Model 1, however, produces modest magnification and distortion of a z = 0.142 source with a shear aligned tan- gentially to the BCG/cluster. A family of models with a range of core radii and asymptotic dispersions for the pri- Blakeslee, Metzger, Kuntschner, & Cot´e 7 potential center is taken to coincide with the peak of the overall X-ray emission. However, if the potential center is taken to be the A2152 BCG, then the models can re- produce the observed position and apparent bowed shape of the galaxy, and the implied magnification is a factor of about 1.9. Conversely, we can say that if the object is sig- nificantly lensed, then the 2.′1 offset of the X-ray centroid from the BCG is due to contamination from another source of X-ray emission. We note that both the cluster veloc- ity dispersion and the core softening radius in our lensing model is degenerate with the source position. However, the higher values that have been reported for the disper- sion (see §2), if centered on the BCG, produce lens models with critical curves that lie outside of the lensed galaxy, thus multiple images would be expected of any background galaxy at the position and redshift of A. We can therefore rule out this case from a lensing argument; even a large core radius model would produce significantly more dis- tortion in a z = 0.142 galaxy. The available redshift data provide evidence for a back- ground association of early-type galaxies at z ≈ 0.134, projected near the center of A2152. Two of the galaxies, which we call G1/G2, have redshifts from the Barmby & Huchra (1998) catalogue. They are separated by 30′′ and are each about 70% more luminous than the A2152 BCG. The third galaxy, called G3, is 35′′ southeast of the A2152 BCG and was picked up by the LRIS slit during our obser- vations. It is 116′′ from G1 and 132′′ from G2, or metric distances of 188-215 h−1 kpc. The velocity range for these three galaxies is ∼ 830 km s−1. While three galaxies do not a cluster make, the luminosities, velocities, and early- type morphologies of the galaxies are highly suggestive. Moreover, there is a strong concentration of small galaxies of unknown redshift centered near the bright G1/G2 pair. It is highly likely that some of the X-ray emission, which is centered just 20′′ from G2, could be due to a grouping of galaxies associated with G1/G2. Therefore, the cen- ter of the true A2152 X-ray emission, and the center of the A2152 cluster potential, might well coincide with the A2152 BCG, as required by our lensing models. We note that the centers of the outermost X-ray contours in the Jones & Forman map of this field drift away from G1/G2 and closer to the A2152 BCG; the optical Hβ emission we find in the BCG spectrum may be additional circumstan- tial evidence for central location in the cluster potential. However, a high-resolution X-ray map is needed to confirm this hypothesis. If there is a background cluster, galaxy A may well be a member, in which case it would be falling away from us through the cluster with a velocity of ∼ 2500 km s−1. Simulations show that the tidal effects of clusters on in- falling galaxies can warp spiral disks into arc-like shapes (Moore et al. 1999a) and even form giant arcs of tidal de- bris (Calc´aneo-Rold´an et al. 2000). Thus, tidal interaction with a local cluster potential, rather than lensing by the foreground A2152 potential, could be responsible for the distorted appearance of the galaxy, as well as for the ob- served strong star formation. Its location and orientation with respect to the A2152 BCG would then be merely for- tuitous. However, the galaxy in question would have to be caught during its initial pass through the cluster, since gas in cluster galaxies tends to be lost on relatively short time-scales, truncating the star formation (e.g., Balogh Fig. 7. -- Contours of the observed galaxy and the inferred source contours from the simple lens model. The upper image shows contours of the observed surface brightness of the lensed galaxy A, spaced by factors of two. The lower image shows the implied source position and the same surface brightness con- tours on the source plane, i.e. how the galaxy would appear if the foreground cluster were absent. The innermost contours, less than 1′′ in size, are rounded in the observed image from the seeing; therefore, these contours become artificially elon- gated in the reconstructed source image. The coordinates are centered on the BCG. 5. SUMMARY AND DISCUSSION Deep R-band Keck imaging taken under excellent seeing conditions has revealed a curved object 25.′′2 (15 h−1 kpc) from the center of the A2152 BCG, along the major axis. The reddening-corrected R magnitude of this "arc" is mR = 18.55 ± 0.03, and follow-up spectroscopy shows that it lies at a redshift z = 0.1423 ± 0.0001. Its position, ori- entation, and redshift all suggest a lensing explanation for the curiously bowed structure. The object has numerous strong emission lines, and the equivalent width of Hα is 117 ± 6 A, from which we infer a star formation rate of (3 ± 1) h−2 M⊙ yr−1. We have presented spatially resolved spectroscopy of the A2152 BCG and SRG, which are separated by 47′′ and are together catalogued as UGC 10187. The relative velocity of the BCG and SRG is 331 ± 7 km s−1, and both galaxies show significant rotation. The rotational velocities along the observed position angle are ∼ 50 km s−1 and ∼ 120 km s−1 for the BCG and SRG, respectively. Our velocity dispersion profile for the BCG extends to a radius of ∼ 16′′ (9.5 h−1 Mpc), about 60% of the way to the radial position of the lensed galaxy. We have modeled the lensing potential of A2152 using constraints from the measured cluster velocity dispersion of 715 km s−1 and our own stellar velocity dispersion mea- surements for the BCG and SRG. The models cannot pro- duce a significant lensing effect on galaxy A if the A2152 8 Lensing in Hercules et al. 1998; Kodama & Bower 2000). The probability that galaxy A is a member of a cluster at z = 0.13 depends of course on the velocity dispersion of the cluster. A deep redshift survey of this field is needed to discern whether or not there is a very substantial background cluster (as suggested by the luminosities of G1/G2 and the position of the X-ray center in this field), and how likely galaxy A's membership is. Such a redshift survey would need to go at least ∼ 2 mag fainter than the Barmby & Huchra (1998) limit in order to pick up the multitude of small galaxies that lie between the A2152 BCG/SRG pair and the G1/G2 background pair. Although superficially A2152 might have seemed an un- remarkable low-redshift cluster (leaving aside the unlikely velocity dispersions in excess 1300 km s−1 found in the lit- erature), our photometric and spectroscopic observations of its central field have raised a host of intriguing new questions about this cluster and the arc-shaped galaxy projected near its BCG. How much mass is there along this line of sight and how is it distributed? Is the A2152 BCG actually at the center of the cluster potential and the observed offset with respect to the X-ray center due to the superposed emission of a background cluster at z = 0.13? If so, is the distorted galaxy a member which is being distorted by the hypothetical background cluster, or is it a field galaxy being lensed by A2152? Along with high- resolution X-ray observations and a deep redshift survey, a weak lensing analysis of this field would help in unraveling the projected mass distribution and provide insight into these questions. We believe that the weight of the evidence favors the view that the A2152 BCG is at the center of its cluster's potential and is significantly lensing galaxy A. Further lensed objects, if present near critical curves, could pro- vide useful future constraints on the central structure of rich clusters. For now, however, a complete understand- ing of this system and definitive answers to the questions raised by this study must await an infusion of new data. We thank Michael Balogh, Alastair Edge, Ben Moore, Ian Smail, and Graham Smith for helpful conversations. This work made use of the NASA/IPAC Extragalactic Database (NED), operated by the Jet Propulsion Labora- tory at Caltech under contract with the National Aeronau- tics and Space Administration. It also made use of Starlink computer facilities. We are grateful to the team of scien- tists and engineers responsible for producing the Low Res- olution Imaging Spectrograph. J.P.B. and P.C. thank the Sherman Fairchild Foundation for support while at Cal- tech. M.R.M's research was supported by Caltech. J.P.B. and H.K. were supported at the University of Durham by a PPARC rolling grant in Extragalactic Astronomy and Cosmology. Note added in proof -- New multi-color data confirm the presence of a rich background cluster around the z = 0.13 G1/G2 galaxy pair; analysis of these data will be presented in a forthcoming paper. REFERENCES Abell, G. O. 1961, AJ, 66, 607 Abell, G. O., Corwin H. G. & Olowin, R. P. 1989, ApJS, 70, 1 Allen, S. W., Fabian, A. C. and Kneib, J. P. 1996, MNRAS, 279, 615 Balogh, M. L., Schade, D., Morris, S. L., Yee, H. K. C., Carlberg, R. G. & Ellingson, E. 1998, ApJ, 504, L75 Barmby, P. & Huchra, J. P. 1998, ApJ, 115, 6 Bender, R. 1990, A&A, 229, 441 Bird, C. M., Davis, D. S., & Beers, T. C. 1995, AJ, 109, 920 Blakeslee, J. P. 1999, AJ, 118, 1506 Blakeslee, J. P. & Metzger, M. R. 1999, ApJ, 513, 592 Blandford, R. D. & Kochanek, C. S. 1987, ApJ, 321, 658 Calc´aneo-Rold´an, C., Moore, B., Bland-Hawthorn, J., Malin, D. & Sadler, E. M. 2000, MNRAS, 314, 324 Campusano, L. D., Kneib, J.-P., & Hardy, E. 1998, ApJ, 496, L79 Cardelli, J. A., Clayton, G. C. & Mathis, J. S. 1989, ApJ, 345, 245 Crawford, C. S., Allen, S. W., Ebeling, H., Edge, A. C. & Fabian, A. C. 1999, MNRAS, 306, 857 Cypriano, E. S., Sodr´e, L., Campusano, L. E., Kneib, J.-P., Giovanelli, R., Haynes, M. P., Dale, D. A., & Hardy, E. 2000, AJ, submitted (astro-ph/0005200) Davoust, E. & Consid`ere, S. 1995, A&AS, 110, 19 Dubinski, J. and Carlberg, R. G. 1991, ApJ, 378, 496 Fukugita, M., Shimasaku, K., & Ichikawa T. 1995, PASP, 107, 945 Grossman, S. A. & Narayan R. 1988, ApJ, 324, L37 Jones, C. & Forman, W. 1999, ApJ, 511, 65 Kennicutt, R. C. 1992, ApJ, 388, 310 Kennicutt, R. C. 1998, ARA&A, 36, 189 Kodama, T. & Bower, R. G. 2000, MNRAS, submitted (astro- ph/0005397) Landolt, A. U. 1992, AJ, 104, 340 LeF`evre, O., Hammer, F., Angonin, M. C., Gioia, I. M., & Luppino, G. A. 1994, ApJ, 422, L5 Lin, H., Kirshner, R. P., Shectman, S. A., Landy, S. D., Oemler, A., 527, 535 Tucker, D. L., & Schechter, P. L., 1996, ApJ, 464, 60 Mellier, Y., Fort, B., & Kneib, J.-P. 1993, ApJ, 407, 33 Miralda-Escud´e, J. 1995, ApJ, 438, 514 Moore, B., Governato, F., Quinn, T., Stadel, J. & Lake, G. 1998, ApJ, 499, L5 Moore, B., Lake, G., Quinn, T. & Stadel, J. 1999a, MNRAS, 304, 465 Moore, B., Quinn, T., Governato, F., Stadel, J. & Lake, G. 1999b, MNRAS, 310, 1147 Navarro, J. F., Frenk, C. S., & White, S. D. M. 1996, ApJ, 462, 563 Navarro, J. F., Frenk, C. S. and White, S. D. M. 1997, ApJ, 490, 493 Oegerle, W. R. & Hoessel, J. G. 1991, ApJ, 375, 15 Oke, J. B., Cohen, J. G., Carr, M., Cromer, J., Dingizian, A., Harris, F. H., Labrecque, S., Lucinio, R., Schaal, W., Epps, H., & Miller, J. 1995, PASP, 107, 307 Postman, M. Huchra, J. P., & Geller, M. J. 1992, ApJ, 384, 404 Postman, M. & Lauer, T. R. 1995, ApJ, 440, 28 Schlegel, D., Finkbeiner, D., & Davis, M. 1998, ApJ, 500, 525 Shapley, H. 1934, MNRAS, 94, 53 Smail, I., Hogg, D. W., Blandford, R., Cohen, J. G., Edge, A. C., & Djorgovski, S. G. 1995, MNRAS, 277, 1 Struble, M. F. & Rood, H. J. 1987, ApJS, 63, 555 Struble, M. F. & Rood, H. J. 1999, ApJS, 125, 35 Tarenghi, M., Chincarini, G., Rood, H. J., & Thompson, L. A. 1980, ApJ, 235, 724 Tarenghi, M., Tifft, W. G., Chincarini, G., Rood, H. J. & Thompson, L. A. 1979, ApJ, 234, 793 Tyson, J. A., Kochanski, G. P., & Dell'Antonio, I. P. 1998 ApJ, 498 L107 Waxman, E. & Miralda-Escude, J. 1995, ApJ, 451, 451 Wegner, G., Colless, M., Saglia, R. P., McMahan, R. K., Davies, R. L., Burstein, D. & Baggley, G. 1999, MNRAS, 305, 259 White, D. A., Jones, C., & Forman, W. 1997, MNRAS, 292, 419 Williams, L. L. R., Navarro, J. F., & Bartelmann, M. 1999, ApJ, Zabludoff, A. I., Geller, M. J., Huchra, J. P., & Ramella, M. 1993, AJ, 106, 1301 Blakeslee, Metzger, Kuntschner, & Cot´e 9 Table 1 Summary of Measurements Object quantity value Arclet G3 Arclet BCG SRG BCG SRG Arclet BCG SRG G3 BCG SRG BCG SRG (a) (a) mR mR fwhm(b) (c) (c) ǫℓ ǫℓ PA(d) PA(d) cz(e) cz(e) cz(e) cz(e) (f) vm vm σ0 σ0 (g) (g) (f) 18.55 19.24 0.′′42 0.10 0.33 10.◦7 161.◦4 42665 13188 13519 39738 49 117 295 271 ± 0.03 0.05 0.′′08 0.01 0.02 0.◦7 2.◦0 40 25 25 25 5 7 7 8 (a)R magnitude, corrected for Galactic extinction (b)seeing-corrected FWHM along the minor axis of arclet (c)ellipticity of the light distribution, at r = 10′′ (d)angle of major axis, east from north, at r = 10′′ (e)heliocentric redshift (km s−1) (f)rotational velocity (km s−1) at r > 5′′ and PA = 122◦ ∼ (g)central velocity dispersion (km s−1) fig1.gif Fig. 1. -- An A2152 finding chart. The image was taken with Keck/LRIS through an R filter. The field is 5.′6 × 6.′0 in size, and the bright galaxy at center is the A2152 BCG. Several other galaxies are labeled: the second-ranked cluster galaxy (SRG); the arclet at z = 0.142 (A); three early-type galaxies all at z ≈ 0.133 (G1, G2, G3); and the only other galaxy in this field with a known redshift (M). The redshifts of the BCG, SRG, and galaxy "M" are all z ≈ 0.044. For reference, the arclet is 25′′ from the center of the BCG, and the BCG/SRG pair are separated by 47′′. The large "X" near the bright pair of background galaxies G1/G2 marks the location of the centroid of the X-ray emission, normally attributed to A2152 (Jones & Forman 1999). This figure "fig1.gif" is available in "gif"(cid:10) format from: http://arxiv.org/ps/astro-ph/0010187v1
0710.3911
1
0710
2007-10-21T09:06:10
Helium stars as supernova progenitors
[ "astro-ph" ]
We follow the evolution of helium stars of initial mass $(2.2 - 2.5) M_\odot$, and show that they undergo off-center carbon burning, which leaves behind ${\mathbf \sim 0.01 M_\odot}$ of unburnt carbon in the inner part of the core. When the carbon-oxygen core grows to Chandrasekhar mass, the amount of left-over carbon is sufficient to ignite thermonuclear runaway. At the moment of explosion, the star will possess an envelope of several $0.1 M_{\odot}$, consisting of He, C, and possibly some H, perhaps producing a kind of peculiar SN. Based on the results of Waldman and Barkat (2007) for accreting white dwarfs, we expect to get thermonuclear runaway at a broad range of $\rho_c \approx (1 - 6) \times 10^9 \mathrm{g cm^{-3}}$, depending on the amount of residual carbon. We verified the feasibility of this scenario by showing that in a close binary system with initial masses $(8.5 + 7.7) M_{\odot}$ and initial period of 150 day the primary produces a helium remnant of $2.3 M_{\odot}$ that evolves further like the model we considered.
astro-ph
astro-ph
**FULL TITLE** ASP Conference Series, Vol. **VOLUME**, **YEAR OF PUBLICATION** **NAMES OF EDITORS** Helium stars as supernova progenitors Roni Waldman 1, Lev R. Yungelson 2 and Zalman Barkat 1 Abstract. We follow the evolution of helium stars of initial mass (2.2 − 2.5) M⊙, and show that they undergo off-center carbon burning, which leaves behind ∼0.01 M⊙ of unburnt carbon in the inner part of the core. When the carbon-oxygen core grows to Chandrasekhar mass, the amount of left-over car- bon is sufficient to ignite thermonuclear runaway. At the moment of explosion, the star will possess an envelope of several 0.1 M⊙, consisting of He, C, and possibly some H, perhaps producing a kind of peculiar SN. Based on the results of Waldman & Barkat (2007) for accreting white dwarfs, we expect to get ther- monuclear runaway at a broad range of ρc ≈ (1 − 6) × 109g cm−3, depending on the amount of residual carbon. We verified the feasibility of this scenario by showing that in a close binary system with initial masses (8.5 + 7.7) M⊙ and initial period of 150 day the primary produces a helium remnant of 2.3 M⊙ that evolves further like the model we considered. 1. Introduction Type Ia supernovae (SN Ia) have a relatively small dispersion of luminosity (the standard deviation in peak blue luminosity is σB ≈ 0.4−0.5 mag., Branch & Miller (1993)) and are being used as distance indicators ("standard candles"), having especial significance in the effort of determining the cosmological parameters of our universe. The long-standing explanation of the SN Ia phenomenon has been the ex- plosive burning of degenerate carbon in the core of a carbon-oxygen white dwarf, which becomes unstable as it grows to Chandrasekhar mass (MCh) either by ac- cretion from a binary companion or by a merger of two white dwarfs, following the angular momentum loss from the system by gravitational wave radiation. However, theoretical models are still far from self-consistently producing an evo- lutionary path towards the progenitor and reproducing crucial features of the observational data, such as the composition of the ejecta. For a detailed re- view of the above see, e.g., Leibundgut (2000); Hillebrandt & Niemeyer (2000); Filippenko (2005). As well, SN Ia can not be regarded as perfectly homogeneous class, since their Hubble diagram exhibits scatter larger than the photometric errors, while spectroscopic and photometric peculiarities have been noted with increasing fre- quency in well-observed SN Ia (e.g., Filippenko 2005). Therefore, there is an obvious need for progenitor scenarios that could ex- plain the diversity among SN Ia. Several explanations have been suggested, such 1Racah Institute of Physics, The Hebrew University, Jerusalem 91904, Israel 2Institute of Astronomy, Russian Academy of Sciences, 48 Pyatnitskaya Str, Moscow, Russia 1 2 Roni Waldman, Lev R. Yungelson and Zalman Barkat as variations in the metallicity of the progenitor, in the carbon to oxygen ratio at its center, or in the central density at the time of ignition (e.g., Timmes et al. 2003; Ropke et al. 2006; Lesaffre et al. 2006). The variation of the latter two parameters is expected to result from the variation in the initial white dwarf mass and in the accretion history. In this work we follow the evolution of helium stars with initial mass ≈ (2.2 − 2.5) M⊙ and show that they might reach thermonuclear explosion and perhaps account for some of the peculiar SNe. 2. Results We followed in detail the evolution of a 2.4 M⊙ helium star, starting from a homogeneous object. We used the TYCHO evolutionary code described in Young & Arnett (2005). The helium burning convective core has an almost constant mass of about 0.82 M⊙, and produces a CO core with XC ≈ 0.27. Subsequently, a radiative helium burning shell develops above the core, and as the core grows to about 1.1 M⊙ carbon is ignited in the core at about 0.3 M⊙ (Fig. 1, left panel). After the end of core helium burning the center of the star becomes increasingly degenerate, and neutrino cooling is increasingly compet- ing with contraction-induced heating. As a result, the maximum temperature is attained at an off-center point and it keeps growing until carbon ignites there (Fig. 1, right panel). Following off-center C-ignition, the center expands and cools, and a series of carbon burning flashes occurs. Eventually burning ceases when carbon is almost exhausted in the core, however as can be seen in Fig. 2, the innermost 0.5 M⊙ of the core has a residual XC ≈ 0.02. After carbon burning ceases, the center continues to heat and contract while the helium burning shell gradually increases the mass of the core. Similarly to AGB evolution, the luminosity of the star grows, causing the envelope to ex- pand and to develop a deepening convective region, which at certain moment penetrates into the helium burning shell and creates conditions for hot bottom burning. During this stage the carbon that accumulates below the helium burn- ing shell quasi-periodically ignites (see Fig. 1, left panel), similarly to the helium flashes occurring in AGB stars. Since very little is known about the mass loss rate of stars of the kind we consider, we applied Reimers (1975)-based rate, which reduced the mass of the star to ≃ 1.8 M⊙ (Fig. 1, left panel). Finally, the core grows to MCh due to the helium burning shell and carbon in the center ignites (Fig. 1, right panel). Convection is initiated, supplying more carbon to the central region, and if the amount of carbon is sufficient to raise the temperature above the oxygen ignition threshold (≃ 1.5×109 K), which indeed happens in this case, oxygen will ignite under degenerate conditions and initiate an explosion similar to the classical SN Ia case. Models with initial masses (2.2−2.5) M⊙ evolve very similarly to the 2.4 M⊙ case, developing a carbon residue which might ignite thermonuclear runaway. We did not follow the evolution further through the hot-bottom burning stage. The 2.1 M⊙ model only ignites carbon at a later stage, below the helium burning shell, similar to the carbon flashes encountered in the 2.4 M⊙ case, however, since in this case carbon has not been previously depleted in the core, burning will probably continue until it reaches the center. We have not followed this Helium stars as supernova progenitors 3 ignition line Xc=0.25 M=2.4 Msun profile before ignition profile before ignition profile at ignition 1010 109 108 107 106 ) 3 m c / r g ( y t i s n e D 105 0 1 2 4 3 5 Temperature (108 K) 6 7 8 Figure 1. Evolution of the initially 2.4 M⊙ helium star model. Left: Kip- penhahn plot after the end of core helium burning. Light (red) filled areas are convective, dark (blue) line shows maximum thermonuclear energy pro- duction rate. Right: Evolution of stellar center on temperature vs. density plane (long dash (green) line); also shown is the ρ − T relation of the inner part of the model prior (short dashed (dark blue) and dotted (purple) lines) and immediately after off-center carbon ignition (dot-dashed (light blue) line). Carbon ignition line for XC = 0.25 is shown as solid (red) line. Figure 2. thermonuclear runaway. Composition of the initially 2.4 M⊙ helium star model before model further. The models of 2.6 M⊙ and more massive ones ignite carbon very close to or at the center, so that the carbon residue is either non-existent or insufficient for thermonuclear runaway. 4 Roni Waldman, Lev R. Yungelson and Zalman Barkat In an earlier work (Waldman & Barkat 2007) we explored the similar case of CO white dwarfs undergoing off-center carbon burning followed by mass ac- cretion. As a function of the amount of residual carbon we got thermonuclear runaway at a broad range of ρc ≈ (1 − 6) × 109g cm−3. Since the structure of the CO core in our helium star models is very similar to that of the mass accreting CO white dwarfs, we expect to get a similar result at runaway. Supernova which results from thermonuclear runaway in the remnant of He-star most probably will not differ photometrically from a "normal" SN Ia, but one may expect presence of strong He-lines in the spectrum, thanks to thick He-mantle of pre-SN, making this SN Ia "peculiar" (N. Chugai, priv. comm.). To complete the picture, we tested whether a hydrogen-deficient star very similar to our initial models could be created as a result of close binary evolution. We begun with a binary of (8.5 + 7.7) M⊙, with an orbital period of 150 day. After the stage of Roche lobe overflow, the primary has a total mass of 2.3 M⊙, a CO core of 1.2 M⊙, and a surface hydrogen mass fraction of 0.14. Later, wind mass loss and sporadic RLOF reduce the mass to 2.1 M⊙. Further evolution of the remnant is similar to the above described. To summarize, we showed that evolution of helium stars with initial mass about (2.2 − 2.5) M⊙, which might be produced in close binaries, may suggest a SN scenario in which thermonuclear runaway in MCh-mass cores is initiated by a very small amount of residual carbon, while the stars still have a several 0.1 M⊙ envelope consisting mostly of helium, carbon and possibly some hydrogen. In order to give a well justified statement on the observational outcome, our pre- SN models should be used for detailed simulations of the explosive runaway and spectra modeling. Acknowledgments. This research was supported in part by the National Science Foundation under grant PHY05-51164, Russian Academy of Sciences Basic Research Program "Origin and Evolution of Stars and Galaxies" and by Russian Foundation for Basic Research grant 07-02-00454. LRY acknowledges N.N. Chugai for fruitful discussions and P.P. Eggleton for providing his evolu- tionary code and advise on its usage. RW acknowledges D. Arnett for providing his TYCHO evolutionary code for public use. References Branch D., Miller D. L., 1993, ApJ 405, L5 Filippenko A. V., 2005, in E. M. Sion, S. Vennes, H. L. Shipman (eds.), White dwarfs: cosmological and galactic probes, Vol. 332 of Astrophysics and Space Science Library, 97 Hillebrandt W., Niemeyer J. C., 2000, ARA&A 38, 191 Leibundgut B., 2000, A&A Rev. 10, 179 Lesaffre P., Han Z., Tout C. A., Podsiadlowski P., Martin R. G., 2006, MNRAS 368, 187 Reimers D., 1975, Memoires of the Societe Royale des Sciences de Liege 8, 369 Ropke F. K., Gieseler M., Reinecke M., Travaglio C., Hillebrandt W., 2006, A&A 453, 203 Timmes F. X., Brown E. F., Truran J. W., 2003, ApJ 590, L83 Waldman R., Barkat Z., 2007, ApJ 665, 1235 Young P. A., Arnett D., 2005, ApJ 618, 908
astro-ph/0201094
1
0201
2002-01-07T16:49:55
How is the Reionization Epoch Defined?
[ "astro-ph" ]
We study the effect of a prolonged epoch of reionization on the angular power spectrum of the Cosmic Microwave Background. Typically reionization studies assume a sudden phase transition, with the intergalactic gas moving from a fully neutral to a fully ionized state at a fixed redshift. Such models are at odds, however, with detailed investigations of reionization, which favor a more extended transition. We have modified the code CMBFAST to allow the treatment of more realistic reionization histories and applied it to data obtained from numerical simulations of reionization. We show that the prompt reionization assumed by CMBFAST in its original form heavily contaminates any constraint derived on the reionization redshift. We find, however, that prompt reionization models give a reasonable estimate of the epoch at which the mean cosmic ionization fraction was ~50%, and provide a very good measure of the overall Thomson optical depth. The overall differences in the temperature (polarization) angular power spectra between prompt and extended models with equal optical depths are less than 1% (10%).
astro-ph
astro-ph
Mon. Not. R. Astron. Soc. 000, 1 -- 5 (2001) Printed 27 October 2018 (MN LATEX style file v1.4) How is the Reionization Epoch Defined ? Marialuce Bruscoli1, Andrea Ferrara2 & Evan Scannapieco2 1Dipartimento di Astronomia, Universit`a degli studi di Firenze, L.go E. Fermi 2, Firenze, Italy 2Osservatorio Astrofisico di Arcetri, L.go E. Fermi 5, Firenze, Italy 27 October 2018 ABSTRACT We study the effect of a prolonged epoch of reionization on the angular power spec- trum of the Cosmic Microwave Background. Typically reionization studies assume a sudden phase transition, with the intergalactic gas moving from a fully neutral to a fully ionized state at a fixed redshift. Such models are at odds, however, with detailed investigations of reionization, which favor a more extended transition. We have mod- ified the code CMBFAST to allow the treatment of more realistic reionization histories and applied it to data obtained from numerical simulations of reionization. We show that the prompt reionization assumed by CMBFAST in its original form heavily con- taminates any constraint derived on the reionization redshift. We find, however, that prompt reionization models give a reasonable estimate of the epoch at which the mean cosmic ionization fraction was ≈ 50%, and provide a very good measure of the overall Thomson optical depth. The overall differences in the temperature (polarization) an- gular power spectra between prompt and extended models with equal optical depths are less than 1% (10%). 1 INTRODUCTION At a redshift of z ≈ 1100 the intergalactic medium (IGM) recombined and remained neutral until the first sources of ionizing radiation formed. While the distribution and evo- lution of these sources are unknown, the overall process of IGM reionization is fairly well understood, and can be di- vided into three phases. First, individual HII regions devel- oped around the sources. Then, these ionized regions grew in number and size until they overlapped, producing a sud- den increase of the photon mean free path. Finally, when all underdense regions and voids were completely ionized, pho- tons penetrated into overdense clumps and filaments, bring- ing the reionization process to completion at a "reionization redshift." It is clear that the mystery of the nature and evolution of the ionizing sources is intimately tied with the time scale over which these phases took place. Thus there is a great deal of physical information that could be gathered if the reionization redshift, zi, and its duration, ∆z, can be firmly established. Apart from the study of the (HI and HeII) Gunn- Peterson effect, which has not yet yielded conclusive re- sults (Becker et al. 2001; Gnedin 2001; Theuns et al. 2001), measurements of Cosmic Microwave Background (CMB) anisotropies have the greatest potential for constraining these important quantities (Griffiths et al. 1999). The scat- tering of CMB photons by free electrons damps the angu- lar power spectrum of primary anisotropies by a factor of e−2τ for large angular multipoles ℓ > ∼ 100 (Tegmark & Zal- darriaga 2000), where τ is the Thomson optical depth. Us- c(cid:13) 2001 RAS ing the currently available data, consistency with the lack of Gunn-Peterson trough and with the observed peak in the angular power spectrum at ℓ ≈ 200 (De Bernardis et al. 2000; Hanany et al. 2000; Padin et al. 2001) is able to constrain 0.02 ≤ τ ≤ 0.44 (De Bernardis et al. 1997; Grif- fiths et al. 1999; Tegmark et al. 2001; Griffiths & Liddle 2001), although these results are somewhat dependent on the cosmological model assumed. Much better constraints are expected from polarization studies to be carried out by future satellites such as SPOrt⋆ and PLANCK†. A second method of quantifying reionization is to ex- amine zi, and several recent studies have attempted to use the CMB to derive this redshift directly, using codes that as- sume a sudden epoch of reionization (e.g, Hu et al. 95; Seljak & Zaldarriaga 1996). For example, Schmalzing et al. (2000) use MAXIMA data in combination with cosmological pa- rameters from independent measurements of Big Bang Nu- cleosynthesis and X-ray cluster data to constrain zi. By per- forming a χ2 analysis, they conclude that zi > 15(8) at the 68% (95%) confidence level. Similarly, Naselsky et al. (2001), in addition to providing support to Schmalzing et al. results, showed that polarization spectra are very sensitive to the reionization redshift. An important caveat is implicit in these applications however, namely the assumption of prompt reionization. While taking ∆z = 0 is a reasonable first step, more de- ⋆ http://sport.tesre.bo.cnr.it † http://astro.estec.esa.nl/SA-general/Projects/Planck 2 M. Bruscoli, A. Ferrara & E. Scannapieco tailed simulations show that this approach paints a picture of reionization in only the broadest of strokes. This is par- ticularly true as the onset of reionization raises the overall temperature of the IGM, suppressing the further formation of objects that are too weakly bound gravitationally to over- come the drastic increase in thermal pressure (Barkana & Loeb 1999). Thus recent numerical simulations of reioniza- tion (Ciardi et al. 2000, hereafter CFGJ; Bruscoli et al. 2000; Gnedin 2000; Benson et al. 2000) have shown that the evolu- tion of the mean ionization fraction is slow, has a nonlinear dependence on redshift, and occurs in an extremely patchy manner. All of these details may considerably affect the de- termination of zi. In this paper, we aim to clarify the effects of realistic reionization scenarios on the observables typically used to examine the reionization epoch. To this end we have modi- fied the most heavily relied on theoretical code for comput- ing CMB fluctuations, CMBFAST (Seljak & Zaldarriaga 1996), to allow the treatment of such reionization histories, helping to clarify the best approach in trying to quantify and define the epoch of reionization. The structure of this work is as follows. In §2 we dis- cuss numerical models of reionization and how these have been incorporated into the CMBFAST code. In §3 we apply these techniques to study the impact of these models on the temperature and polarization spectrum of the CMB, and conclusions are given in §4. 2 REIONIZATION HISTORY FROM SIMULATIONS To assess the effects of more physical reionization histories on the determination of zi, we have modified the Boltzmann code CMBFAST such that the redshift evolution of the volume- averaged mean hydrogen ionization fraction, xe(z), can be specified, reading in data from more detailed simulations of reionization.‡ Our modified version of CMBFAST is then able to consider three types of reionization histories: (i) prompt reionization models in which the total optical depth, τ , is chosen (ii) prompt models in which the reionization redshift, zi, is fixed, and (iii) models in which xe(z) is read from an external file. As in Bruscoli et al. (2000), we consider a set of reion- ization simulations studied in CFGJ, which model reioniza- tion from stellar sources, including Population III objects. These simulations were developed in a critical density CDM universe with cosmological parameters h = 0.5, σ8 = 0.6, and Ωb = 0.06 (where h is the Hubble constant in units of 100 km s−1 Mpc−1, σ8 is the rms variation of density pertur- bations at the 8h−1 comoving Mpc length scale, and Ωb is the baryonic density in units of the critical density) but span a large range of model parameters, and thus are suitable for drawing conclusions as to the effect of extended reionization in more general cosmologies. The spatial distribution of the sources and ionized re- gions at various cosmic epochs and the overall evolution of ‡ A copy of this version of CMBFAST is publicly available at http://www.arcetri.astro.it/science/cosmology/CMB xe(z) are obtained from high-resolution cosmological simula- tions within a periodic box of comoving length L = 2.55h−1 Mpc. The source properties are calculated taking into ac- count a self-consistent treatment of both radiative feedback from ionizing and H2 -- photodissociating photons and stel- lar feedback regulated by supernovae from massive stars. There are two main free parameters in the simulations: the fraction of total baryons converted into stars, fb⋆, and the escape fraction of ionizing photons from a given galaxy, fesc. A critical discussion of these parameters is given in CFGJ, and here we examine four different models, fixing (fb⋆ = 0.012, fesc = 0.2) in run A, (0.004, 0.2) in run B, (0.15, 0.2) in run C , and (0.012, 0.1) in run D. As run A, in which zi ≈ 11, gives the best agreement between the derived evolution of the cosmic star formation rate and observations at z < ∼ 4 (Steidel et al. 1998, CFGJ), we choose this as our fiducial model, and rely on runs B, C, and D to study the effects of model uncertainties on our conclusions. 3 PROMPT VS. SIMULATED REIONIZATION Equipped with this more general tool, we now quantitatively contrast the prompt reionization xe,p(z) with the one de- rived from the simulations, xe,s(z). As CMBFAST can be run either by assigning zi or τ , we are able to examine both these cases separately. 3.1 Assigning the Reionization Redshift In Fig. 1 we plot the redshift evolution of the hydrogen ionization fraction and the Thomson optical depth for the prompt and simulated cases, choosing in the prompt case the same value of zi = 10.9 as deduced from run A. In this plot, the smoother, nonlinear increase of xe,s(z) is evident, with ∆z ≈ 15. Note also that the optical depth for the two reionization histories are quite different, and that τs = 0.08 for run A and τp = 0.04 for prompt reionization. If we define the discrepancy between the prompt and simulated visibility functions g(z) as ∆g g = gs(z) − gp(z) gs(z) , (1) ∼ 1000). where g(z) is the probability that a photon reaching the observer last scattered in the redshift range z and z + dz, then we find that ∆g/g ≈ 5% for a large range of redshifts between recombination and the beginning of reionization (25 < ∼ z < The most important comparison between the prompt and fiducial cases, however, is made in terms of the dif- ferent resulting CMB temperature and polarization angular power spectra, Cℓ and Pℓ, as these are the observable quan- tities. These are shown in the upper left panel of Fig. 2, in which differences are particularly evident in the polarization spectra, which are more sensitive to the reionization history. Reionization introduces a characteristic bump in the lower multipoles (ℓ = 5 − 10) that tends to shift to higher ℓ's when τ is increased. The amplitude of the bump in the Pℓ spec- trum decreases with τ and therefore the signal is weaker for the prompt reionization case. In particular the bump am- plitude in the angular spectrum Pℓ is roughly proportional c(cid:13) 2001 RAS, MNRAS 000, 1 -- 5 How is the Reionization Epoch Defined ? 3 Figure 1. H ionization fraction, xe, and Thomson optical depth, τ , versus redshift for run A (solid lines) and for prompt reioniza- tion (dotted lines) when CMBFAST is run choosing zi = 10.9. The dots represent the output from the simulation. to τ 2 (for τ < (Zaldarriaga 1997; Fabbri 1999). ∼ 1), while its position scales as ℓpeak ∝ z1/2 i To highlight these differences, we plot the normalized discrepancy of temperature and polarization angular spec- tra [defined analogously to eq. (1)] between the models in the lower left panel. Here find that ∆Cℓ/Cℓ < 0.07 while, around ℓ ≈ 10, ∆Pℓ/Pℓ ≈ 1. These results make it clear that although the prompt and simulated reionization his- tories share the same value of zi, they produce noticeably different angular power spectra, and thus correct constraints on the reionization redshift cannot be obtained by equating zi in the prompt and simulated runs. 3.2 Assigning the Optical Depth An alternative approach is to compare our fiducial run with a prompt model that differs in its zi but has the same Thom- son optical depth, τ = 0.08. This results in a value for the prompt reionization redshift of zi,p = 17.1. Note that this value is much larger than the simulated redshift of reioniza- tion zi,s = 10.9, and corresponds to the redshift at which the simulated ionization fraction is only xe,s ≈ 0.5. Comparing these two runs we find, however, that the discrepancy between the two visibility functions is only ∆g/g ∼ 10−4. This small change in the visibility function results in an equally small change in the observed CMB anisotropies. This can be seen on the top right panel of in Fig. 2 in which the angular power spectra are almost in- distinguishable. Indeed, the fractional discrepancy between < these runs is only ∆Cℓ/Cℓ ∼ 0.06 as shown in the bottom right panel in this figure. < ∼ 0.01 and ∆Pℓ/Pℓ Thus the two different reionization models produce comparable angular power spectra if τ is specified rather than zi. In fact if we try to recover the reionization redshift from the standard formula zi = 8.9(τ /hΩb)2/3Ω1/3 (Peebles 1993, Tegmark et al. 2000) we get zi = 17.1, again a redshift at which xe ≈ 0.5. It is clear then that by analyzing CMB 0 c(cid:13) 2001 RAS, MNRAS 000, 1 -- 5 Figure 2. Top left: temperature (Cℓ) and polarization (Pℓ) an- gular power spectra versus the multipole ℓ for run A (solid line) and for prompt reionization (dotted line) when CMBFAST is run with zi = 10.9. Bottom left: the normalized temperature and po- larization discrepancy as a function of angular multipole ℓ for this case. Here the solid and dotted lines refer to Cℓ and Pℓ respec- tively. Top right: temperature (Cℓ) and polarization (Pℓ) angular power spectra versus the multipole ℓ for run A (solid line) and for prompt reionization (dotted line) when CMBFAST is run with τ = 0.08. Bottom right: the normalized temperature and polariza- tion discrepancy as a function of angular multipole ℓ for this case. Here the solid and dotted lines refer to Cℓ and Pℓ respectively. power spectra using a prompt model, one can draw reason- able conclusions as to the overall Thomson optical depth, whereas the reionization redshift is much more uncertain. 3.3 Quantification of Errors & Tests of Our Approach In order to quantify the difference between the fiducial model and the prompt model with the same optical depth further, we plot in Fig. 3 the quantity χ2(ℓmax) = ℓmax X 2 (2ℓ + 1) 2 ( ∆Cℓ Cℓ )2fsky (2) as a function of the multipole number ℓmax assuming full sky coverage, fsky = 1. For multipole values greater than 100, χ2 exceeds unity, the value corresponding to the cos- mic variance error of 2C 2 ℓ /(2ℓ + 1). This means that one can find a statistically significant difference between the prompt and fiducial models only by looking at angular multipoles with ℓ > ∼ 100. However, in Fig. 2 we saw that the largest dif- ferences between the two histories occur at lower multipole values (ℓ = 20 − 40). Thus although these reionization his- 4 M. Bruscoli, A. Ferrara & E. Scannapieco Figure 3. χ2 of Cℓ (solid line) and of Pℓ (dotted line) versus ℓmax (see eq. 2). tories are considerably different, it is almost impossible to discriminate between them observationally. Finally, we extend our analysis to the runs B, C, and D of CFGJ, comparing them with prompt runs with equal values of τs. These results are summarized in Table 1. As in run A, the values of zi,p found for these models correspond to times at which xe,s ≈ 0.5, when the IGM was in the midst of changing from a neutral state to an ionized one. Is is clear from this table that it is not possible to put a constraint on the reionization redshift directly from CMBFAST. In fact, as can be easily seen in Table 1, the differ- ences between zi,s and zi,p cannot be easily parameterized because they depend on several effects that influence the du- ration ∆zs of the reionization process. These effects include radiative and stellar feedbacks, the star formation efficiency, the photon escape fraction, the spatial and mass distribu- tion of the ionizing sources, and even the clumpiness of IGM, which is beyond the scope of the CFGJ simulations. We have conducted a number of checks and convergence tests to assess the robustness of our approach. Our first check was to examine the overall, numerical value of the optical depth, which differs slightly from the input value due to the finite integration time step. Here we found that the output τ s differ by less than 5 × 10−5 from the specified values. ∼ 10−4, i.e. 30 times smaller This corresponds to ∆Cℓ/Cℓ than the actual difference between the simulated and prompt runs. < Next, we carried out a series of convergence tests, to ascertain the effects of k-mode sampling on our results. Here we restrict our tests to run A for which we increased the k resolution by 50%. A comparison between the origi- nal and resampled simulated runs showed that at most an- gular scales ∆Cℓ/Cℓ and ∆Pℓ/Pℓ are within 0.1%; larger differences (up to ∼ 0.7%) are found at higher multipole numbers (ℓ > ∼ 700). Comparing the resampled run A with a resampled prompt run with equal input optical depth, we found ∆Cℓ/Cℓ and ∆Pℓ/Pℓ values that were virtually in- distinguishable from those obtained with low k resolution runs. Finally, we examined the effect of time sampling at both k resolutions, comparing two prompt runs with equal optical depth and k resolution, but forcing one of them to the same time step as in simulated run A. Again, we found ∆Cℓ/Cℓ ∼ 10−4, far smaller than the discrepancies be- and ∆Pℓ/Pℓ tween the prompt and simulated runs. < 4 CONCLUSIONS One of the first stages of nonlinear structure formation, reionization marked an important transition from a dark and relatively simple universe to one filled with a dazzling array of stars, galaxies, quasars, and other nonlinear ob- jects. And although one of our best probes of this transition is through the measurement of CMB fluctuations, the pro- cess of reionization itself is much more dependent on the complicated astrophysical issues important at low redshifts than the linear issues important at z ≈ 1100. In this work, we have explored this transition, quan- tifying the impact of realistic scenarios of reionization on the angular power spectrum of the Comic Microwave Back- ground. While standard estimates assume prompt reioniza- tion, we have considered instead a range of simulated mod- els, each with a prolonged reionization epoch. We find that equating the redshift of full IGM ionization between these simulations and models that assume instantaneous reioniza- tion leads to widely discrepant temperature and polarization spectra. On the other hand, equating prompt and extended models with the same overall optical depth leads to differ- ences in anisotropies that are nearly undetectable. In this case the redshift of complete ionization is lost in the compli- cated details of the phase transition, and comparisons yield zi values corresponding to roughly the point of 50% ioniza- tion in the simulations, although even this value is model de- pendent. It is clear then that while zi is useful as schematic tool, it is the total optical depth that is most accurate in providing a definition of the reionization epoch. MB thanks M. Zaldarriaga for discussions and hospi- tality at IAS, Princeton where part of this work has been carried out. ES has been supported in part by an NSF MPS- DRF fellowship. REFERENCES Ciardi, B., Ferrara, A. & Abel, T. 2000, ApJ, 533, 594 Ciardi, B., Ferrara, A., Governato, F. & Jenkins, A. 2000, MN- RAS, 314, 611 (CFGJ) Barkana, R. & Loeb, A. 1999, ApJ, 523, 54 Becker, R. H. et al. 2001, astro-ph/0108097 De Bernardis, P. et al. 2000, Nature, 404, 955 De Bernardis, P., Balbi, A., De Gasperis, G., Melchiorri, A. & Vittorio, N. 1997, ApJ, 480, 1 Fabbri, R., 1999, New Astronomy Reviews, 43, 215 Gnedin, N.Y., 2001, astro-ph/0110290 Griffiths, L. M. & Liddle, A. R. 2001, MNRAS, 324, 769 Griffiths, L.M., Barbosa, D. & Liddle, A.R. 1999, MNRAS, 308, 854 Hanany, S. et al. 2000, ApJ, 545, L5 Hu, W. 2000, ApJ, 529, 12 Hu, W., Scott, D., Sugiyama, N. & White, M. 1995, Phys. Rev. D, 52, 5498 c(cid:13) 2001 RAS, MNRAS 000, 1 -- 5 How is the Reionization Epoch Defined ? 5 Table 1: Reionization parameters Run ∆zs (a) (b) zi,s (c) τs (d) zi,p xe,s(zi,p)(e) A B C D 15 17 15 17 10.9 0.080 17.1 8.3 15 8.3 0.059 13.9 0.098 19.8 0.064 14.7 0.46 0.51 0.48 0.47 ∆Cℓ Cℓ ∆Pℓ Pℓ < ∼ 0.01 < ∼ 0.06 < ∼ 0.007 < ∼ 0.01 < ∼ 0.008 < ∼ 0.1 < ∼ 0.1 < ∼ 0.1 Table 1. (a)Duration of reionization; (b)Redshift of complete reionization (simulated); (c)Thomson Optical Depth (simulated); (d)Redshift of complete reionization (prompt); (e)Hydrogen ion- ization fraction at zi,p (simulated). Naselsky, P. et al. 2001, astro-ph/0102378 Padin, S. et al. 2001, ApJ, 549, L1 Peebles, P.J.E. 1993, Principles of Physical Cosmology, Princeton, Princeton University Press Tegmark, M., Zaldarriaga, M. & Hamilton, A.J.S. 2001, Phys. Rev. D, 43007 Tegmark, M. & Zaldarriaga, M. 2000, ApJ, 544, 30 Theuns, T., Zaroubi, S. & Kim, T.S. 2001, astro-ph/0110552 Schmalzing, J., Sommer-Larsen, J. & Goetz, M. 2000, astro- ph/0010063 Seljak, U. & Zaldarriaga, M. 1996, ApJ, 469, 437 Steidel, C.C. et al. 1999, ApJ, 519, 1 Zaldarriaga, M., 1997, Phys. Rev. D, 55, 1822 c(cid:13) 2001 RAS, MNRAS 000, 1 -- 5
astro-ph/9509120
1
9509
1995-09-21T18:28:30
Optical Identification of Quasar 0917+7122 in the Direction of an Extreme-Ultraviolet Source
[ "astro-ph" ]
We report the optical identification of an $R=18.3$ mag, $z=2.432$ quasar at the position of a 6 cm radio source and a faint \rosat PSPC X-ray source. The quasar lies within the error circles of unidentified extreme-UV (EUV) detections by the \euve and \rosat WFC all-sky surveys at $\sim 400$ \AA\ and $\sim 150$ \AA, respectively. A 21 cm H I emission measurement in the direction of the quasar with a $21'$-diameter beam yields a total H I column density of $N_{H}=3.3\times 10^{20}$ cm$^{-2}$, two orders of magnitude higher than the maximum allowed for transparency through the Galaxy in the EUV. The source of the EUV flux is therefore probably nearby ($\ltorder 100$ pc), and unrelated to the quasar.
astro-ph
astro-ph
Optical Identi(cid:12)cation of Quasar + in the Direction of an Extreme-Ultraviolet Source Dan Maoz, Eran O. Ofek, Amotz Shemi School of Physics & Astronomy and Wise observatory Tel-Aviv University, Tel-Aviv  , ISRAEL Aaron J. Barth, Alexei V. Filippenko Department of Astronomy University of California, Berkeley, CA , USA M. S. Brotherton, Beverley J. Wills, D. Wills McDonald Observatory and Astronomy Department University of Texas, Austin, TX , USA Felix J. Lockman National Radio Astronomy Observatory P.O. Box , Green Bank, WV  -, USA to appear in Astronomy & Astrophysics (Research Note) 5 9 9 1 p e S 1 2 0 2 1 9 0 5 9 / h p - o r t s a  Abstract We report the optical identi(cid:12)cation of an R = : mag, z = : quasar at the position of a  cm radio source and a faint ROSAT PSPC X-ray source. The quasar lies within the error circles of unidenti(cid:12)ed extreme-UV (EUV) detections by the EUVE and ROSAT WFC all-sky surveys at (cid:24)  A and (cid:24)  A, respectively. A  cm (cid:23) (cid:23) H I emission measurement in the direction of the quasar with a  -diameter beam yields a total H I column density of N = : (cid:2)  cm , two orders of magnitude H  (cid:0) higher than the maximum allowed for transparency through the Galaxy in the EUV. The source of the EUV (cid:13)ux is therefore probably nearby (  pc), and unrelated < (cid:24) to the quasar.  Introduction The ROSAT Wide Field Camera (WFC) and the Extreme Ultraviolet Explorer (EUVE) missions carried out all-sky surveys at extreme-ultraviolet (EUV) wavelengths in  -  and  - , respectively (Pounds et al.  ; Malina et al.  ; Bowyer et al.  ). The EUVE all-sky survey was carried out in four broad bands, centered at approximately  A,  A,  A, and  A. The ROSAT WFC all-sky survey was (cid:23) (cid:23) (cid:23) (cid:23) conducted in two EUV bands, centered at about  A and  A, and similar to the (cid:23) (cid:23)  A and  A bands in the EUVE survey. The two experiments had comparable (cid:23) (cid:23) sensitivities in their common bands. While the ma jority of the EUV sources detected by the two missions have been identi(cid:12)ed, mostly with white dwarfs and active late-type stars, about % of the EUV sources still have no identi(cid:12)ed optical counterparts. As part of an ongoing program of optical identi(cid:12)cation of EUV sources (Maoz, Ofek, & Shemi  ) it was noticed that the ROSAT source RE + and the EUVE source EUVE J +. are separated by only  on the sky, each within the positional error circle of the other. (The typical %-con(cid:12)dence error-circles for EUVE and ROSAT, while sub ject to some uncertainties, are of order  .) RE + was detected only in the  A (cid:23) ROSAT band at the : (cid:27) level, whereas EUVE J +. was detected only in the  A EUVE band at the : (cid:27) level. These sources are not included in the latest (cid:23)  compilations of EUV sources by the ROSAT and EUVE teams (Pye et al.  ; Lewis et al.  ; Bowyer et al.  ) because their signi(cid:12)cance is slightly below the cuto(cid:11) adopted for inclusion. They are, however, likely to be real (J. Pye, private communication; X. Wu, private communication). Maoz et al. ( ) obtained B and R CCD images of the (cid:12)eld of these EUV sources, and noticed a blue (B (cid:0) R = :, R = : mag) ob ject lying between the positions of the EUV sources, and within their respective error-circles. A database search showed that the blue ob ject is positionally coincident with a  cm radio source (Gregory & Condon  ) and a faint ROSAT PSPC (.{. keV) X-ray source (White, Giommi, & Angelini  ). A (cid:12)nding chart of the (cid:12)eld is given in Fig. . The blue ob ject, and the positions and error circles of the various detections, are marked. A summary of the positions, count-rates, and estimated (cid:13)uxes at all wavelengths is given in Table . The positionally-consistent radio through X-ray emission led us to investigate this (cid:12)eld further. In particular, we were intrigued by the possibilities that this is an unknown type of Galactic EUV source, or a quasar detected in the EUV through a hole in the Galaxy's interstellar medium (ISM). Note that, apart from the detection of about  bright ({ mag) active galactic nuclei, all in the  A bands, no (cid:23) extragalactic ob jects have been detected in the EUV (Lewis et al.  ; Barber et al.  ; Marshall, Fruscione, & Carone  ). EUV radiation is very strongly absorbed by the Galaxy's ISM. We report here optical spectroscopy of the blue ob ject, and a  cm H I emission measurement in its direction.  Optical Spectroscopy A  {  A spectrum of the blue ob ject was obtained on  February   (cid:23) UT with the Large Cassegrain Spectrograph on the .-m telescope at McDonald Observatory. A Craf-Cassini  (cid:2) -pixel CCD detector was used with a  l/mm grating and a  slit at position angle to give a spectral resolution of less (cid:14) than  A. The total integration time was  s. The spectrum reveals two broad (cid:23) (FWHM { km s ) emission lines at  A and  A (C IV (cid:21) and (cid:0) (cid:23) (cid:23) C III] (cid:21) ), identifying the ob ject as a high-redshift quasar.  Additional CCD spectra were obtained on  February   UT with the Kast double spectrograph (Miller & Stone  ) at the Cassegrain focus of the -m Shane re(cid:13)ector at Lick Observatory. The total integration time was s in the range { A, plus a short ( s) exposure in the range { A to verify the (cid:23) (cid:23) reliability of the (cid:13)ux calibration in the dichroic crossover region ((cid:24) { A). (cid:23) The seeing was about  , and thin cirrus may have been present. Although the air mass was relatively high ((cid:24) :), the e(cid:11)ects of atmospheric dispersion were minimized by orienting the long slit (of width  ) along the parallactic angle; thus, the relative (cid:13)ux calibration should be accurate. The spectral resolution was (cid:24)  A below  A, (cid:23) (cid:23) and (cid:24) { A above it. (cid:23) Fig.  shows the Lick spectrum of the quasar. The Ly(cid:11), C IV, and C III] emission lines are prominent, and indicate a redshift of z = :. There is a strong unresolved (FWHM  A) absorption line near the peak of the Ly(cid:11) emission. The spectral slope (cid:23) (F / (cid:23) ) is (cid:11) (cid:25) (cid:0):, somewhat steeper than that of most quasars (e.g. Richstone & (cid:23) (cid:11) Schmidt  ). The absolute (cid:13)uxes of the Lick and McDonald spectra agree to (cid:24) %, but are about % lower than that implied by the photometric R-band measurement (obtained  months earlier, on July ,   UT). This is reasonable considering some seeing losses through the narrow slits, and the possibly non-photometric conditions of the spectroscopic observations. The equivalent widths in the observer's frame of the Ly (cid:11) + NV, C IV, and C III lines are, respectively,  A,  A, and  A. (cid:23) (cid:23) (cid:23) Their (cid:13)ux ratios are . : . : .  -cm H I Measurements The identi(cid:12)cation of the blue ob ject as a quasar raises the possibility that we are seeing a quasar in the EUV through a hole in the ISM. In the H I maps of Stark et al. ( ) the quasar (Galactic coordinates l = : , b = : ) lies inside I I (cid:14) I I (cid:14) several concentric contours of decreasing H I column density. The lowest contour in this \dip" is still fairly high, corresponding to about : (cid:2)  cm , but a deeper  (cid:0) tunnel may have been smoothed out by the map's resolution (>  ). The lowest total (cid:14) N columns that have been measured in the Galaxy are a few  cm (Lockman, H  (cid:0) Jahoda, & McCammon,  ), whereas to obtain transparency to light at  {   (cid:23)  (cid:0) A a column density lower than (cid:24)  (cid:2)  cm is required (e.g. Cox & Reynolds  ). The total count rate in the ROSAT PSPC detection is about that expected from an -mag quasar, but there are too few () total counts to determine the N H absorbing column by (cid:12)tting the x-ray spectrum. We therefore carried out a  cm measurement of the H I column density in the direction of the quasar. Observations of Galactic  cm H I emission toward + were made with the  m NRAO telescope at Green Bank, WV, which has a half-power beam-width (cid:0) of  in the  cm line. The H I spectra covered about (cid:6) km s around zero velocity relative to the local standard of rest (LSR). They were corrected for stray  cm H I radiation using the technique described by Lockman et al. ( ), although the actual correction amounted to less than %. The measured Galactic H I column density toward the quasar is : (cid:2)  cm for an assumed H I spin temperature of  (cid:0)  K. The estimated (cid:27) uncertainty on the value of N is <  (cid:2)  cm from all H  (cid:0) sources, including the opacity correction. The Galactic H I spectrum in a  (cid:2)  (cid:12)eld (cid:14) (cid:14) around + has two main spectral components, one near zero velocity LSR and the other near (cid:0) km s . The total N across this (cid:12)eld varies systematically from H (cid:0) about  (cid:2)  cm north of the quasar to about  (cid:2)  cm in the south because  (cid:0)  (cid:0) of a change in the zero velocity line strength. Toward + itself, the zero and (cid:0) km s components have widths of  and  km s (FWHM), respectively, (cid:0) (cid:0) and the negative velocity component contains  (cid:2)  cm . The uniformly large  (cid:0) values of N in the sky around +, the presence of two spectral components, H and the fairly large line widths all argue that it is unlikely that there are extreme (cid:13)uctuations in the total Galactic N on angular scales smaller than the  beam of H the  meter telescope in this part of the sky. Thus the present data should give an accurate value of the Galactic N toward the quasar. H  Conclusions As part of a program to search for optical counterparts to unidenti(cid:12)ed EUV sources, we have optically identi(cid:12)ed a z = : quasar coincident with a previously detected radio and X-ray source. Although the quasar lies within the error circles of an EUV source detected by both EUVE and the ROSAT /WFC,  cm H I emission measure-  ments in this direction show a high total H I column density. It is therefore unlikely that the quasar and the EUV source are associated. The EUV source is likely to be nearby, within (cid:24)  pc (Warwick et al.  ), and only by chance coincidence on the line-of-sight to the quasar. As shown in more detail by Maoz et al. ( ), all the stars within : of either of the EUV detections (i.e. well beyond the error circles shown in Fig.), down to a limit of R = , B = , have B (cid:0) R colors and magnitudes consistent with late-type dwarfs at distances of  pc or more. At such distances, these stars are unlikely candidate optical counterparts to the EUV source. This indicates that the EUV source is either very faint optically, or is an unknown type of nearby Galactic star that mimics the color and magnitude of a distant late-type dwarf. In the former case, the EUV/optical (cid:13)ux ratio, (cid:23)F ( A)=(cid:23)F ( A) > , i.e. the source is some kind of extremely (cid:23) (cid:23) (cid:23) (cid:23) hot ob ject. A third alternative is that the source emits transiently, and was \turned o(cid:11) " at the time of the optical observations. Upcoming observations with the Hubble Space Telescope and EUVE may provide further clues to its nature. Acknowledgements We thank K. Anderson, N. Craig, A. Laor, H. Netzer, J. Pye, and X. Wu for helpful discussions and suggestions. Astronomy at the Wise Ob- servatory is supported by grants from the Ministry of Science and Technology and from the Israel Academy of Science. The work of A. V. F. and A. J. B. was supported by the National Science Foundation through grant AST{ . The University of Texas observations are partially supported by the Space Telescope Science Institute grant HST GO .-A RQ-Q. STScI is operated by AURA, Inc., under NASA contract NAS- The National Radio Astronomy Observatory is operated by As- sociated Universities, Inc., under a cooperative agreement with the National Science Foundation. This work has made use of the NASA/IPAC Extragalactic Database (NED), which is operated by JPL, Caltech, under contract with NASA. References Barber, C.R., Warwick, R.S., McGale, P.A., & Pye, J.P.  , MNRAS, in press Bowyer, S. et al.  , ApJS, ,   Bowyer, S. et al.  , ApJS, submitted, CEA preprint No.  Cox., D. P., & Reynolds, R. J.  , ARA&A, ,  Gregory, P.C., & Condon, J.J.  , ApJS, ,  Lewis, J., Bowyer, S., Lampton, M., Wu, X., Jelinsky, P., Saba, V., Lieu, R., & Malina, R.F.  , BAAS, ,  Lockman, F.J., Jahoda, K., & McCammon, D.  , ApJ, ,  Malina, R.F., et al.  , AJ, ,  Maoz, D., Ofek, E.O., & Shemi, A.  , in preparation Marshall, H.L., Fruscione, A., & Carone, T.E.  , ApJ, in press Miller, J. S., & Stone, R. P. S.  , Lick Obs. Tech. Rep., No.  Pounds, K.A., et al.  , MNRAS, ,  Pye, J.P., et al.  , MNRAS, in press Richstone, D.O., & Schmidt, M.  , ApJ, ,  Stark, A.S., Gammie, C.F., Wilson, R.W., Bally, J., Linke, R.A., Heiles, C., & Hurwitz, M.  , ApJS,  ,  Warwick, R.S., Barber, C.R., Hodgkin, S.T., & Pye, J.P.  , MNRAS, ,  White, N.E., Giommi, P., & Angelini L.  , IAU Circ. No.   Table : Positions and Fluxes Band Instrumental Flux R.A. ( ) Dec. ( ) Positional  cm {  mJy :  :  :  Traceback (most recent call last): File "/usr/bin/pdf2txt.py", line 115, in <module> if __name__ == '__main__': sys.exit(main(sys.argv)) File "/usr/bin/pdf2txt.py", line 109, in main interpreter.process_page(page) File "/usr/lib/python2.6/site-packages/pdfminer/pdfinterp.py", line 832, in process_page self.render_contents(page.resources, page.contents, ctm=ctm) File "/usr/lib/python2.6/site-packages/pdfminer/pdfinterp.py", line 845, in render_contents self.execute(list_value(streams)) File "/usr/lib/python2.6/site-packages/pdfminer/pdfinterp.py", line 870, in execute func(*args) File "/usr/lib/python2.6/site-packages/pdfminer/pdfinterp.py", line 780, in do_EI if 'W' in obj and 'H' in obj: TypeError: argument of type 'PSLiteral' is not iterable   (cid:14) Count Rate Error R { . mag :  :  :   : : (cid:14) B { . mag :  :  :   : : (cid:14) EUVE  A  counts ks : :  : :    (cid:24)  (cid:23) (cid:0) y (cid:14) ROSAT  A  counts ks : :  : :    (cid:24)  (cid:23) (cid:0) y (cid:14) ROSAT .{. keV . counts ks : :  : :     (cid:0) y (cid:14) y Units of  erg s cm . (cid:0) (cid:0) (cid:0) 
astro-ph/9601110
1
9601
1996-01-22T09:08:40
Constraining Omega using weak gravitational lensing by clusters
[ "astro-ph" ]
The morphology of galaxy clusters reflects the epoch at which they formed and hence depends on the value of the mean cosmological density, Omega. Recent studies have shown that the distribution of dark matter in clusters can be mapped from analysis of the small distortions in the shapes of background galaxies induced by weak gravitational lensing in the cluster potential. We construct new statistics to quantify the morphology of clusters which are insensitive to limitations in the mass reconstruction procedure. By simulating weak gravitational lensing in artificial clusters grown in numerical simulations of the formation of clusters in three different cosmologies, we obtain distributions of a quadrupole statistic which measures global deviations from spherical symmetry in a cluster. These distributions are very sensitive to the value of Omega_0 and, as a result, lensing observations of a small number of clusters should be sufficient to place broad constraints on Omega_{0} and certainly to distinguish between the extreme values of 0.2 and 1.
astro-ph
astro-ph
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000) Printed 18 September 2018 (MN LATEX style file v1.3) Constraining Ω0 using weak gravitational lensing by clusters Gillian Wilson1, Shaun Cole2, Carlos S. Frenk3 Department of Physics, University of Durham, Science Laboratories, South Rd, Durham DH1 3LE [email protected] [email protected] [email protected] 18 September 2018 ABSTRACT The morphology of galaxy clusters reflects the epoch at which they formed and hence depends on the value of the mean cosmological density, Ω0. Recent studies have shown that the distribution of dark matter in clusters can be mapped from analysis of the small distortions in the shapes of background galaxies induced by weak gravitational lensing in the cluster potential. We construct new statistics to quantify the morphology of clusters which are insensitive to limitations in the mass reconstruction procedure. By simulating weak gravitational lensing in artificial clusters grown in numerical simula- tions of the formation of clusters in three different cosmologies, we obtain distributions of a quadrupole statistic which measures global deviations from spherical symmetry in a cluster. These distributions are very sensitive to the value of Ω0 and, as a result, lensing observations of a small number of clusters should be sufficient to place broad constraints on Ω0 and certainly to distinguish between the extreme values of 0.2 and 1. Key words: gravitational lensing 1 INTRODUCTION In a low density universe, density fluctuations cease to grow after a redshift z ∼ 1 − 1, where Ω0 is the present value of Ω0 the cosmological density parameter (see e.g. Peebles 1980, §11 & §13). Introducing a cosmological constant, λ0 (which we shall express in units of 3H 2 0 , where H0 is the present value of the Hubble constant) makes this cessation more abrupt. Hence, in low Ω0 models, both with and without a cosmological constant, clusters form at moderate redshift (z = 1 -- 4), and subsequently accrete very little material. In this period, which can span many dynamical times, the in- ternal structure of the clusters relaxes to produce smooth, nearly spherical configurations. In contrast, in Ω0 = 1 mod- els structure formation occurs continuously, rich galaxy clus- ters form in abundance only at very low redshift (z = 0.2 -- 0.3), and continue to accrete material even at the present epoch. Hence, if Ω0 = 1, many clusters are expected to show evidence of recent merger events and to have irregular mor- phologies. Over the past few years a number of optical and X-ray studies have suggested that a significant proportion of clus- ters, perhaps ∼ 40%, show evidence of substructure (e.g. Geller & Beers 1982; Dressler & Schectman 1988; West & Bothun 1990; Forman & Jones 1990; Bird 1994; West, Jones, & Forman 1995. Motivated by these interesting but contro- versial observations suggesting recent cluster growth, Evrard et al. (1994) and Mohr et al. (1995) demonstrated that, in agreement with analytic predictions (Richstone, Loeb, & Turner 1992; Lacey & Cole 1993), the internal structure of clusters is sensitive to the cosmological model. Evrard et al. (1994) constructed X-ray surface brightness maps from their hydrodynamic N-body simulations of cluster formation and compared these to X-ray maps of 65 clusters observed with the Einstein Imaging Proportional Counter. They concluded that galaxy clusters with the observed range of X-ray mor- phologies were more likely to have arisen in a high Ω0 cos- mology. In parallel with these developments, it has become clear that the structure of the dynamically dominant dark matter in clusters can be reliably mapped by analysing the distor- tions in the images of background galaxies induced by weak gravitational lensing in the cluster potential. Our aim in this paper is to assess whether the surface overdensity maps revealed by such analyses may be used to quantify cluster morphology and hence to constrain Ω0. To this end we sim- ulate the lensing of background galaxies by clusters drawn from three different cosmological models. These are the same clusters studied by Evrard et al. (1994). For each cluster we create a mock CCD frame as described in Wilson, Cole, & 2 G. Wilson, S. Cole, and C. S. Frenk Frenk (1996) (hereafter WCF) and then use the inversion technique of Kaiser & Squires (1993) to reconstruct a map of the projected cluster overdensity. The remainder of this paper is organised as follows. The main features of the Kaiser & Squires reconstruction method are summarized in Section 2.1. Section 2.2 intro- duces new statistics for measuring cluster shapes, specifi- cally tailored to be insensitive to uncertainties inherent in the reconstructed surface overdensity maps. In Section 2.3 we describe the simulated galaxy clusters to be used as lenses and whose density profiles we examine. Section 2.4 reviews how we generate the background distribution of galaxies and create mock CCD frames of the lensed galaxies. Our results are presented in Section 3 where we compare the reconstruc- tions to the original clusters and assess how the measured distribution of cluster shapes depends on the cosmological model. We conclude in Section 4 with a discussion and sum- mary of our main results. 2 METHODS 2.1 Lensing Reconstruction The lensing reconstruction technique employed in this paper was devised by Kaiser & Squires (1993) (hereafter KS). It produces a map of the estimated overdensity, σ(~θ), where σ(~θ) is the deviation of the true cluster surface density, S, from the mean surface density, ¯S, within the area being considered, measured in units of the critical surface density Scrit: σ(~θ) = S − ¯S Scrit , (2.1) where ~θ is the position vector on the plane of the lens. Scrit is defined as Scrit = c2 4πG Dos DolDls , (2.2) where D denotes angular-diameter distance and the sub- scripts refer to the observer, lens and source galaxy. This critical surface density is the minimum required to produce multiple images of a source object. As shown in WCF, the KS technique recovers the mass distribution in clusters extremely well when the lensing is weak i.e. when the bending angle varies slowly with po- sition. Our CCD simulations in that paper demonstrated that observational complications such as seeing and noise act to diminish the reconstructed surface density. However, we showed that it was possible to correct fully for this diminu- tion and proposed a method to estimate a multiplicative compensation factor, f. 2.2 Quantifying Shapes Lensing reconstructions such as that of KS reliably return relative values of the surface overdensity, σ(~θ) (equation 2.1), but do not fix its absolute value. Since Scrit depends on the geometry of the lensing configuration through the angular- diameter distance relationship, the redshift distribution of source galaxies must be known before the mean value of Scrit (equation 2.10 below) and hence the absolute surface overdensity S − ¯S can be obtained from equation 2.1. In addition, the KS technique is only sensitive to variations in surface density from the mean. This is because a uniform slab of material across the whole lens plane does not distort the images of galaxies lying behind. Thus, the mean surface density, ¯S, is also unknown unless the region analysed is suf- ficiently large to encompass the whole of the lensing material so that the surface density near the edge of the region can be taken as the zero-point. In order to minimise the effects of these uncertainties, we define dipole and quadrupole statistics which depend not on the absolute surface overdensity, but only on the values of σ(~θ) relative to each other. These statistics, therefore, are explicitly independent of ¯S, Scrit and any compensation factor f . We define the dipole, D(A), by [d2 1 + d2 2] 1 2 3 2 A = d 3 2 A D(A) = where d1 = Z H(σ − σcon)x dA, d2 = Z H(σ − σcon)y dA and A = Z H(σ − σcon) dA. (2.3) (2.4) (2.5) (2.6) Here, σcon is the surface overdensity of the contour level above which we choose to evaluate the statistic and A is the area within that contour. H is the Heaviside step function, H(σ − σcon), which is equal to 0 if σ < σcon and equal to 1 if σ ≥ σcon. The integrals are evaluated over the area within the contour. Similarly the quadrupole, Q(A), for the same area A is defined as [q2 1 + q2 2] 1 2 A2 = q A2 Q(A) = where q1 = Z H(σ − σcon)(x2 − y2) dA and q2 = Z H(σ − σcon)2xy dA. (2.7) (2.8) (2.9) The x and y coordinates are measured relative to the cluster centre. (We shall discuss exactly how we define the cluster centre in Section 3.) In practice, the integrals become sums over pixels in the reconstructed maps. Note that both these statistics are dimensionless. If one pictures a reconstructed surface density map as a contour plot, then these statistics are independent of the value of σ labelling each contour because we select the contour by the area A that it encloses rather than by its level. Thus, once a suitable area A has been chosen, say 5 arcmin2, the statistics are independent of :- (i) Any dilution of the lensing signal due to seeing. Constraining Ω0 using weak gravitational lensing by clusters 3 (ii) Nonlinearity effects which may result in an underes- timate of the surface overdensity in regions of high S. (iii) The (unknown) value of Scrit. The values of the statistics D(A) and Q(A) will, of course, depend on the noise level in the reconstructed surface den- sity maps. It is this effect on these statistics that we aim to quantify with our simulations of lensing. 2.3 The Cluster Simulations As our lenses we used a set of eight N-body gasdynamic sim- ulations of the formation of galaxy clusters, further details of which may be found in Evrard et al. (1994). The clusters evolved from the same eight sets of initial density fields but in three different cosmologies: (i) A biased Einstein-de Sitter model [ Ω0 = 1, σ8 = 0.59, where σ8 is the rms fluctuation of mass within spheres of radius 8h−1Mpc, and h is the present value of the Hubble constant in units of 100 kms−1Mpc−1]. (ii) An unbiased (σ8 = 1.0) open model with Ω0 = 0.2 and λ0 = 0. (iii) An unbiased (σ8 = 1.0) low density flat model with Ω0 = 0.2 and λ0 = 0.8. In each of four periodic boxes of size L = 15, 20, 25 and 30 h−1 Mpc, Evrard et al. created two constrained realizations of Gaussian random density fields, with the standard cold dark matter power spectrum appropriate for H0 = 50 km s−1 Mpc−1. In each case, using the technique of Bertschinger (1987), they imposed the constraint that, when smoothed with a Gaussian of width Rf = 0.2L, there be a peak at the centre of the box with height 2.5 -- 5 times the rms density fluctuation on this scale. Evrard's (1988) P3M/SPH hydrodynamic N-body code was used to evolve the particle distributions to the present epoch. Two sets of 323 particles represented the dark mat- ter and gas respectively. A baryon content of Ωb = 0.1 was assumed for all the models. Gravity, PdV work and shock heating were incorporated for the gas but the effects of ra- diative cooling were ignored. The spatial resolution was ap- proximately 0.005L, varying from 75 to 150 h−1 kpc depend- ing on the box length. For our lensing simulations we chose the output epoch from each cosmology closest to redshift z = 0.18. This value is fairly typical in observational studies (Wilson et al. 1996) and corresponds to the value adopted in our previous simulations (WCF). In the simulations of Evrard et al. (1994) the mass of each cluster was approximately proportional to the volume of the simulation box. Our aim here is to study the morphol- ogy of clusters detected by gravitational lensing with similar signal-to-noise ratio in each cosmology. Thus, for our pur- poses it is more convenient to have a set of clusters all of the same mass. To achieve this we simply rescaled the mass of each particle in the smaller clusters so that the total mass was the same as that in the largest cluster i.e. the one in the L = 30h−1 Mpc box. This required multiplying each mass by the factor (30h−1Mpc/L)3. In order to preserve the correct density, we also multiplied the simulation length by 30h−1Mpc/L. The typical mass of a rich cluster is known approx- imately from observations. For a fair intercomparison of the models, the clusters considered should all have approx- imately the same mass within a fiducial radius such as the Abell radius (1.5h−1Mpc). The simulated clusters in the open and flat Ω0 = 0.2 cosmologies were somewhat less massive and so we scaled them as in the previous paragraph to have approximately the same total mass as the average Ω0 = 1 cluster (∼ 1.36 × 1015h−1M⊙ within an Abell ra- dius). The maximum scaling required was a factor of 2.4 in mass. After this rescaling the initial conditions are no longer appropriate for a standard CDM model but instead have power spectra with slopes which differ slightly from the cor- rect ones. However, analytic arguments show that moderate changes in the slope of the power spectrum are far less im- portant in determining the formation history of a cluster than the actual value of Ω0 (Lacey & Cole 1993). Further- more, recent high-resolution simulations of the formation of dark matter halos in an Ω0 = 1 universe by Cole & Lacey (1995) demonstrate that a statistic similar to the one we use below to characterize cluster shapes is insensitive to spectral index and halo mass (cf their Figure 17). Thus, our rescaling should have a negligible effect on our results. We tripled the size of our cluster sample by choosing three perpendicular axes through the centre of each cluster and treating the projection along each axis as a different cluster. This resulted in 72 clusters, 24 from each of the three cosmologies. 2.4 CCD images To simulate typical observational datasets, we constructed artificial B-band CCD images of lensed field galaxies, build- ing them up pixel by pixel, in the manner described in WCF. We employed the same distributions of galaxy ellipticity, scalelength, redshift, magnitude, noise and seeing as in that paper: • Redshift and magnitude distributions For the redshift distribution of the model galaxies, we adopted the mB = 25 distribution predicted by the ana- lytic model of galaxy formation of Cole et al. (1994). Since the critical density, Scrit, depends on the source redshifts through equation (2.2), we sampled this redshift distribu- tion discretely and produced a set of source planes spanning a range of redshifts. The net effect on the reconstruction equations is that Scrit is replaced by the mean value, ¯Scrit, defined by crit = Z ¯S−1 1 Scrit(z) p(z) dz, (2.10) where p(z) is the probability that a galaxy lies at red- shift z. For our adopted redshift distribution, ¯Scrit ≃ 6 × 1015hM⊙/Mpc2, with the exact value depending on the cos- mology. The distribution of apparent magnitudes we gen- erated directly from the B-band source counts of Metcalfe et al. (1995). • Scalelength distribution We assumed that all the background galaxies are disks with exponential profiles. The scalelength was chosen from a 4 G. Wilson, S. Cole, and C. S. Frenk a) b) Figure 1. Examples of surface overdensity maps for four clusters from each of the three cosmologies. In each group the left-hand column shows the original surface density maps and the right-hand column the corresponding reconstruction. The greyscales have been adjusted according to the central surface overdensity in each case. From top to bottom the clusters are from the Ω0 = 1, open Ω0 = 0.2 and flat Ω0 = 0.2 models with the same initial fluctuation spectrum. The original CCD frame from which the reconstruction was made subtended an angle of ∼ 10 arcmins but these maps show only the inner 6.7 by 6.7 arcmins. At the cluster redshift, z = 0.18, 1 arcmin is equivalent to 0.12h−1Mpc for the Ω0 = 1 cosmology. The contour plotted encloses an area of 3.8 arcmin2. A Gaussian smoothing of width θsm = 0.25 arcminutes has been used in the KS reconstruction procedure. uniform distribution in the range 0.25 arcseconds to 0.65 arc- seconds, as suggested by the observations of Tyson (1994). • Ellipticity distribution Ellipticities for the background galaxies were generated by randomly sampling the empirical ellipticity distribution derived from a single frame in 0.7 -- 0.9 arcseconds seeing by Brainerd, Blandford, & Smail (1995). We created one CCD frame per cluster projection and then analysed it using FOCAS (Jarvis & Tyson 1981) as described in WCF. We assumed moderately good seeing of 1 arcsecond full-width half-maximum. 3 RESULTS Figures 1a-d show examples of the projected mass distribu- tion in the central regions of some of our clusters. In each group, from top to bottom, we illustrate clusters grown from identical initial conditions in the Ω0 = 1, open Ω0 = 0.2, and flat Ω0 = 0.2 cosmologies respectively. The left-hand panels give the original surface density, smoothed with a Gaus- sian of θsm = 0.25 arcminutes, the same smoothing as we employed in the KS reconstruction. The right-hand panels show the corresponding reconstructions. The field of view in each plot is 6.7 arcmins (equivalent to 0.8h−1 Mpc, for Ω0 = 1), and the contour on each map encloses an area of 3.8 arcmin2. The surface overdensities on the left-hand side have been normalised so that the mean value is zero over the full Constraining Ω0 using weak gravitational lensing by clusters 5 c) Figure 1. continued. d) CCD frame, reflecting the fact that the mean value from the KS reconstruction is automatically set to zero. To en- hance the substructure in the reconstructed maps, we have used different greyscales in the left- and right-hand panels. The former appear uniformly black near the centre. Limita- tions of scale, particularly in the open Ω0 = 0.2 case, con- ceal the reality that the density distribution is in fact very highly peaked. The maximum surface density can be critical or even greater. In contrast, the maximum surface overden- sity in the right-hand panels is ≃ 0.2 or 0.3 of critical. This diminution is partly due to observational effects such as see- ing and noise, and partly due to the effects of nonlinearity - weak lensing reconstruction techniques generally underesti- mate σ in regions where the surface overdensity rises above a few tenths of critical. This failure of the method results in reconstructed profiles that saturate near the cluster cen- tre producing a broad plateau which, in the open Ω0 = 0.2 model, extends over many hundreds of pixels with similar density values. Figure 1 illustrates how clusters in the Ω0 = 1 cos- mology are still forming at z = 0.18. In most cases, large clumps are still infalling and the dominant mass conden- sation is usually quite elongated. By contrast, clusters in the open Ω0 = 0.2 cosmology are more centrally concen- trated and closer to spherical symmetry. Clusters from the flat Ω0 = 0.2 (λ0 = 0.8) cosmology are similar to these but they have slightly lower central overdensity and are a little more elongated. In general, the reconstructed shapes match the originals rather well. With the exception of a few noisy pixels, the essential features of the different cluster morphologies are preserved by the reconstruction. We now use the dipole and quadrupole statistics, de- fined in Section 2.2, to quantify the distribution of cluster shapes in each cosmology. Initially we evaluated the statis- tics D(A) and Q(A) (equations 2.3 and 2.7) after identifying the cluster centre with the pixel of maximum surface over- density. We found that unless the reconstructions have very high signal-to-noise, this choice of centre results in noisy es- timates of D(A) and to a lesser extent of Q(A). For clusters with steep density profiles the reconstructed profiles have a saturated core and, as a result, the highest density peak is simply the highest noise peak anywhere in this central plateau. In light of this we tried a second choice of cluster centre which is more robust than simply selecting the dens- 6 G. Wilson, S. Cole, and C. S. Frenk a) b) Figure 2. The quadrupole, Q(A), as a function of the area over which it is calculated. Two examples are shown of clusters in each of the Ω0 = 1 (left panel) and open Ω0 = 0.2 models (right panel). est pixel. This time we took the centre of coordinates to be the position at which D(A) = 0 (i.e. the centroid of the area within the contour). The choice of centre now depends on the area, A, selected, but is much less sensitive to the pres- ence of noise. Having chosen the centre such that D(A) = 0 we then use only Q(A) as a measure of the asymmetry of the cluster. Figure 2a shows Q(A) as a function of area, A, for two clusters from the Ω0 = 1 cosmology. For A < 2 arcmin2, Q(A) tends to be small because of the dominant contribu- tion from the central relaxed core, although its value fluctu- ates due to noise. As larger areas (and hence smaller surface overdensities) are considered, some of the more distant in- falling clumps are included and Q(A) increases. Q(A) then varies slowly, gradually falling as an increasing number of randomly distributed noise peaks enter into the analysis. Fig. 2b shows Q(A) for two clusters from the open Ω0 = 0.2 cosmology. Q(A) is fairly stable over a wide range of area, never rising above a value of 0.1. This behaviour reflects the smooth and nearly spherical mass distribution in these clusters. We now examine the distribution of Q(A), at a fixed area A, for our sample of clusters in each of the 3 cosmolo- gies. Figure 3 shows cumulative distributions of Q(A) ob- tained from the 24 clusters in each model (see Section 2). The three panels are for A = 3.8 arcmin2, 6.3 arcmin2, and 12.6 arcmin2. The two solid lines are for Ω0 = 1, the two dot- ted lines for the open Ω0 = 0.2 model, and the two dashed lines for the flat Ω0 = 0.2 cosmology. For each pair, the line marked by triangles, usually with the smaller quadrupole value, corresponds to the original, smoothed clusters. The unmarked line corresponds to the reconstructed clusters. Examining the original clusters first, it is apparent that an open Ω0 = 0.2 universe forms clusters with the smallest quadrupoles, followed by the flat Ω0 = 0.2 model and fi- nally the Ω0 = 1 case. After reconstruction, the quadrupole value tends to increase, at least for the Ω0 = 1 and the flat Ω0 = 0.2 clusters. This is primarily due to the inclusion of outlying noisy pixels in the analysis. Figure 3 shows that even with the imperfections of the weak lensing reconstructions, there remains a strong and measurable difference in the expected distribution of clus- ter shapes in high and low Ω0 cosmologies. The dependence of these distributions on λ0 is very weak and so this statis- tic cannot be used to constrain the cosmological constant. This does mean, however, that Ω0 can be constrained in- dependently of the unknown value of λ0. For example, for A = 3.8 arcmin2 we see that 75% of clusters in an Ω0 = 1 universe have a Q(A) value in excess of 0.1, whereas less than 10% of clusters in the two Ω0 = 0.2 universes are so aspherical. Alternatively, the median value of Q(A) is ap- proximately 0.2 for both A = 3.8 and 6.3 arcmin2 in the Ω0 = 1 universe, but less than 0.06 for both the Ω0 = 0.2 universes. These large differences suggest that weak lensing Constraining Ω0 using weak gravitational lensing by clusters 7 4 DISCUSSION AND CONCLUSIONS We have shown how global properties of the dark matter distribution in clusters, as revealed by weak gravitational lensing analyses, can be used to measure the cosmologi- cal density parameter, Ω0. Our approach complements and extends earlier work by Evrard et al. (1994) who first ap- plied in practice the well established correlation between the morphology of clusters and the underlying cosmology to estimate the value of Ω0. Their analysis exploited the fact that the distribution of hot X-ray emitting gas in clusters is expected to reflect the structure of the underlying mass distribution. By probing this distribution directly, gravita- tional lensing bypasses the need to assume that the mor- phology of the gas faithfully mimics the morphology of the cluster mass. Possible (although unlikely) non-gravitational processes that may affect this correspondence are therefore not a concern, nor is there a worry that the test may be affected by contaminating X-ray signals from, for example, AGNs. A second advantage of the lensing approach is that it allows the matter distribution to be probed at larger cluster radii than the centrally concentrated X-ray emission, that is, at radii where the distinction between different cosmological models is particularly strong. The major disadvantage of the lensing method is its sensitivity to projection effects. The observed galaxy shear pattern is a function of the product of the projected sur- face density along the line-of-sight and the effective cross- section for lensing (which depends on distance through equa- tion 2.2). Mass clumps at large distances from the cluster contribute to the lensing signal and, in principle, to the mea- sured quadrupole. In practice, this is unlikely to be a strong effect, particularly when the analysis is restricted to rela- tively small areas around the cluster centre. Nevertheless the size of this contamination needs to be assessed using different simulations from those discussed here. X-ray anal- yses are much less sensitive to projection effects and so the two approaches are complementary and should be used in combination. In some respects, our detailed results must be regarded as preliminary. The simulations by Evrard et al. which we have analyzed have a number of limitations. Firstly, the ini- tial conditions in all the cosmological models were laid down with the power spectrum appropriate to an Ω0 = 1 CDM universe. This inconsistency between the power spectrum and the cosmology was further exacerbated in our case by the need to rescale the simulations in order to ensure a lens- ing signal with comparable signal-to-noise in all cosmolo- gies. Secondly, the baryon fraction in all the simulations was fixed to be 10% of the critical density so that in the Ω0 = 0.2 models, the gas contributes half of the total grav- itating mass. Finally, the clusters we have analyzed were chosen to correspond to high peaks in the density field, but they do not represent a proper statistical sample. Although we argue that all these approximations are unlikely to affect our results significantly, it is clearly desirable to repeat our analysis with purpose-built simulations. The main result of our work is that the distribution of a low-order statistic sensitive to global deviations from spher- ical symmetry in the mass distribution of clusters depends Figure 3. The cumulative probability distribution of the quadrupole, Q(A), at a fixed area, A, for A = 3.8 arcmin2 (top), 6.3 arcmin2 (middle), and 12.6 arcmin2 (bottom). The cluster centre has been chosen as the point about which the dipole van- ishes. The two solid lines are for the Ω0 = 1 model, the two dotted lines for the open Ω0 = 0.2 model, and the two dashed lines for the flat Ω0 = 0.2 model. In each case the curves marked by trian- gles correspond to the original clusters while the unmarked lines correspond to the reconstructed clusters. observations of a small number of clusters, approximately five or so, can distinguish between these two values of Ω0. A larger cluster sample is required to place finer constraints on Ω0. We stress that the distributions of Q(A) plotted in Fig- ure 3 are typical of those expected from KS mass reconstruc- tions using weak lensing data obtained in realistic observing conditions. However, they are not intended to be definitive descriptions of Q(A) for these cosmologies. Clearly, the exact form of the distributions will depend on the observing con- ditions, the seeing and limiting magnitude, as well as on the redshifts of the clusters. In practice, simulations that mimic specific observational datasets will need to be performed for this test of Ω0 to be reliably applied. 8 G. Wilson, S. Cole, and C. S. Frenk strongly on Ω0. The distribution of this quadrupole statistic can be robustly recovered from the shear pattern of galax- ies weakly lensed by the cluster gravitational potential. We have shown that under standard observing conditions, lens- ing data for a handful of clusters should distinguish between cosmological models with Ω0 = 1 and Ω0 = 0.2. ACKNOWLEDGEMENTS We would like to thank Gus Evrard for very kindly supply- ing us with his cluster simulations. We would also like to thank Chris Metzler for many useful discussions. SMC ac- knowledges the support of a PPARC Advanced Fellowship. REFERENCES Bertschinger E., 1987, Ap.J.L., 323, L103 Bird C. M., 1994, A.J., 107, 1637 Brainerd T. G., Blandford R. D., Smail I., 1995, preprint Cole S., Aragon-Salamanca A., Frenk C. S., Navarro J. F., Zepf S. E., 1994, M.N.R.A.S., 271, 781 Dressler A., Schectman S., 1988, A.J., 95, 284 Evrard A. E., Mohr J. J., Fabricant D. G., Geller M. J., 1994, Ap.J.L., 419, L9 Forman W. F., Jones C. J., 1990, in Oegerle W. R., Fitchett M. J., Danly L., ed, Clusters of Galaxies. Cambridge University Press, p. 257 Geller M. J., Beers T. C., 1982, P.A.S.P., 94, 421 Jarvis J. F., Tyson J. A., 1981, A.J., 86, 476 Kaiser N., Squires G., 1993, Ap.J., 404, 441 Lacey C., Cole S., 1993, M.N.R.A.S., 262, 627 Metcalfe N., Shanks T., Fong R., Roche N., 1995, preprint Mohr J. J., Evrard A. E., Fabricant D. G., Geller M. J., 1995, Ap.J., 447, 8 Peebles P. J. E., 1980, The Large-Scale Structure of the Universe. Princeton University Press Richstone D., Loeb A., Turner E. L., 1992, Ap.J., 393, 477 Tyson J. A., 1994, in Schaeffer R., ed, Proc. Les Houches Sum- mer School, Cosmology and Large Scale Structure. Elsevier Scientific Publishers West M. J., Bothun G. D., 1990, Ap.J., 350, 36 West M. J., Jones C., Forman W., 1995 Wilson G., Cole S., Frenk C. S., 1996, M.N.R.A.S., in press Wilson G., Smail I., Frenk C. S., Ellis R. S., Couch W. J., 1996, M.N.R.A.S., in preparation This paper has been produced using the Royal Astronomical Society/Blackwell Science LATEX style file.
astro-ph/0004279
1
0004
2000-04-19T15:18:59
The response of a dwarf nova disc to real mass transfer variations
[ "astro-ph" ]
We present simulations of dwarf nova outbursts taking into account realistic variations of the mass loss rate from the secondary. The mass transfer variation has been derived from 20 years of visual monitoring and from X-ray observations covering various accretion states of the discless cataclysmic variable AM Herculis. We find that the outburst behaviour of a fictitious dwarf nova with the same system parameters as AM Her is strongly influenced by these variations of the mass loss rate. Depending on the mass loss rate, the disc produces either long outbursts, a cycle of one long outburst followed by two short outbursts, or only short outbursts. The course of the transfer rate dominates the shape of the outbursts because the mass accreted during an outburst cycle roughly equals the mass transferred from the secondary over the outburst interval. Only for less than 10% of the simulated time, when the mass transfer rate is nearly constant, the disc is in a quasi-stationary state during which it periodically repeats the same cycle of outbursts. Consequently, assuming that the secondary stars in non-magnetic CV's do not differ from those in magnetic ones, our simulation indicates that probably all dwarf novae are rarely in a stationary state and are constantly adjusting to the prevailing value of the mass transfer rate from the secondary.
astro-ph
astro-ph
A&A manuscript no. (will be inserted by hand later) Your thesaurus codes are: 08(02.01.2;08.02.1;08.09.2 AM Her;08.14.2) ASTRONOMY AND ASTROPHYSICS 0 0 0 2 r p A 9 1 1 v 9 7 2 4 0 0 0 / h p - o r t s a : v i X r a The response of a dwarf nova disc to real mass transfer variations M.R. Schreiber, B.T. Gansicke and F.V. Hessman Universitats-Sternwarte, Geismarlandstr.11, D-37083 Gottingen, Germany the date of receipt and acceptance should be inserted later Abstract We present simulations of dwarf nova out- bursts taking into account realistic variations of the mass loss rate from the secondary. The mass transfer variation has been derived from 20 years of visual monitoring and from X-ray observations covering various accretion states of the discless cataclysmic variable AM Herculis. We find that the outburst behaviour of a fictitious dwarf nova with the same system parameters as AM Her is strongly influ- enced by these variations of the mass loss rate. Depending on the mass loss rate, the disc produces either long out- bursts, a cycle of one long outburst followed by two short outbursts, or only short outbursts. The course of the trans- fer rate dominates the shape of the outbursts because the mass accreted during an outburst cycle roughly equals the mass transferred from the secondary over the outburst in- terval. Only for less than 10% of the simulated time, when the mass transfer rate is nearly constant, the disc is in a quasi-stationary state during which it periodically repeats the same cycle of outbursts. Consequently, assuming that the secondary stars in non-magnetic CV's do not differ from those in magnetic ones, our simulation indicates that probably all dwarf novae are rarely in a stationary state and are constantly adjusting to the prevailing value of the mass transfer rate from the secondary. Key words: accretion, accretion discs - binaries: close - stars: individual: AM Her - novae, cataclysmic variables. 1. Introduction Cataclysmic variable systems (CV's), in which a white dwarf accretes material from a Roche lobe filling sec- ondary star (see Warner 1995 for an enyclopaedic re- view) can be broadly divided into three subclasses: in non-magnetic systems, the white dwarf accretes via an accretion disc; in the magnetic polars, the infalling mat- ter couples directly onto the strong magnetic field of the white dwarf before it can build a disc and is funnelled to accretion region(s) near the magnetic pole(s) of the white dwarf; finally, in CV's containing a weekly mag- netic white dwarf (intermediate polars), a partial disc may exist, with the mass flowing from the inner edge of the disrupted disc through magnetically funnelled accretion curtains onto the white dwarf. Dwarf novae are non-magnetic cataclysmic variable stars which show characteristic eruptions with photomet- ric amplitudes in the range of 2 -- 8 magnitudes 1. These eruptions typically last a few days to several weeks and recur quasi-periodically on timescales of weeks up to many years. They are generally thought to result from thermal instabilities associated with hydrogen ionization in the ac- cretion disc (see Cannizzo 1993a for a review and Ludwig et al. 1994 for a detailed parameter study). An important boundary condition of the disc-limit- cycle model is the mass transfer rate from the secondary to the accretion disc. In most studies of dwarf nova out- bursts this mass transfer rate is kept constant (Cannizzo 1993b; Ludwig & Meyer 1997; Hameury et al. 1998, 1999). Duschl and Livio (1989) were the first to examine com- bined mass transfer and disc outbursts, though within the framework of the mass transfer instability model where individual and short-lived mass transfer events are capa- ble of producing single outbursts. Smak (1991, 1999) dis- cussed mass transfer variations in the context of the su- peroutburst phenomenon in dwarf novae of the SU UMa type and dwarf nova outbursts with enhanced mass trans- fer during outburst. King & Cannizzo (1998) and Leach et al. (1999) tested how the accretion disc in a dwarf nova system behaves if the mass transfer from the secondary varies abruptly be- tween different levels. They found that these mass transfer variations produce only subtle effects on normal dwarf no- vae, including variations in the outburst shape, and that the dwarf novae keep on having outbursts even if the trans- fer rate is reduced close to zero. In spite of these efforts we are left with the question of what the real variations of the mass tranfer rates from the secondaries in dwarf nova systems are and how they can influence the outburst behaviours of the accretion discs. Fortunately, nature provides an answer to the first ques- 1 Also intermediate polars are known to show outbursts in their partial discs; e.g. EX Hya, GK Per 2 Schreiber et al.: The response of a dwarf nova disc to real mass transfer variations tion in the form of discless cataclysmic variables, the po- lars or AM Her systems. In these systems, the mass trans- fer rate can be estimated directly from the observations because there is no accretion disc acting as a mass buffer. Thus, we can use the long-term light curve of an AM Her system as a measure for the mass transfer variations in a fictitious but realistic dwarf nova system with the same system parameters. In the present paper, we present the results of such a numerical experiment. In the following section, we briefly review the equa- tions for the viscous and thermal evolution of the accretion disc, discuss the numerical methods used in our code, and compare the results of two standard models with other fine mesh computations. Thereafter we derive the mass Mtr(t). Fi- transfer rate in AM Her as a function of time, nally, we apply this mass transfer rate to a fictitious dwarf nova and discuss the effects that can be observed in the outburst behaviour. We solve Eqs. (1) and (2) using a combined Finite- Element / Finite-Difference algorithm (FE for the spatial part and FD for the time-evolution). Apart from our own work (Schreiber & Hessman 1998) the method of Finite El- ements has not been used in this context. As this method proved to be extremely robust, it warrants a somewhat more detailed description. u(x) = Pn The idea of FE is to divide the region of interest (the disc radii between Rin and Rout) into n − 1 ele- ments and to expand the function u(x) which is supposed to solve the differential equation with suitable functions i=1 aiϕi(x) for every element. In order to get a continuous solution over all elements, the functions (ϕ) of every element have to be transformed to the so-called lo- cal basis functions Ni and the coefficients to the so-called nodes (ci) before collected together (Gruber & Rappaz 1985, Schwarz 1991). To solve the differential equation, the function 2. Finite Element Methods in the context of disc evolution The classical equation describing the viscous evolution of the surface density Σ in a geometrically thin, axisymmet- ric accretion disc is obtained by combining the vertically averaged Navier-Stokes and mass-conservation equations: Mtr(t, R) 1 1 2 ∂ = (R (1) 3 R 2πR ∂ ∂R ∂R(cid:16)R 2 νΣ)(cid:17) + ∂Σ ∂t where ν = (2/3)αRgT /µΩ is the kinematic viscosity, α denotes the viscosity parameter, Rg the gas constant and µ the mean molecular weight. The first term on the right hand side is the classical disc diffusion term and the sec- ond term describes the external mass-deposition. For the purposes of this study, we follow Cannizzo (1993b) and simply add the mass lost from the secondary Mtr(t) in a Gaussian distribution at a fixed radius near the outer edge. Similarly, one can derive an energy equation (in the central temperature T : ∂T ∂t = 2(H − C + J) cpΣ − RgT µcp 1 R ∂(RvR) ∂R − vR ∂T ∂R , (2) where cp denotes the specific heat, vR = −3ν R ∂ log(νΣr ∂ log R 1 2 ) is the local radial flow velocity, H = 9 viscous heating, C = σT 4 eff is the radiative cooling and 8 νΩ2Σ represents J = 3 2 cpν Σ R ∂ ∂R (R ∂T ∂R ) the radial energy flux carried by viscous processes (see Smak 1984; Mineshige & Osaki 1983; Mineshige 1986; Ichikawa & Osaki 1992 and Cannizzo 1993b, for discus- sions). u(x) = n Xk=1 ciNi(x) (3) has to meet the requirement formulated by Garlerkin: the integral of the residuum (which one gets by inserting Eq. (3) into the differential equation) weighted with the functions Nj(x)(j = 1, ..., n) has to vanish. This require- ment, the interchange of integration and summation and partial integration lead to matrix-equations of the form B c + Ac = D, (4) with A = (aij ), B = bij and D = di (i = 1, .., n; j = 1, .., n for n nodes). To solve the differential Eqs. (1) and (2), we have to fill A, B, D in the sense mentioned above and calculate c from 2 and Eq. (4). After transforming to the variables X = 2R S = XΣ we derive for the surface density from Eq. (1): 1 n X 2 aij = Xk=1 νk(cid:16)Z 12 +Z 12 X 2 Ni bij = Z NiNjdX ∂Ni ∂X Nk ∂Nj ∂X dX ∂Nk ∂X ∂Nj ∂X dX(cid:17) di = n Xk=1 Z 2 Mtr,kNk πX 2 NjdX. Similarly it is easy to obtain for the central temperature Eq. (2): n k aij = Xk=1(cid:16)Z p(1) +Z p(2) k Nk bij = Z NiNjdR ∂Ni ∂R Nk ∂Nj ∂R dR ∂Ni ∂R NjdR +Z p(3) k NiNkNjdR(cid:17) Schreiber et al.: The response of a dwarf nova disc to real mass transfer variations 3 Figure 1. Comparison with Cannizzo (1993b, his Fig.6d). From top to bottom we used 40, 60, 80, 100 and 200 nodes. Convergence is obtained for 100 nodes. Figure 2. Comparison with Ludwig & Meyer (1997). From top to bottom we again used 40, 60, 80 ,100 and 200 nodes. Our code acurately reproduces the sequence of one long outburst followed by two short outbursts. dj = n Xk=1 Z p(4) k NkNjdR. The coefficients p(i) k are given by p(1) k = 3cpνΣ 3cpνΣ p(2) k = R Rg − vR ∂RvR ∂R p(3) k = p(4) k = µcpR 9/4νΣΩ2 − σTeff . cpΣ The index k refers to the value of the quantities at node number k, which is equivalent to the radius Rk. Notice that we only used linear basis functions in this paper. As mentioned above, Eq. (4) has to be solved to get c(t + ∆t) from c(t). Using simple finite differences leads to c(t + ∆t) = c(t) 1 2 ∆t(B−1Ac(t) +B−1Ac(t + ∆t)(+B−1D)). (5) In order to test our code, we carried out two sets of calculations using the binary parameters and cooling func- tions from Cannizzo (1993b) M1 = 1M⊙, αh = 0.1, αc = 0.02, Rin = 5.0 × 108cm, Rout = 4.0 × 1010cm, Mtr = 1.5 × 109M⊙/yr and Ludwig & Meyer (1997) M1 = 0.63M⊙, αh = 0.2, αc = 0.04, Rin = 8.4 × 108cm, Rout = 1.7 × 1010cm, Mtr = 5 × 1015g/s. The resulting light curves are shown in Figs. 1 and 2. Our code reproduces the sequence of only relatively long out- bursts found by Cannizzo (1993b) for the parameter of SS Cygni as well as the sequence of one long outburst fol- lowed by two short outbursts found by Ludwig & Meyer (1997) to describe VW Hydri. The short outbursts arise when there is not enough mass stored in the disc and there- fore the heating wave gets reflected before it has reached the outer edge of the disc. We find that at least 100 nodes are necessary for long- term convergence. The outburst and quiescence duration decreases with an increasing number of nodes because - with finer zoning - the length of time spent on the viscous plateau becomes shorter (Cannizzo 1993b). This effect is smaller in Fig. 2 because the disc is smaller in this system. We conclude that our FE-code produces results which are in excellent agreement with those of other fine-mesh computations. 3. The mass loss rate of the secondary star in AM Herculis The strong magnetic field of the white dwarf primary in polars prevents the formation of an accretion disc. With- out an accretion disc acting as a buffer for the transferred mass, the mass loss rate from the secondary equals the mass accretion rate on the white dwarf at every moment, Macc = Mtr (the free-fall time is <∼ 1 h). As an obser- vational consequence, any variation in rate at which the secondary star loses mass through the L1 point will re- sult in a quasi-immediate change of the observed accre- tion luminosity. The brightest polar, AM Her, has been intensely monitored at optical wavelengths by observers of the AAVSO for more than 20 years and shows an ir- regular long-term variability, switching back and forth be- tween high- and low states of accretion on timescales of days to months. The problem of deriving the mass loss history of the secondary star is then equivalent to that of determining the accretion luminosity Lacc as a function of time. As the 4 Schreiber et al.: The response of a dwarf nova disc to real mass transfer variations Figure 3. The mass transfer rate in AM Her as a function of time. bulk of the accretion luminosity is emitted in the X-ray regime, a bolometric correction relating the densly moni- tored optical magnitude to the total luminosity has to be derived. This approach has been followed in detail by Hess- man et al. (A&A, submitted) using X-ray observations ob- tained at multiple epochs. We summarize only briefly the results here. The accretion luminosity is computed from the observed accretion-induced flux Facc as Lacc(t) = 4πd2Facc(t) (6) with d = 90pc (Gansicke et al. 1995). We further use Facc ≈ FSX + 3 × FHX with FSX and FHX the observed soft and hard X-ray fluxes respectively. The factor three ac- counts for the additional cyclotron radiation emitted from the accretion column and for the thermal reprocession of bremsstrahlung and cyclotron radiation intercepted by the white dwarf and emitted in the ultraviolet (Gansicke et al. 1995). The mass loss ( = transfer) rate is then Mtr(t) = Lacc(t)Rwd GMwd (7) where G is the gravitational constant and Rwd and Mwd are the white dwarf radius and mass, respectively. As the actual properties of the white dwarf in AM Her are still the subject of controversial discussions (Gansicke et al. 1998, Cropper et al. 1999), we use the parameters of an average white dwarf, Mwd = 0.6 M⊙ and Rwd = 8.4 × Mtr(t) is shown in Fig. 3. The average value of the 108 cm. mass transfer rate in AM Her is Mav = 7.88 × 1015g s−1 = 1.24 × 10−10M⊙yr−1. The derived mass transfer rates of AM Her are in general agreement with results published in the literature. For example, Beuermann & Burwitz (1995) M⊙yr−1 found transfer rates between 0.8 and 2.0 × 10−10 and Greeley et al. (1999) estimated a mass transfer rate of 2 × 1016 g s−1 for the high state of AM Her from far ultraviolet spectra. 4. Results 4.1. The fictitious dwarf nova We devised a fictitious dwarf nova with a non-magnetic primary of mass Mwd = 0.6 M⊙ and an orbital period of P = 3.08 hr, i.e. a non-magnetic twin of AM Her. For these binary parameters, we obtain Rin ≃ Rwd = 8.4 × 108 cm for the inner disc radius and Rout = 2.2 × 1010 cm for the outer edge of the disc. We then used our FE code to follow the structure of the accretion disc in our fictitious dwarf nova for 7000 d, applying the variable mass transfer rate Mtr(t) derived above and standard viscosity parameters αh = 0.2 and αc = 0.04. We show 500 day-long samples of our calculations in Figs. 4 -- 8. In each figure, the top panel show the mass transfer rate as a function of time (solid line) and the av- erage mass transfer rate log( Mav[g s−1]) = 15.90 (dotted line). The panel below displays the disc mass Mdisc nor- malized with the averaged disc mass M disc = 1.64 × 1023g. The two lower panels display the light curves calculated with the varying mass transfer rate and the light curves calculated with the constant average mass transfer rate respectively. For the constant average mass transfer rate the disc goes through a ∼ 60 day-long cycle including one long outburst followed by two short outbursts. The long outbursts are those in which the entire disc is transformed into the hot state while the short outbursts arise when the outward moving heating wave is reflected as a cooling wave before it has reached the outer edge of the disc. Our numerical experiment clearly demonstrates that the outburst light curve of the fictitious system is strongly affected by the variations of the mass transfer rate. Even in the case of relatively small fluctuations effects on the outburst behaviour (16.0 < log( Mtr[g s−1]) < 16.4) are clearly present in the light curve. In Fig. 4, the mass trans- fer rate is always high during the 500 days but the disc switches between an accretion state with only long out- bursts and states where one or two short outbursts follow Schreiber et al.: The response of a dwarf nova disc to real mass transfer variations 5 Figure 4. From top to bottom as a function of time: the mass transfer rate (top, solid line), the averaged mass transfer rate (top, dotted line), the normalized disc mass and the light curves produced by the fictitious dwarf nova with the variable transfer rate and the averaged transfer rate adopted from AM Her. a long outburst. When the tranfer rate decreases some- what (∼ day 200), the disc does not save enough mass to create consecutive long outbursts until the mass transfer rate increases again (∼ day 350). In addition to this effect, our experiment shows that a sharp decrease in the mass transfer rate instantaneously changes the outburst behaviour of the accretion disc. In Fig. 5 the disc first behaves as in Fig. 4 but when the trans- fer rate drops sharply (day 4340), the long outbursts im- mediately vanish and the duration of the quiescent phase increases somewhat. Another remarkable point is that even during rel- atively long periods of very low transfer rates (Fig. 6, days 4700 -- 5000) the disc does not stop its outburst ac- tivity but produces only short outbursts with slowly de- creasing amplitudes and increasing quiescence intervals. This confirms the findings of King & Cannizzo (1998). Fig. 7 shows 500 days of our simulation in which the adopted mass transfer from the secondary varies strongly on a timescale of roughly 20 days. This is the most fre- quent case during our calculation but there are rare peri- ods of nearly constant mass transfer. Fig. 8 shows the light curve of the fictitious system during 500 days in which the mass transfer rate nearly exactly equals its averaged value (top). As a result the light curves computed with the real mass transfer rate and the averaged value look equal. In summary, one can say that the variations of the mass transfer rate leads the disc to switch between three states in which only long outbursts occur (log( Mtr) ≥ 16.3), one long outburst is followed by one or two short Figure 5. The same as Fig. 4 but another snapshot of the simulation. The sharp decline of the mass transfer rate is immediately reproduced by the accretion disc. Figure 6. The same as Fig. 4 but another snapshot of the simulation. Even the long period of low transfer rates does not stop the outburst activity of the disc. outbursts (16.3 > log( Mtr) > 15.7), and only short out- bursts occur (log( Mtr) ≤ 15.7). In order to understand the described behaviour of the disc we take into account the viscous timescale tv ∼ R2/ν which gives an estimate of the timescale for a disc annulus to move a radial distance R. For the quiescent state tv,c and the outburst state tv,h and with R = Rout (where the mass transferred from the secondary is added to the disc) we obtain for the viscous timescale tv,c ∼ R2 ν ∼ 2000 d, tv,h ∼ R2 ν ∼ 15 d. (8) (9) 6 Schreiber et al.: The response of a dwarf nova disc to real mass transfer variations The mass added to the disc during quiecence is stored in the disc because it moves inward on the long timescale tv,c whereas during an outburst even mass from outer regions can reach the white dwarf within a viscous timescale. This makes it possible that the mass accreted onto the white dwarf during a long outburst can be up to roughly one third of the disc mass (∼ day 4220). Therefore the disc can relax to equillibrium with the mass transfer rate in only one outburst in the case of high transfer rates (Fig. 4). The prompt response of the disc to the sharp decline in the mass transfer rate (Fig. 5) can be understood in the same way: due to the short viscous time (tv,h) the disc ac- cretes a substantial fraction of the disc mass (∼ 1/4Mdisc) during the last long outburst which immediately prevents long outbursts when the mass transfer rate becomes low (∼ day 4340). Finally, the long period of low transfer rates (Fig. 6) do not prevent outbursts because the mass accreted during the short outbursts is only a few percent of the disc mass. Therefore the disc relaxes to the mass transfer rate on a longer timescale. 4.2. The mass transfer and mass accretion rates An important point in understanding the physics of ac- creting binaries is to know how far the outburst behaviour and hence the resulting light curves depend on real varia- tions of the mass tranfer rate. To answer this question, we compare the averaged mass accretion rate onto the white dwarf with the mass transfer rate. Figure 9 shows that the time in which the disc relaxes to an equilibrium with the mass transfer rate depends on the occurence of long outbursts: when the accretion rate is averaged over 20 days (dotted line in Fig. 9) the mass transfer and the accretion rate correspond roughly only during periods where only long outbursts occur (days 40 -- 150, see also Fig. 4) but for the periods where the disc goes through a cycle of short and long out- bursts the accretion rate has to be averaged over 60 days to match the mass transfer rate (days 150 -- 500). If the mass transfer rate drops steeply and stays in a low -- state (days 4700 -- 5000 in Fig. 10, see also Fig. 6), the accretion rate needs more than 60 days to follow this behaviour, because the disc produces only short outbursts in which only a small percentage of the disc mass is involved. In order to make this plausible we give a relaxation timescale tr as the ratio of the viscous timescale in out- burst with the relative mass fraction accreted during an outburst: tr ∼ tv,h Mdisc ∆ Mdisc (10) For high mass transfer rates this timescale is around 70 days and so the correspondence of the averaged ac- cretion rate and the mass transfer rate in Fig. 9. is not surprising. In the case of low transfer rates (Fig. 10, day Figure 7. The same as Fig. 4 but a snapshot of the sim- ulation where the transfer rate strongly varies. Figure 8. The same as Fig. 4 but a snapshot of the sim- ulation where the transfer rate nearly stays constant. 4700 -- 5000) this timescale is longer (∼ 300 days) because only roughly 5 percent of the disc mass are accreted during an outburst. 4.3. Dependence on the primary mass In the numerical experiment above, we have assumed an average white dwarf mass for the primary in AM Her. The literature holds a large spectrum of white dwarf mass es- timates for AM Her, Mwd = 0.39 M⊙ (Young & Schneider 1981), Mwd = 0.69 M⊙ (Wu et al.1995), Mwd = 0.75 M⊙ (Mukai & Charles 1987), Mwd = 0.91 M⊙ (Mouchet 1993) and Mwd = 1.22 M⊙ (Cropper et al. 1998). Based on the observed ultraviolet spectrum of AM Her and on its well-established distance, Gansicke et al. (1998) estimated Schreiber et al.: The response of a dwarf nova disc to real mass transfer variations 7 7000 days of disc evolution with the new Mtr(t), αh = 0.1, αc = 0.02 and Rout = 2.8 × 1010 cm. In Fig. 11 we show 500 days of our calculation with Mwd = 1.0 M⊙. The disc produces only short outbursts and the outburst cycle of four outbursts with decreasing amplitude is hardly changed even by drastic variations of the mass transfer rate (day 4700). The different responses that our fictitious dwarf novae with 0.6M⊙ and 1.0M⊙ white dwarfs show to the vari- able mass transfer rate are easy to understand: both the increased primary mass and the decreased radius of the white dwarf reduce - as mentioned above - the derived average mass transfer rate. In addition, the outer disc radius of the fictitious sys- tem with Mwd = 1.0M⊙ increases. The disc becomes more massive, i.e. M disc = 4.59 × 1023g, because of its increased size. Due to the reduced mass transfer rate and the in- creased disc size, the heating waves are able to reach the outer edge of the disc only during the first outburst of the outburst cycle and only in cases where the mass transfer rate is high (log Mtr ≥ 16.1). Even in this rare situation the disc stays only a few days in the hot state and ac- cretes only a small fraction of the disc mass (∼ 1/8 Mdisc). For lower transfer rates (the extremly more frequent case shown in Fig. 11) the heating front gets reflected before it has reached the outer edge of the disc. Hence, only a small percentage of the disc mass is involved in any out- burst. Therefore, and due to the longer viscous timescale, the relaxation timescale given in Eq. (10) is always larger than a few years. Summing up, the adopted primary mass and the aver- age accretion rate play an important role on the influence that the variable mass transfer rate has on the outburst behaviour. 5. A note on the disc limit-cycle model A number of new features have been added to the well known limit-cycle model (Meyer & Meyer-Hofmeister 1981; Smak 1982; Cannizzo, Gosh & Wheeler 1982). Hameury et al. (1998, 1999, 2000) have shown how the light curves change if the disc size is allowed to vary and which effects irradiation has on the outburst behaviour. Additionally these authors discussed other sources of un- certainties such as the tidal torque and the evaporation of the inner parts of the disc (Meyer & Meyer-Hofmeister 1994). They concluded that all these effects must be in- cluded in order to obtain meaningful physical informa- tion on e.g. the viscosity from the comparison of pre- dicted and observed light curves. Another point of im- portance is the interaction of the accretion stream leav- ing the secondary star and the accretion disc. Schreiber & Hessman (1998) tested the influence of stream overflow on the disc evolution in dwarf novae. They found that signif- icant stream overflow can lead to reversion of the inward- Figure 9. The mass transfer rate (solid line) and the accretion rate onto the white dwarf as function of time. The accretion rate is averaged over 60 days (dashed line) and averaged over 20 days (dotted line). Figure 10. The mass transfer rate (solid line) and the accretion rate onto the white dwarf as function of time. The accretion rate is averaged over 60 days (dashed line). Rwd ≈ 1.1 × 109 cm, and, using the Hamada-Salpeter (1961) mass-radius relation for carbon cores, 0.35 M⊙ < Mwd < 0.53 M⊙. As the Hamada-Salpeter mass-radius relation is valid for cold white dwarfs, the finite temper- ature ≈ 20 000 K of the white dwarf in AM Her would al- low also somewhat higher masses, Mwd ≈ 0.65M⊙, which is very close to the average mass of field white dwarfs, 0.6M⊙, that we used. Even though we exclude a massive white dwarf based on the observational evidences, we repeated our simulation with Mwd = 1.0M⊙ in order to test the sensitivity of our results. In a first step, we recompute Mtr(t) from Eq. (7) with Mwd = 1.0 M⊙ and a corresponding Rwd = 5.4 × 108 cm. The resulting mean accretion rate is 3.0 × 1015g s−1, a fac- tor 2.6 lower than before. Then, we simulated once more 8 Schreiber et al.: The response of a dwarf nova disc to real mass transfer variations travelling cooling front and create an outward-travelling heating front. This behaviour would produce small dips in the light curve during the declining phase. The model we used here contains neither irradiation nor allows it the disc radi to vary. The stream mass is simply deposited in a small Gaussian distribution near the outer edge of the disc. Therefore we do not attempt at present a comparison with observed light curves. Nev- ertheless, all calculated light curves with constant mass transfer from the secondary relax in a quasi-stationary outburst cycle (of one or more outbursts) which repeat periodically. Our model shows in a qualitative way that real mass transfer variations may have a dominant influ- ence on the outburst behaviour at least in systems with relatively small discs and strong mass transfer. 6. Discussion and conclusions There is no reason to assume that the mass transfer varia- tions of the secondary observed in AM Her are not present in non-magnetic systems. Our numerical experiment in- cluding realistic variations of the mass transfer rate in a dwarf nova system is, therefore, a significant step towards a better understanding of dwarf nova light curves, and, thereby, of the underlying disc limit cycle. The light curve produced by our fictitious system switches between three states depending on the actual mass transfer rate. High transfer rates lead to only long outbursts where the entire disc is transformed into the hot state. If the transfer rate is near the average value the disc goes through a cycle of three outbursts, one long outburst followed by two short ones. Even long periods of low transfer rates do not force the disc to stop its outburst activity: long outbursts are suppressed and the duration of quiescence increases but the disc always produces short outbursts. From this follows that the low-states of VY Sculptoris stars (a subgroup of novalike variables) could not be caused by low transfer rates alone (see also Leach et al. 1999). We find that in our fictitious system the mass accreted during an outburst cycle is dominated by the course of the mass transfer rate if the mass transfer rate varies significantly. The disc always relaxes to equilibrium with the mass input from the secondary. Thus, our experiment strongly supports King & Cannizzo's (1998) claim that dwarf nova accretion discs are probably never in a sta- tionary state but are constantly adjusting to the prevailing Mtr. Only during periods where the mass transfer value of rate is nearly exactly constant the disc periodically repeat the quasy stationary outburst cycle. Such periods are rare but occur in AM Her. The strong influence of the mass transfer rate on the outburst behaviour of the fictitious system clearly indi- cates that probably most (if not all) the deviations from periodic outburst cycles seen in the light curves of dwarf novae are caused by variations of the mass transfer rate. Figure 11. The same as in Fig. 6 but assuming a more massive primary Mwd = 1.0 M⊙ Acknowledgements. We thank Daisaku Nagami and the referee for their helpful comments and suggestions. MRS would like to thank the Deutsche Forschungsgemeinschaft for financial sup- port (Ma 1545 2-1). BTG thanks for support from the DLR under grant 50 OR 99 03 6. References Beuermann K., Burwitz, 1995, ASP Conf. Ser. 85, Cape Work- shop on Magnetic Cataclysmic Variables, ed. D.A.H. Buck- ley and B. Warner (San Francisco:ASP), p.99 Cannizzo J.K., Gosh P., Wheeler J.C. 1982, ApJ 260, L83 Cannizzo J.K., 1993a, in Accretion Discs in Compact Stellar Systems, ed. J.C. Wheeler (Singapore: World Scientific), 6 Cannizzo J.K., 1993b, ApJ 419, 318 Cropper M., Ramsay G., Wu K., 1998, MNRAS, 293, 222 Cropper M., Wu K., Ramsay G., Kocabiyik A., 1999, MNRAS 306, 684 Duschl W.J. & Livio M., 1989, A&A 209, 183 Gansicke B.T., Beuermann K., de Martino D., 1995, A&A 303, 127 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., Raymond J.C., 1999, ApJ 513, 491 Gruber R. Rappaz J., 1985, Finite Element Methods in Linear Ideal Magnetohydrodynamics (Springer Verlag Heidelberg) Hamada T., Salpeter E.E., 1961, ApJ 134, 683 Hameury J.-M., Menou K., Dubus G., Lasota J.-P., Hur´e J.- M., 1998, MNRAS 298, 1048 Hameury J.-M., Lasota J.-P., Dubus G., 1999, MNRAS 303, 39 Hameury J.-M., Menou K., Lasota J.-P., Warner B., 2000, A&A 353, 244 Hessman F.V., Gansicke B.T., Mattei J., A&A submitted Ichikawa S., Osaki, Y., 1992, PASJ 44, 14 King A.R., Cannizzo J.K., 1998, ApJ 499, 348 Ludwig K., Meyer F., 1997, A&A 329, 559 Schreiber et al.: The response of a dwarf nova disc to real mass transfer variations 9 Ludwig K., Meyer-Hofmeister E., Ritter H., 1994, A&A 290, 473 Leach R., Hessman F.V., King A.R., Stehle R., Mattei J., 1999, MNRAS 305, 225 Meyer F., Meyer-Hofmeister E., 1981, A&A 104, L10 Meyer F., Meyer-Hofmeister E., 1994, A&A 288, 175 Mineshige S., Osaki Y., 1983, PASJ 44, 15 Mineshige S., 1986, PASJ 38, 831 Mouchet M., 1993, in White dwarf: Advances in Observation and Theory, ed Bartsow M., p. 411 (Dodrecht: Kluwer) Mukai K., Charles P.A., 1987, MNRAS 226, 209 Schreiber M.R., Hessman F.V., 1998, MNRAS 301, 626 Schwarz H.R., 1991, Finite Element Methods (Teubner) Smak J., 1982, Acta Astron. 32, 199 Smak J., 1984, Acta Astron. 34, 161 Smak J., 1991, Acta Astron. 41, 269 Smak J., 1999, Acta Astron. 49, 383S Warner B., 1995, Cataclysmic Variables (Cambridge Univ. Press, Cambridge) Wu K., Changmugam G., Shaviv G., 1995, ApJ 455, 260 Young P., Schneider D.P., 1981, ApJ 245, 1043
astro-ph/0309435
2
0309
2003-11-10T09:55:28
The entropy distribution in clusters: evidence of feedback?
[ "astro-ph" ]
The entropy of the intracluster medium at large radii has been shown recently to deviate from the self-similar scaling with temperature. Using N-body/hydrodynamic simulations of the LCDM cosmology, we demonstrate that this deviation is evidence that feedback processes are important in generating excess entropy in clusters. While radiative cooling increases the entropy of intracluster gas, resulting in a good match to the data in the centres of clusters, it produces an entropy-temperature relation closer to the self-similar scaling at larger radii. A model that includes feedback from galaxies, however, not only stabilises the cooling rate in the simulation, but is capable of reproducing the observed scaling behaviour both in cluster cores and at large radii. Feedback modifies the entropy distribution in clusters due to its increasing ability at expelling gas from haloes with decreasing mass. The strength of the feedback required, as suggested from our simulations, is consistent with supernova energetics, providing a large fraction of the energy reaches low-density regions and is originally contained within a small mass of gas.
astro-ph
astro-ph
Preprint astro-ph/0309435 The entropy distribution in clusters: evidence of feedback? Scott T. Kay⋆ Department of Physics and Astronomy, University of Sussex, Falmer, Brighton BN1 9QJ 12 October 2018 ABSTRACT The entropy of the intracluster medium at large radii has been shown recently to deviate from the self-similar scaling with temperature. Using N -body/hydrodynamic simulations of the ΛCDM cosmology, we demonstrate that this deviation is evidence that feedback processes are important in generating excess entropy in clusters. While radiative cooling increases the entropy of intracluster gas, resulting in a good match to the data in the centres of clusters, it produces an entropy-temperature relation closer to the self-similar scaling at larger radii. A model that includes feedback from galaxies, however, not only stabilises the cooling rate in the simulation, but is capable of reproducing the observed scaling behaviour both in cluster cores and at large radii. Feedback modifies the entropy distribution in clusters due to its increasing ability at expelling gas from haloes with decreasing mass. The strength of the feedback required, as suggested from our simulations, is consistent with supernova energetics, providing a large fraction of the energy reaches low-density regions and is originally contained within a small mass of gas. Key words: methods: N -body simulations -- hydrodynamics -- X-rays: galaxies: clusters -- galaxies: clusters: general 1 INTRODUCTION Entropy1 has become the standard quantity for describing the distribution of the intracluster medium. In a self-similar population of clusters, where entropy originates solely due to gravitational infall via shock-heating, entropy scales pro- portional to the system's virial temperature, S ∝ Tvir. Such a model, however, does not reproduce the observed X-ray properties of groups and clusters. The most striking example of this failure is the X-ray luminosity-temperature (LX −TX) relation: the self-similar scaling, LX ∝ T α X , where α = 2 (Kaiser 1986) is flatter than the observed relation for clus- ters, α ∼ 3 (e.g. Edge & Stewart 1991; Mushotzky & Scharf 1997). This deficit in luminosity, primarily in low-mass sys- tems, is due to an excess of entropy in cluster cores (Evrard & Henry 1991; Kaiser 1991), a phenomenon first measured by Ponman, Cannon & Navarro (1999). Understanding the origin of this entropy is central to our understanding of clus- ter physics (Bower 1997; Voit et al. 2002, 2003). Directly heating the intracluster gas is the most common solution to this problem, however a variety of theoretical studies con- clude that the obvious candidates -- supernovae -- are at best only marginally capable of generating the required excess en- ⋆ E-mail: [email protected] 1 We define entropy as S = kT /nγ−1 the ratio of specific heats for a monatomic ideal gas. e keV cm2, where γ = 5/3 is c(cid:13) 0000 RAS tropy (e.g. Balogh, Babul & Patton 1999; Kravtsov & Yepes 2000; Loewenstein 2000; Wu, Fabian & Nulsen 2000; Bower et al. 2001; Borgani et al. 2002). More energetic forms of supernovae (so-called hypernovae) may be important. Al- ternative sources of energy that are clearly observed to be interacting with the intracluster medium are Active Galactic Nuclei (AGN), capable of distributed heating through their release of bubbles and jets (e.g. Quilis, Bower & Balogh 2001; Bruggen & Kaiser 2002; Omma et al. 2003). Recently, attention has also focused on the effects of radiative cooling, which selectively removes the low-entropy material to form galaxies, causing higher-entropy material to flow in to replace it (Knight & Ponman 1997; Pearce et al. 2000; Bryan 2000; Muanwong et al. 2001, 2002; Dav´e, Katz & Weinberg 2002; Wu & Xue 2002). Cooling is an appealing mechanism as it involves no free parameters (other than the metallicity of the gas, which is observable): the observed level of excess entropy is consistent with the removal of low- entropy gas with cooling times shorter than the age of the Universe, scaling as S ∝ T 2/3 (Voit & Bryan 2001). Cooling, if unchecked however, overproduces the mass in galaxies, requiring a feedback mechanism in order to regulate itself (White & Rees 1978; Cole 1991; Balogh et al. 2001), but as was pointed out by Voit & Bryan (2001), their argument still holds so long as feedback efficiently transports cold gas away from cluster cores. Perhaps the likely source of excess entropy is from a combination of cooling and heating processes (Kay, Thomas 2 S. T. Kay & Theuns 2003; Tornatore et al. 2003). For example, Kay et al. (2003) studied the combined effects of cooling and feed- back (by reheating cold galactic gas) in simulations of groups and concluded that feedback, as well as regulating cooling, could contribute to the excess core entropy in systems with virial temperatures smaller than the heating temperature. In larger systems where this temperature inequality was re- versed, the inability of reheated gas to escape from the halo reduces the average entropy of the gas, and if the heating temperature is low enough, will affect the core entropy. Attention has now shifted to observing the entropy of intracluster gas at larger radii. In a recent paper, Pon- man, Sanderson & Finoguenov (2003) measured the entropy- temperature relation at a significant fraction of the virial radius (R500) for a sample of 66 galaxies, groups and clus- ters: the Birmingham-CfA Cluster Scaling Project (see also Sanderson et al. 2003; Sanderson & Ponman 2003). In their study, they measured the same deviation from self-similarity (S ∝ T 2/3) already observed in the core. This scaling be- haviour was also verified recently by Pratt & Arnaud (2003), using XMM-Newton observations of Abell 1413 & 1983. In this paper, we will argue that these observations are evi- dence that feedback plays a significant role in forming the entropy distribution in clusters. The rest of this paper is organized as follows. In Sec- tion 2 we describe the simulations used to perform this study. Our results are presented in Section 3 and summa- rized in Section 4. 2 METHOD 2.1 Simulation details We have performed four simulations of the ΛCDM cosmol- ogy, setting Ωm = 0.3, ΩΛ = 0.7, Ωb = 0.045, h = 0.7 & σ8 = 0.9, consistent with the WMAP results. Each run started from the same initial conditions: a regular cubic mesh of 2,097,152 (1283) particles (each of gas and dark matter) within a box of comoving length 60 h−1 Mpc. This set the gas and dark matter particle masses to be mgas ∼ 1.3 × 109 and mCDM ∼ 7.3 × 109 h−1 M⊙ respectively. The runs were started at z = 49 and evolved to z = 0, using a gravitational softening length equal to 20 h−1 kpc in comov- ing co-ordinates. The code used in this study is a parallel (MPI) ver- sion of gadget (Springel, Yoshida & White 2001), an N - body/hydrodynamics code that uses a PM-tree to calcu- late gravitational forces and Smooth Particle Hydrodynam- ics (SPH, e.g. Monaghan 1992) to calculate gas forces. This version of gadget uses entropy as the state variable in the time-integration of the gas (Springel & Hernquist 2002). Differences between each of the four runs are solely due to the processes incorporated that are able to change the entropy of the gas. These details are described below. 2.1.2 Radiative model Gas in this model could also lose entropy through radia- tive processes. We implemented the isochoric approxima- tion described in Thomas & Couchman (1992), using equi- librium cooling tables from Sutherland & Dopita (1993). We fixed the metallicity of the gas to Z = 0.3Z⊙. Dense gas (n > 10−3cm−3, δ > 100) which cooled below T = 1.2×104K was converted into collisionless 'stars'. This model signif- icantly overproduces the mass fraction of cooled baryons (observed to be 5-10 per cent; Cole et al. 2001; Balogh et al. 2001), producing a global cooled fraction of ∼ 34 per cent at z = 0. 2.1.3 Feedback model To stabilize the cooling rate, we only allow some of the cooled gas to form stars and reheat the rest of the cold ma- terial. Similarly to Kay et al. (2003) we incorporate feedback effects on a probabilistic basis. (A more sophisticated heat- ing mechanism is unwarranted for simulations of this reso- lution, however this method is designed to capture the gross behaviour of gas being transported out of galaxies.) Rather than instantaneously convert all cooled particles into stars, we assign a value, fheat to each cooled gas particle, defined to be the fractional mass of cooled material that is reheated, and draw a random number, r, uniformly from the unit in- terval. If r < fheat then the particle is heated, otherwise it becomes collisionless. Rather than heating particles to a fixed temperature, we instead assigned them a fixed entropy, as it is this quantity that determines how far the gas can rise buoyantly out of the cluster. We ran two simulations with feedback, hereafter re- ferred to as Weak Feedback and Strong Feedback runs. In the former case we set fheat = 0.5 (i.e. a cooled particle has an equal chance of forming stars or being reheated) and supply an entropy Sheat = 100 keV cm2 to each reheated particle. For the Strong Feedback simulation, we set fheat = 0.1 and Sheat = 1000 keV cm2. For a cold gas particle at the same density, both runs require comparable amounts of energy per mass of collisionless material but in the Strong Feedback case, the energy is added to one fifth of the mass of gas in the Weak Feedback case. Note that an associated temperature depends on the density of the gas. At n = 10−3cm−3 (where most of the gas is heated in the simulation) both runs require ∼ 1 keV of en- ergy per particle, comparable with the energy available from supernovae. However, cold gas within galaxies (our simula- tions do not resolve the internal structure of these objects) is generally at much higher densities, where much higher tem- peratures would be required to reach Sheat (if the gas flows out adiabatically). The solution to this problem is either that a large fraction of energy is transported to low-density regions in kinetic form (e.g. Strickland & Stevens 2000) or that more energetic phenomena are at work, for example hypernovae or AGN or both. 2.1.1 Non-radiative model 2.2 Cluster selection In this run, gas was only able to increase its entropy through shocks. This model reproduces self-similar scalings rather well. Clusters in each of the simulations were selected using the method described in detail in Muanwong et al. (2002). A minimal-spanning tree is first created, of all dark matter c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Entropy distribution in clusters 3 Figure 1. Entropy at fixed overdensity versus temperature for clusters in the Non-radiative (open symbols) and Radiative (filled symbols) runs at z = 0. Squares represent entropy values at 0.1R200 and triangles at R500. Solid lines represent least-squares fits to the simulated data at R500. The data-points are results from Ponman et al. (2003), with the dashed line illustrating the self-similar slope (S ∝ T ), normalized to their 8 hottest clusters. The horizontal dotted line is the entropy floor, S ∼ 124 keV cm2, from Lloyd-Davies et al. (2000). particles whose overdensity exceeds the virial value δ ∼ 178Ω−0.55(z) (Eke, Navarro & Frenk 1998). The tree is then pruned using a linking length, l = 0.5δ−1/3 times the mean interparticle separation. Spheres are then grown around the position of maximum density in each clump of particles un- til the enclosed mean density (of all particles) exceeds ∆ times the comoving critical density. In this paper we focus on results for ∆ = 500 and ∆ = 200. Finally, clusters with fewer than 500 particles each of baryons and dark matter are discarded. 3 RESULTS 3.1 Entropy-Temperature relation We begin by showing results from our Non-radiative and Radiative simulations. Fig. 1 illustrates entropy at a fixed density contrast against soft-band X-ray temperature for clusters in these two runs. The temperatures are calculated as TX = ΣimiρiΛsoft(Ti, Z)Ti ΣimiρiΛsoft(Ti, Z) , (1) where mi, ρi and Ti are the mass, density and temperature of hot gas particle i (Ti > 105K) and Z = 0.3Z⊙. We adopt the soft-band cooling function, Λsoft, from Raymond & Smith (1977) for an energy range 0.3 -- 1.5 keV. We calculate a mass- weighted entropy for each cluster by averaging the entropy of hot gas particles within a shell of width 20 h−1 kpc (our results are insensitive to changing the size of this shell by a factor of 2), centred on the radius of interest. Squares in the figure represent entropy values at 0.1R200, a radius typically chosen to highlight excess entropy c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Figure 2. Entropy at fixed overdensity versus temperature for clusters in the Weak Feedback (top panel) and Strong Feedback (bottom panel) runs at z = 0. Squares represent entropy values at 0.1R200 and triangles at R500. Solid lines represent least-squares fits to the simulated data. The data-points are results from Pon- man et al. (2003), with the dashed line illustrating the self-similar slope (S ∝ T ), normalized to their 8 hottest clusters. The hori- zontal dotted line is the entropy floor, S ∼ 124 keV cm2, from Lloyd-Davies et al. (2000). in groups (Ponman et al. 1999). As shown previously (Muan- wong et al. 2002; Kay et al. 2003), the Non-radiative model (open symbols) agrees well with the self-similar prediction (S ∝ T ), apart from at low temperatures where there is some scatter due to insufficient resolution in the core, while the Radiative model (filled symbols) is in reasonable agree- ment with the observations (the error-bars shown here are from Ponman et al. 2003). We also show in this figure (triangles) results for a lower density contrast of 500, corresponding to ∼ 2/3R200. Again, our Non-radiative results agree well with the self- similar scaling but (allowing extrapolation) are around 40 per cent higher than the expected result when normalized to the hottest 8 clusters in the Ponman et al. sample (up- per dashed line). Note however, that the Radiative results at R500 are higher than the observational data at all temper- 4 S. T. Kay atures, but scale almost self-similarly (S ∝ T 0.9 X ). Cooling increases the entropy in clusters at all radii but in this case, disagrees with the observed slope. It is therefore apparent that cooling has been too efficient at generating excess en- tropy in the most massive clusters. Fig. 2 illustrates the same relationships for our Weak Feedback and Strong Feedback runs. In the former case, we see that this model does a reasonable job at R500, flatter than the self-similar relation, but the entropy at 0.1R200 is considerably lower than observed. In this model, the gas is not receiving enough entropy to escape from the inner (X- ray dominant) region of the halo, but the feedback is having a larger effect on the gas in lower-mass systems, even at R500. (Note that reheated gas can lose some of its entropy through cooling.) The Strong Feedback model, however, matches both re- lations very well. The increase in entropy allows the gas to escape to larger radii than previously, but the fate of this gas depends on the system size. In low-temperature systems, a significant fraction of hot gas gains enough entropy to es- cape the system entirely (Kay et al. 2003), resulting in an entropy excess that is shown here to be present even at large radii. This effect diminishes in higher-temperature systems, where the entropy level of the gas is sufficiently high that less material was able to escape. 3.2 Profiles To illustrate more clearly the effects of cooling and heating on the entropy distribution of our clusters, we plot in Fig. 3 entropy profiles for two systems with kTX ∼ 1 & 3 keV respectively. At large radii in both cases, the Non-Radiative model has an entropy profile in reasonable agreement with the relation, S ∝ R1.1, as expected from spherical accretion shocks (Tozzi & Norman 2001). Both profiles flatten within ∼10 per cent of R200. Cooling increases the entropy of the gas at all radii (by removing a significant fraction of gas over the age of the uni- verse) except for in the very centre where the cooling time is very short, causing the entropy to drop sharply. These effects are more prominent in the smaller system, reflect- ing the increasing efficiency of cooling with decreasing halo mass. The Weak Feedback profiles are similar to the Radiative profiles, except that they are lower in normalization. The feedback in this case is enough to prevent some gas from cooling but not to move it around in the cluster significantly. The Strong feedback model on the other hand shows greater differences. In the 1 keV system the entropy profile is above the Radiative profile out to ∼ 0.4R200, where they both be- come comparable. Feedback therefore has increased the core entropy over and above what is possible from cooling alone, by moving some of the hot gas (which did not cool in the Ra- diative model) to larger radii. In the hotter system, however, the Strong Feedback profile is slightly lower than the Radia- tive profile out to ∼ 0.2R200, where it decreases further, and at R500 is only slightly higher than the Non-radiative profile. In this system, a larger fraction of gas was retained by the halo due to its deeper gravitational potential well. We note that the Strong Feedback profiles are a bit flat- ter than R1.1 at R500 (supported by observations). Gen- erally, systems above 1 keV have slopes between 0.8-0.9. Figure 3. Entropy profiles for two haloes (top panel: kTX ∼ 3 keV and bottom panel: kTX ∼ 1 keV) at z = 0, from the Non- radiative simulation (solid curve), Radiative simulation (dashed curve), Weak Feedback simulation (dot-dashed curve) and the Strong Feedback simulation (triple-dot-dashed curve). The ver- tical dotted line marks the softening radius and the solid vertical line R500. The solid diagonal line illustrates the predicted slope from spherical accretion shock models, S ∝ R1.1. Whether this discrepancy is a problem with this model re- quires further investigation, but will greatly benefit from both larger simulations and a larger sample of high-quality cluster data. 4 SUMMARY In this paper we used results from N -body/hydrodynamic simulations of the ΛCDM cosmology with various degrees of cooling and feedback, in order to understand the re- cent observational claim (Ponman et al. 2003; Pratt & Ar- naud 2003) that the entropy of intracluster gas does not scale self-similarly with temperature out to at least R500 (∼ 2/3R200). In particular, we compared our results to the entropy-temperature relation found by Ponman et al. at c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 0.1R200 and at R500, where both are found to be flatter (S ∝ T ∼2/3) than the self-similar scaling (S ∝ T ). Our re- sults can be summarized as follows. clusters scale • Non-radiative self-similarly in the entropy-temperature plane to good approximation, both at 0.1R200 and at R500. All but the largest clusters are observed to have an excess of entropy relative to the self-similar model at both radii. • Radiative cooling raises the entropy in clusters at all radii. While this reproduces the entropy-temperature rela- tion well at 0.1R200, it produces a relation that is too steep at R500. In this case, the relation is almost self-similar, sug- gesting that clusters above a few keV have too much entropy compared to observations. This is due to the hot gas mass being too low at all radii. • A feedback model in which gas receives a sufficient (1000 keV cm2) amount of entropy results in excellent agree- ment with the observations at both radii. The reheated gas has enough entropy to escape the X-ray core but the fraction of material unable to leave the cluster (particularly at large radii) increases with system size, flattening the entropy- temperature relation. In detail, the entropy profiles are a bit flatter (at R500) than suggested by observations, prompting future investigation into whether this discrepancy is signifi- cant. We conclude that the entropy level required to match the observations is high, but supernovae are energetically capable, providing the energy is efficiently transported into lower density regions and contained within a small mass of gas. Alternatively, more energetic phenomena (hypernovae or Active Galactic Nuclei) may play a part. ACKNOWLEDGEMENTS We thank Peter Thomas for many helpful discussions and comments on the original manuscript, Trevor Ponman for providing useful comments and supplying data-points from the Birmingham-CfA Cluster Scaling Project and Volker Springel for making his code gadget available to us. The simulations used in this paper were carried out using the Cosmology Machine at the Institute for Computational Cos- mology, Durham as part of the Virgo Consortium pro- gramme of investigations into the formation of structure in the Universe. STK is supported by PPARC. REFERENCES Balogh M. L., Babul A., Patton D. R., 1999, MNRAS, 307, 463 Entropy distribution in clusters 5 Edge A. C., Stewart G. C., 1991, MNRAS, 252, 414 Eke V. R., Navarro J. F., Frenk C. S., 1998, ApJ, 503, 569 Evrard A. E., Henry J. P., 1991, ApJ, 383, 95 Kaiser N., 1986, MNRAS, 222, 323 Kaiser N., 1991, ApJ, 383, 104 Kay S. T., Thomas P. A., Theuns T., 2003, MNRAS, 343, 608 Knight P. A., Ponman T. J., 1997, MNRAS, 289, 955 Kravtsov A. V., Yepes G., 2000, MNRAS, 318, 227 Lloyd -- Davies E. J., Ponman T. J., Cannon D. B., 2000, MNRAS, 315, 689 Loewenstein M., 2000, ApJ, 532, 16 Monaghan J.J., 1992, ARA&A, 30, 543 Muanwong O., Thomas P. A., Kay S. T., Pearce F. R., Couchman H. M. P., 2001, MNRAS, 552, L27 Muanwong O., Thomas P. A., Kay S. T., Pearce F. R., 2002, MNRAS, 336, 527 Mushotzky R. F., Scharf C. A., 1997, ApJ, 482, L13 Navarro J. F., Frenk C. S., White S. D. M., 1995, MNRAS, 275, 720 Omma H., Binney J., Bryan G., Slyz A., 2003, MNRAS, submitted (astro-ph/0307471) Pearce F. R., Thomas P. A., Couchman H. M. P., Edge A. C., 2000, MNRAS, 317, 1029 Ponman T. J., Cannon D. B., Navarro J. F., 1999, Nature, 397, 135 Ponman T. J., Sanderson A. J. R., Finoguenov A., 2003, MNRAS, 343, 331 Pratt G. W., Arnaud M., 2003, A&A, 408, 1 Quilis V., Bower R. G., Balogh M. L., 2001, MNRAS, 328, 1091 Raymond J. C., Smith B. W., 1977, ApJS, 35, 419 Sanderson A. J. R., Ponman T. J., Finoguenov A., Lloyd- Davies E. J., Markevitch M., 2003, MNRAS, 340, 989 Sanderson A. J. R., Ponman T. J., 2003, MNRAS, accepted (astro-ph/0307457) Springel V., Yoshida N., White S. D. M., 2001, New As- tronomy, 6, 79 Springel V., Hernquist L., 2002, MNRAS, 333, 649 Strickland D. K., Stevens I. R., 2000, MNRAS, 314, 511 Sutherland R. S., Dopita M. A., 1993, ApJS, 88, 253 Tornatore L., Borgani S., Springel V., Matteucci F., Menci N., Murante G., 2003, MNRAS, 342, 1025 Tozzi P., Norman C., 2001, ApJ, 546, 63 Voit G. M., Bryan G. L., 2001, Nature, 414, 425 Voit G. M., Bryan G. L., Balogh M. L., Bower R. G., 2002, ApJ, 576, 601 Voit G. M., Balogh M. L., Bower R. G., Lacey C. G., Bryan G. L., 2003, ApJ, 593, 272 White S. D. M., Rees M. J., 1978, MNRAS, 183, 341 Wu K. K. S., Fabian A. C., Nulsen P. E. J., 2000, MNRAS, 318, 889 Balogh M. L., Pearce F. R., Bower R. G., Kay S. T., 2001, Wu X.-P., Xue Y.-J., 2002, ApJ, 569, 112 MNRAS, 326, 1228 Bower R. G., 1997, MNRAS, 288, 355 Bower R. G., Benson A. J., Lacey C. G., Baugh C. M., Cole S., Frenk C. S., 2001, MNRAS, 325, 497 Bruggen M., Kaiser C. R., 2002, Nature, 418, 301 Bryan G. L., 2000, ApJ, 544, L1 Cole S., 1991, ApJ, 367, 45 Cole S. et al., 2001, MNRAS, 326, 255 Dav´e R., Katz N., Weinberg D. H., 2002, ApJ, 579, 23 c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
0811.2467
2
0811
2009-01-28T03:35:34
Warm gas accretion onto the Galaxy
[ "astro-ph" ]
We present evidence that the accretion of warm gas onto the Galaxy today is at least as important as cold gas accretion. For more than a decade, the source of the bright H-alpha emission (up to 750 mR) along the Magellanic Stream has remained a mystery. We present a hydrodynamical model that explains the known properties of the H-alpha emission and provides new insights on the lifetime of the Stream clouds. The upstream clouds are gradually disrupted due to their interaction with the hot halo gas. The clouds that follow plough into gas ablated from the upstream clouds, leading to shock ionisation at the leading edges of the downstream clouds. Since the following clouds also experience ablation, and weaker H-alpha (100-200 mR) is quite extensive, a disruptive cascade must be operating along much of the Stream. In order to light up much of the Stream as observed, it must have a small angle of attack (~20 deg) to the halo, and this may already find support in new HI observations. Another prediction is that the Balmer ratio will be substantially enhanced due to the slow shock. We find that the clouds are evolving on timescales of 100-200 Myr, such that the Stream must be replenished by the Magellanic Clouds at a fairly constant rate (>0.1 Msun/yr). The ablated material falls onto the Galaxy as a warm drizzle; diffuse ionized gas at 10^4 K is an important constituent of galactic accretion. We consider the stability of HI clouds falling towards the Galactic disk and show that most of these must break down into smaller fragments that become partially ionized. The Galactic halo is expected to have huge numbers of smaller neutral and ionized fragments. When the ionized component is accounted for, the rate of gas accretion is ~0.4 Msun/yr, roughly twice the rate deduced from HI observations alone.
astro-ph
astro-ph
The Galaxy Disk in Cosmological Context Proceedings IAU Symposium No. 254, 2008 J. Andersen, J. Bland-Hawthorn & B. Nordstrom, eds. c(cid:13) 2008 International Astronomical Union DOI: 00.0000/X000000000000000X Warm gas accretion onto the Galaxy J. Bland-Hawthorn School of Physics, University of Sydney, Australia, NSW 2006 email: [email protected] 9 0 0 2 n a J 8 2 ] h p - o r t s a [ 2 v 7 6 4 2 . 1 1 8 0 : v i X r a Abstract. We present evidence that the accretion of warm gas onto the Galaxy today is at least as important as cold gas accretion. For more than a decade, the source of the bright Hα emission (up to 750 mR†) along the Magellanic Stream has remained a mystery. We present a hydrodynamical model that explains the known properties of the Hα emission and provides new insights on the lifetime of the Stream clouds. The upstream clouds are gradually disrupted due to their interaction with the hot halo gas. The clouds that follow plough into gas ablated from the upstream clouds, leading to shock ionisation at the leading edges of the downstream clouds. Since the following clouds also experience ablation, and weaker Hα (100−200 mR) is quite extensive, a disruptive cascade must be operating along much of the Stream. In order to light up much of the Stream as observed, it must have a small angle of attack (≈ 20◦) to the halo, and this may already find support in new HI observations. Another prediction is that the Balmer ratio (Hα/Hβ) will be substantially enhanced due to the slow shock; this will soon be tested by upcoming WHAM observations in Chile. We find that the clouds are evolving on timescales of 100−200 Myr, such that the Stream must be replenished by the Magellanic Clouds at a fairly constant rate (& 0.1 M⊙ yr−1). The ablated material falls onto the Galaxy as a warm drizzle; diffuse ionized gas at 104K is an important constituent of galactic accretion. The observed Hα emission provides a new constraint on the rate of disruption of the Stream and, consequently, the infall rate of metal-poor gas onto the Galaxy. We consider the stability of HI clouds falling towards the Galactic disk and show that most of these must break down into smaller fragments that become partially ionized. The Galactic halo is expected to have huge numbers of smaller neutral and ionized fragments. When the ionized component of the infalling gas is accounted for, the rate of gas accretion is ∼0.4 M⊙ yr−1, roughly twice the rate deduced from HI observations alone. Keywords. Galaxies: interaction, Magellanic Clouds -- Galaxy: evolution -- ISM: individual (Smith Cloud) -- shock waves -- instabilities -- hydrodynamics 1. Introduction It is now well established that the observed baryons over the electromagnetic spectrum account for only a fraction of the expected baryon content in Lambda Cold Dark Matter cosmology. This is true on scales of galaxies and, in particular, within the Galaxy where easily observable phases have been studied in great detail over many years. The expected baryon fraction (ΩB/ΩDM ≈ 0.17) of the dark halo mass (1.4 × 1012 M⊙; Smith et al. 2007) leads to an expected baryon mass of 2.4 × 1011 M⊙ but a detailed inventory reveals only a quarter of this mass (Flynn et al. 2006‡). Moreover, the build-up of stars in the Galaxy requires an accretion rate of 1−3 M⊙ yr−1 (Williams & McKee 1997; Binney et al. 2000), substantially larger than what can be accounted for from direct observation. We can extend the same argument to M31 where the total baryon mass is . 1011 M⊙ (Tamm et al. 2007). For the Galaxy, the predicted baryon mass may be a arcsec−2 at Hα. † 1 Rayleigh (R) = 106/4π photons cm−2 s−1 sr−1, equivalent to 5.7 × 10−18 erg cm−2 s−1 ‡ A decade ago, it was claimed that MACHOs may be important in the halo but these can only make up a negligible fraction by mass (Tisserand et al. 2007). 1 2 J. Bland-Hawthorn lower bound if the upward correction in the LMC-SMC orbit motion reflects a larger halo mass (Kallivayalil et al. 2006; Piatek et al. 2008; cf. Wilkinson & Evans 1999). Taken together, these statements suggest that most of the baryons on scales of galaxies have yet to be observed. So how do galaxies accrete their gas? Is the infalling gas confined by dark matter? Does the gas arrive cold, warm or hot? Does the gas rain out of the halo onto the disk or is it forced out by the strong disk-halo interaction? These issues have never been resolved, either through observation or through numerical simulation. HI observations of the nearby universe suggest that galaxy mergers and collisions are an important aspect of this process (Hibbard & van Gorkom 1996), but tidal interactions do not guarantee that the gas settles to one or other galaxy. The most spectacular interaction phenomenon is the Magellanic HI Stream that trails from the LMC-SMC system (10:1 mass ratio) in orbit about the Galaxy. Since its discovery in the 1970s, there have been repeated attempts to explain the Stream in terms of tidal and/or viscous forces (q.v. Mastropietro et al. 2005; Connors et al. 2005). Indeed, the Stream has become a benchmark against which to judge the credibility of N-body+gas codes in explaining gas processes in galaxies. A fully consistent model of the Stream continues to elude even the most sophisticated codes. Here, we demonstrate that Hα detections along the Stream (Fig. 1) are providing new insights on the present state and evolution of the HI gas. At a distance of D ≈ 55 kpc, the expected Hα signal excited by the cosmic and Galactic UV backgrounds are about 3 mR and 25 mR respectively (Bland-Hawthorn & Maloney 1999, 2002), significantly lower than the mean signal of 100−200 mR, and much lower than the few bright detections in the range 400 − 750 mR (Weiner, Vogel & Williams 2002). This signal cannot have a stellar origin since repeated attempts to detect stars along the Stream have failed. Some of the Stream clouds exhibit compression fronts and head-tail morphologies (Bruns et al. 2005) and this is suggestive of confinement by a tenuous external medium. But the cloud:halo density ratio (η = ρc/ρh) necessary for confinement can be orders of magnitude larger than that required to achieve shock-induced Hα emission (e.g. Quilis & Moore 2001). Indeed, the best estimates of the halo density at the distance of the Stream (ρh ∼ 10−4 cm−3; Bregman 2007) are far too tenuous to induce strong Hα emission at a cloud face. It is therefore surprising to discover that the brightest Hα detections lie at the leading edges of HI clouds (Weiner et al. 2002) and thus appear to indicate that shock processes are somehow involved. We summarize a model, first presented in Bland-Hawthorn et al. (2007), that goes a long way towards explaining the Hα mystery. The basic premise is that a tenuous external medium not only confines clouds, but also disrupts them with the passage of time. The growth time for Kelvin-Helmholtz (KH) instabilities is given by τKH ≈ λη0.5/vh where λ is the wavelength of the growing mode, and vh is the apparent speed of the halo medium (vh ≈ 350 km s−1; see §2). At the distance of the Stream, the expected timescale for KH instabilities is less than for Rayleigh- Taylor (RT) instabilities (see §3). For cloud sizes of order a few kiloparsecs and ξ ≈ 104, the KH timescale can be much less than an orbital time (τMS ≈ 2πD/vh ≈ 1 Gyr). Once an upstream cloud becomes disrupted, the fragments are slowed with respect to the LMC-SMC orbital speed and are subsequently ploughed into by the following clouds. In §2, the new hydrodynamical models are described and the results are presented; we discuss the implications of our model and suggest avenues for future research. In §3, we discuss the stability of HI clouds (high velocity clouds) moving through the corona toward the Galactic disk and briefly consider the Smith Cloud, arguably the HVC with the best observed kinematic and photometric parameters. 2. A new hydrodynamical model There have been many attempts to understand how gas clouds interact with an ambient medium (Murray, White & Blondin 1993; Klein, McKee & Colella 1994). In order to capture the evolution of a system involving instabilities with large density gradients correctly, grid based methods (Liska & Wendroff 1999; Agertz et al. 2007) are favoured over other schemes (e.g. Smooth Particle Hydrodynamics). We have therefore investigated the dynamics of the Warm gas accretion onto the Galaxy 3 Magellanic Stream with two independent hydrodynamics codes, Fyris (Sutherland 2008) and Ramses (Teyssier 2002), that solve the equations of gas dynamics with adaptive mesh refinement. The results shown here are from the Fyris code because it includes non-equilibrium ionization, but we get comparable gas evolution from either code†. The brightest emission is found along the leading edges of clouds MS II, III and IV with values as high as 750 mR for MS II. The Hα line emission is clearly resolved at 20 − 30 km s−1 FWHM, and shares the same radial velocity as the HI emission within the measurement errors (Weiner et al. 2002; Madsen et al. 2002). This provides an important constraint on the physical processes involved in exciting the Balmer emission. In order to explain the Hα detections along the Stream, we concentrate our efforts on the disruption of the clouds labelled MS I−IV (Bruns et al. 2005). The Stream is trailing the LMC- SMC system in a counter-clockwise, near-polar orbit as viewed from the Sun. The gas appears to extend from the LMC dislodged through tidal disruption although some contribution from drag must also be operating (Moore & Davis 1994). Recently, the Hubble Space Telescope has determined an orbital velocity of 378±18 km s−1 for the LMC. While this is higher than earlier claims, the result has been confirmed by independent researchers (Piatek et al. 2008). Besla et al. (2007) conclude that the origin of the Stream may no longer be adequately explained with existing numerical models. The Stream velocity along its orbit must be comparable to the motion of the LMC; we adopt a value of vMS ≈ 350 km s−1. 400 300 200 100 0 -20 -40 -60 -80 -100 angular distance from LMC along Stream (degrees) Figure 1. Hα measurements and upper limits along the Stream. The filled circles are from the WHAM survey by Madsen et al. (2002); the filled triangles are from the TAURUS survey by Putman et al. (2003). The dashed line model is the Hα emission measure induced by the ionizing intensity of the Galactic disk (Bland-Hawthorn & Maloney 1999; 2002); this fails to match the Stream's Hα surface brightness by at least a factor of 3. 2.1. Model parameters Here we employ a 3D Cartesian grid with dimensions 18 × 9 × 9 kpc [(x, y, z) = (432, 216, 216) cells] to model a section of the Stream where x is directed along the Stream arc and the z axis points towards the observer. The grid is initially filled with two gas components. The first is a hot thin medium representing the halo corona. Embedded in the hot halo is (initially) cold HI material with a total HI mass of 3 × 107 M⊙. are provided at † Further details and comparative simulations http://www.aao.gov.au/astro/MS. on the codes 4 J. Bland-Hawthorn -7.5 -8 -8.5 -9 4 3 2 1 8 6 4 2 0 -1 freefall drag disk halo present work Wolfire 500 400 300 200 100 0 4 2 0 40 20 0 1 -1 0 1 0 2 (a) gravitational acceleration due to the disk+halo − all plots are shown as a Figure 2. function of vertical height above the disk at the Solar Circle; (b) infall velocity for a point mass starting from the Magellanic Stream with halo drag (CD = 1 discussed in §3; dashed curve) and without (solid); (c) coronal gas pressure for our model (solid line) compared to the Wolfire model (dashed) -- the dot indicates the halo pressure used in our hydrodynamical model; (d) magnetic field strength for β = 1 (dashed), β = 0.3 (dotted), β = 0.1 (solid); (e) minimum lengthscale for RT instability (discussed in §3); (e) timescale for RT instability (discussed in §3). The cold gas has a fractal distribution and is initially confined to a cylinder with a diameter of 4 kpc and length 18 kpc (Fig. 5); the mean volume and column densities are 0.02 cm−3 and 2 × 1019 cm−2 respectively. The 3D spatial power spectrum (P (k) ∝ k−5/3) describes a Kolmogorov turbulent medium with a minimum wavenumber k corresponding to a spatial scale of 2.25 kpc, comparable to the size of observed clouds along the Stream. We consider the hot corona to be an isothermal gas in hydrostatic equilibrium with the gravitational potential, φ(R, z), where R is the Galactocentric radius and z is the vertical scale height. We adopt a total potential of the form φ = φd + φh for the disk and halo respectively; for our calculations at the Solar Circle, we ignore the Galactic bulge. The galaxy potential is defined by φd(R, z) = −cdv2 φh(R, z) = chv2 circ/(R2 + (ad +qz2 + b2 circ ln((ψ − 1)/(ψ + 1)) d)2)0.5 (2.1) (2.2) h + R2 + z2)/r2 h)0.5. The scaling constants are (ad, bd, cd) = (6.5, 0.26, 8.9) and ψ = (1 + (a2 kpc and (ah, rh) = (12, 210) kpc with ch = 0.33 (e.g. Miyamoto & Nagai 1975; Wolfire et al. 1995). The circular velocity vcirc ≈ 220 km s−1 is now well established through wide-field stellar surveys (Smith et al. 2007). We determine the vertical acceleration at the Solar Circle using g = −∂φ(Ro, z)/∂z with Ro = 8 kpc. The hydrostatic halo pressure follows from Warm gas accretion onto the Galaxy ∂φ ∂z = − 1 ρh ∂P ∂z 5 (2.3) After Ferrara & Field (1994), we adopt a solution of the form Ph(z) = Po exp((φ(Ro, z) − φ(Ro, 0))/σ2 h) where σh is the isothermal sound speed of the hot corona. To arrive at Po, we adopt a coronal halo density of ne,h = 10−4 cm−3 at the Stream distance (55 kpc) in order to explain the Magellanic Stream Hα emission (Bland-Hawthorn et al. 2007), although this is uncertain to a factor of a few. We choose Th = 2 × 106 K to ensure that OVI is not seen in the diffuse corona consistent with observation (Sembach et al. 2003); this is consistent with a rigorously isothermal halo for the Galaxy. Our solution to equation (2.3) is shown in Fig. 2(c) and it compares favorably with the pressure profile derive by others (e.g. Wolfire et al. 1995; Sternberg, McKee & Wolfire 2002). A key parameter of the models is the ratio of the cloud to halo pressure, ξ = Pc/Ph. If the cloud is to survive the impact of the hot halo, then ξ & 1. A shocked cloud is destroyed in about the time it takes for the internal cloud shock to cross the cloud, during which time the cool material mixes and ablates into the gas streaming past. Only massive clouds with dense cores can survive the powerful shocks. An approximate lifetime† for a spherical cloud of diameter dc is τc = 60(dc/2 kpc)(vh/350 kms−1)−1(η/100)0.5 Myr. (2.4) For η in the range of 100−1000, this corresponds to 60−180 Myr for individual clouds. With a view to explaining the Hα observations, we focus our simulations on the lower end of this range. For low η, the density of the hot medium is nh = 2 × 10−4 cm−3. The simulations are undertaken in the frame of the cold HI clouds, so the halo gas is given an initial transverse velocity of 350 km s−1. The observations reveal that the mean Hα emission has a slow trend along the Stream which requires the Stream to move through the halo at a small angle of attack (20o) in the plane of the sky (see Fig. 5). Independent evidence for this appears to come from a wake of low column clouds along the Stream (Westmeier & Koribalski 2008). Thus, the velocity of the hot gas as seen by the Stream is (vx, vy) = (−330, −141) km s−1. The adiabatic sound speed of the halo gas is 200 km s−1, such that the drift velocity is mildly supersonic (transsonic), with a Mach number of 1.75. A unique feature of the Fyris simulations is that they include non-equilibrium cooling through time-dependent ionisation calculations (cf. Rosen & Smith 2004). When shocks occur within the inviscid fluid, the jump shock conditions are solved across the discontinuity. This allows us to calculate the Balmer emission produced in shocks and additionally from turbulent mixing along the Stream (e.g. Slavin et al. 1993). We adopt a conservative value for the gas metallicity of [Fe/H]=-1.0 (cf. Gibson et al. 2000); a higher value accentuates the cooling and results in denser gas, and therefore stronger Hα emission along the Stream. Figure 3. The dependence of the evolving fractions of Ho and H+ on column density as the shock cascade progresses. The timesteps are 70 (red), 120 (magenta), 170 (blue), 220 (green) and 270 Myr (black). The lowest column HI becomes progressively more compressed with time but the highest column HI is shredded in the cascade process; the fraction of ionized gas increases with time. The pile-up of electrons at low column densities arises from the x-ray halo. 2.2. Results The results of the simulations are shown in Figs. 3 − 5; we provide animations of the disrupting stream at http://www.aao.gov.au/astro/MS. In our model, the fractal Stream experiences a "hot wind" moving in the opposite direction. The sides of the Stream clouds are subject to gas ablation via KH instabilities due to the reduced pressure (Bernouilli's theorem). The ablated gas is slowed dramatically by the hot wind and is transported behind the cloud. As higher order † Here we correct a typo in equation (1) of Bland-Hawthorn et al. (2007). 6 J. Bland-Hawthorn modes grow, the fundamental mode associated with the cloud size will eventually fragment it. The ablated gas now plays the role of a "cool wind" that is swept up by the pursuing clouds leading to shock ionization and ablation of the downstream clouds. The newly ablated material continues the trend along the length of the Stream. The pursuing gas cloud transfers momentum to the ablated upstream gas and accelerates it; this results in Rayleigh-Taylor (RT) instabilities, especially at the stagnation point in the front of the cloud. We rapidly approach a nonlinear regime where the KH and RT instabilities become strongly entangled, and the internal motions become highly turbulent. The simulations track the progression of the shock fronts as they propagate into the cloudlets. In Fig. 3, we show the predicted conversion of neutral to ionized hydrogen due largely to cascading shocks along the Stream. The drift of the peak to higher columns is due to the shocks eroding away the outer layers, thereby progressing into increasingly dense cloud cores. The ablated gas drives a shock into the HI material with a shock speed of vs measured in the cloud frame. At the shock interface, once ram-pressure equilibrium is reached, we find vs ≈ vhη−0.5. In order to produce significant Hα emission, vs & 35 km s−1 such that η . 100. In Fig. 4, we see the steady rise in Hα emission along the Stream, reaching 100− 200 mR after 120 Myr, and the most extreme observed values after 170 Myr. The power-law decline to bright emission measures is a direct consequence of the shock cascade. The shock-induced ionization rate is 1.5 × 1047 phot s−1 kpc−1. The predicted luminosity-weighted line widths of 20 km s−1 FWHM (Fig. 4, inset) are consistent with the Hα kinematics. In Fig. 5, the Hα emission is superimposed onto the projected HI emission: much of it lies at the leading edges of clouds, although there are occasional cloudlets where ionized gas dominates over the neutral column. Some of the brightest emission peaks appear to be due to limb brightening, while others arise from chance alignments. The simulations track the degree of turbulent mixing between the hot and cool media brought on by KH instabilities (e.g. Kahn 1980). The turbulent layer grows as the flow develops, mixing up hot and cool gas at a characteristic temperature of about 104K. In certain situations, a sizeable Hα luminosity can be generated (e.g. Canto & Raga 1991) and the expected line widths are comparable to those observed in the Stream. Indeed, the simulations reveal that the fractal clouds develop a warm ionized skin along the entire length of the Stream. But the characteristic Hα emission (denoted by the shifting peak in Fig. 4) is comparable to the fluorescence excited by the Galactic UV field (Bland-Hawthorn & Maloney 2002). We note with interest that narrow Balmer lines can arise from pre-cursor shocks (e.g. Heng & McCray 2007), but these require conditions that are unlikely to be operating along the Stream. Figure 4. The evolving distribution of projected Hα emission as the shock cascade progresses. The timesteps are explained in Fig. 3. The extreme emission measures increase with time and reach the observed mean values after 120 Myr; this trend in brightness arises because denser material is ablated as the cascade evolves. The mean and peak emission measures along the Stream are indicated, along with the approximate contributions from the cosmic and Galactic UV backgrounds. Inset: The evolving Hα line width as the shock cascade progresses; the velocity scale is with respect to the reference frame of the initial HI gas. The solid lines are flux-weighted line profiles; the dashed lines are volume-weighted profiles that reveal more extreme kinematics at the lowest densities. Figure 5. The initial fractal distribution of HI at 20 Myr, shown in contours, before the wind action has taken hold. The upper figure is in the x − z frame as seen from above; the lower figure is the projected distribution on the plane of the sky (x − y plane). Both distributions are integrated along the third axis. The logged HI contours correspond to 18.5 (dotted), 19.0, 19.5, 20.0, and 20.5 (heavy) cm−2. The greyscale shows weak levels of Hα along the Stream where black corresponds to 300 mR. The predicted HI (contours) and Hα (greyscale) distributions after 120 Myr. The angle of attack in the [x,y,z] coordinate frame is indicated. The Hα emission is largely, but not exclusively, associated with dense HI gas. Warm gas accretion onto the Galaxy 7 2.3. Discussion We have seen that the brightest Hα emission along the Stream can be understood in terms of shock ionization and heating in a transsonic (low Mach number) flow. For the first time, the Balmer emission (and associated emission lines) provides diagnostic information at any position along the Stream that is independent of the HI observations. Slow Balmer-dominated shocks of this kind (e.g. Chevalier & Raymond 1978) produce partially ionized media where a significant fraction of the Hα emission is due to collisional excitation. This can lead to Balmer decrements (Hα/Hβ ratio) in excess of 4, i.e. significantly enhanced over the pure recombination ratio of about 3, that will be fairly straightforward to verify in the brightest regions of the Stream. The shock models predict a range of low-ionization emission lines (e.g. OI, SII), some of which will be detectable even though suppressed by the low gas-phase metallicity. There are likely to be EUV absorption-line diagnostics through the shock interfaces revealing more extreme kinematics (Fig. 4, inset), but these detections (e.g. OVI) are only possible towards fortuitous background sources (Sembach et al. 2001; Bregman 2007). The predicted EUV/x-ray emissivity from the post-shock regions is much too low to be detected in emission. The characteristic timescale for large changes is roughly 100−200 Myr, and so the Stream needs to be replenished by the outer disk of the LMC at a fairly constant rate (e.g. Mastropietro et al. 2005). The timescale can be extended with larger η values (equation (2.4)), but at the expense of substantially diminished Hα surface brightness. In this respect, we consider η to be fairly well bounded by observation and theory. What happens to the gas shedded from the dense clouds? Much of the diffuse gas will become mixed with the hot halo gas suggesting a warm accretion towards the inner Galactic halo. If most of the Stream gas enters the Galaxy via this process, the derived gas accretion rate is ∼ 0.4M⊙ yr−1. The higher value compared to HI (e.g. Peek et al. 2008) is due to the gas already shredded, not seen by radio telescopes now. In our model, the HVCs observed today are unlikely to have been dislodged from the Stream by the process described here. These may have come from an earlier stage of the LMC-SMC interaction with the outer disk of the Galaxy. The "shock cascade" interpretation for the Stream clears up a nagging uncertainty about the Hα distance scale for high-velocity clouds. Bland-Hawthorn et al. (1998) first showed that distance limits to HVCs can be determined from their observed Hα strength due to ionization by the Galactic radiation field, now confirmed by clouds with reliable distance brackets from the stellar absorption line technique (Putman et al. 2003; Lockman et al. 2008; Wakker et al. 2007). HVCs have smaller kinetic energies compared to the Stream clouds, and their interactions with the halo gas are not expected to produce significant shock-induced or mixing layer Hα emission, thereby supporting the use of Hα as a crude distance indicator. Here, we have not attempted to reproduce the HI observations of the Stream in detail. This is left to a subsequent paper where we explore a larger parameter space and include a more detailed comparison with the HI and Hα power spectrum, inter alia. We introduce additional physics, in particular, the rotation of the hot halo, a range of Stream orbits through the halo gas, and so on. If we are to arrive at a satisfactory understanding of the Stream interaction with the halo, future deep Hα surveys will be essential. It is plausible that current Hα observations are still missing a substantial amount of gas, in contrast to the deepest HI observations. We can compare the particle column density inferred from HI and Hα imaging surveys. The limiting HI column density is about NH ≈ hnHiL ≈ 1018 cm−2 where hnHi is the mean atomic hydrogen density, and L is the depth through the slab. By comparison, the Hα surface brightness can be expressed as an equivalent emission measure, Em ≈ hn2 eiL ≈ hneiNe. Here ne and Ne are the local and column electron density. The limiting value of Em in Hα imaging is about 100 mR, and therefore Ne ≈ 1018/hnei cm−2. Whether the ionized and neutral gas are mixed or distinct, we can hide a lot more ionized gas below the imaging threshold for a fixed L, particularly if the gas is at low density (hnei ≪ 0.1 cm−3). A small or variable volume filling factor can complicate this picture but, in general, the ionized gas still wins out because of ionization of low density HI by the cosmic UV background (Maloney 1993). In summary, even within the constraints of the cosmic microwave background (see Maloney & Bland-Hawthorn 1999), a substantial fraction of the gas 8 J. Bland-Hawthorn can be missed if it occupies a large volume in the form of a low density plasma (e.g. Rasmussen et al. 2003). 3. Direct infall of HI onto the disk The conditions operating along the Magellanic Stream are unlikely to be representative of all HI clouds that move through the Galactic corona. A related process is the infall of individual HI clouds towards the Galactic disk. The Galactic halo is home to many HVCs of unknown origin (Wakker 2001; Lockman et al. 2008). The survival and stability of these clouds is a problem that has long been recognized (e.g. Benjamin & Danly 1997) which we now discuss. It is likely that many or all halo clouds have experienced some deceleration during their transit through the lower halo. Using equation (2.3), we determine a freefall velocity for a cloud starting at the distance of the Stream (Fig. 1(b)) that is more than twice what is inferred for clouds at the Solar Circle (Wakker 2001), although some HVCs clearly have high space motions (e.g. Lockman et al. 2008). To explain this observation, Benjamin & Danly (1997) investigated a drag equation for a cloud moving through a stationary medium, µc vc = 1 2 CDρh(z)v2 c − µcg(z) (3.1) where µc is the surface density of the cloud. Equation (3.1) only holds as long as the cloud stays together. The drag coefficient CD is a measure of the efficiency of momentum transfer to the cloud. For the high Reynolds numbers typical of astrophysical media, incompressible objects have CD ≈ 0.4 (e.g. a rough sphere) which indicates that the turbulent wake behind the plunging object efficiently transfers momentum to the braking medium. The leading face of a compressible cloud may become flattened, such that the approaching medium is brought to rest in the reference frame of the cloud; in this instance, CD & 1 may be more appropriate (we adopt CD = 1 here). A solution specific to our model is shown in Fig. 2(b) where the freefall velocity is now slowed by about 35%. In practice, the cloud's projected motion can be considerably less than its 3D space velocity (e.g. Lockman et al. 2008). In all likelihood, infalling HVCs have experienced significant deceleration through ram pressure exerted by the corona. But even before the cloud reaches terminal velocity, the cloud is expected to break up (Murray & Lin 2004). So how do clouds resist the destructive forces of RT and shock instabilities? In §3.1, we investigate the stabilizing influence of magnetic fields when a cloud passes through a magnetized medium. The halo magnetic field is poorly constrained at the present time (e.g. Sun et al. 2008). We describe the uniform magnetic field in terms of the pressure of the halo medium, or B2 8π = βPh (3.2) such that B ≈ 1 µG at the distance of the Stream (55 kpc) if the field is in full equipartition with the corona (see Fig. 1(d)). But there is evidence that the field is weaker than implied here (β ≈ 0.3; Sun et al. 2008), at least within 5 kpc of the Galactic plane For the warm, denser low- latitude gas (Reynolds layer), we adopt the new parameter fits of Gaensler et al. (2008) from a re-analysis of pulsar data. The lower β value finds support from recent magnetohydrodynamic simulations of the Reynolds layer (Hill et al. 2008). 3.1. Stability limits and growth timescales We consider the surface of a high velocity cloud as a boundary between two fluids. In practice, the Galactic ionizing radiation field imparts a multiphase structure to the cloud. At all galactic latitudes within the Stream distance, HVCs with column densities of order 1020 cm−2 or higher have partially ionized skins to a column depth of roughly 1019 cm−2 for sub-solar gas due to the Galactic ionizing field (see Bland-Hawthorn & Maloney 1999; Wolfire et al. 2003). Between the warm ionized skin and the cool inner regions is a warm neutral medium of twice the skin thickness; both outer layers have a mean particle temperature of . 104K. Warm gas accretion onto the Galaxy 9 The cloud is denser than the halo gas. Because of the gravitational field, RT instabilities can grow on the boundary. Furthermore, KH instabilities may also develop due to the relative motion of the cloud with respect to an external medium. Recent work has shown that buoyant bubbles in galaxy clusters are stabilized against RT and KH instabilities by viscosity and surface tension due to magnetic fields in the boundary (De Young 2003; Kaiser et al. 2005; Jones & De Young 2005). Here we examine whether HVC boundaries are similarly stabilized against disruption in the Galactic halo. When there is no surface tension, no viscosity and no relative motion between the two media, the growth rate of the RT instability for a perturbation with wavenumber k is ω = √gk, where g is the gravitational acceleration at the fluid boundary. The wavenumber is related to a perturbation length scale, ℓ = 2π/k. The instability requires a few e-folding timescales to fully develop; the timescale is given by tgrow = ω−1 =s ℓ 2πg . (3.3) In the presence of a magnetic field, the transverse component (Btr) provides some surface tension which can help to suppress RT instabilities below a lengthscale of ℓmin = B2 tr 2ρcg (3.4) (Chandrasekhar 1961). Here ρc is the mass density of the denser medium, i.e. the cloud, and Btr is the average value of the transverse magnetic field at the boundary. In order to illustrate when RT instabilities become important, we assume a flat rotation curve for the Galaxy (e.g. Binney & Dehnen 1997) v2 circ R g ≈ = 1.6 × 10−8" vcirc 220 km s−1"2„ R Ro«−1 cm s−2. (3.5) This is only a rough approximation to the form expected from equations (2.1) and (2.2). We stress that the actual behaviour discussed below, and shown in Figs. 2(e) and (f), solves for the gravitational potential correctly. Shortly after the discovery of HVCs, it was thought that they may be self-gravitating. But this would place them at much greater distances than the Magellanic Stream (e.g. Oort 1966) which is now known not to be the case (e.g. Putman et al. 2003). Instead, we consider two cases: (i) HVCs in pressure equilibrium with the coronal gas; (ii) HVCs with parameters fixed by direct observation. In (i), because the temperature is not strongly dependent on radius, but the number density decreases rapidly with increasing radius, we expect the increased pressure to compress the clouds at lower latitudes. We estimate the impact of RT instabilities using equations (2.3) and (3.3): for a cloud tem- perature of Tc = 104K (see §2) in pressure equilibrium with the hot halo, the electron density is given by ne,c ≈ ne,h Th Tc = 0.02„ R 55 kpc«−2„ Th 2 × 106K«„ Tc 104K«−1 (3.6) cm−3. We use equations 3.4, 3.6 and 3.2 to estimate ℓmin as a function of Galactocentric radius. The minimum length scale for instability is ℓmin R ∼ 8πβkBTc mpv2 circ (3.7) = 0.004„ Tc 104K«„ β 0.1«" vcirc 220 km s−1"−2 10 J. Bland-Hawthorn and its associated growth timescale using equation (3.3) tgrow ∼s 4βkBTc mpv2 circ Ω−1 (3.8) = 1.1 Myr„ Tc 104K« 1 2 „ β 0.1« vcirc Ro« 220 km s−1"−2„ R 1 2 " where the angular rotation rate is given by Ω = vcirc/R. Because we have assumed that B2 ∝ ne,h (equipartition) and ne,c ∝ ne,h (pressure equilib- rium), neither the minimum scale length or its growth timescale depend on the halo density or temperature. They do depend on the temperature of the clouds and the ionization state. If the clouds are hotter than 104K, then ne,c is overestimated under the assumption of pressure equilibrium. This would lead to larger minimum instability lengthscales and growth timescales. If the Galactic rotation curve drops faster than the flat profile implied by equation (2.3), we would have underestimated both the minimum instability scale length and its associated growth timescale at large radii. Under the assumption of pressure equilibrium, the falling clouds become more compressed as they approach the disk which can hasten cooling. This effect may help to stabilize against break up, particularly if a cool shell develops (cf. Sternberg & Soker 2008). In the absence of gravitational instability, the flow is stable against the KH instability if (Chandrasekhar 1961) where U is the relative velocity between the two fluids. When ρc > ρh, this requirement becomes U 2 < B2 tr(ρc + ρh) 2πρcρh (3.9) mp U <s 8βkBTh = 115„ β 0.1« 1 2 1 2 „ Th 2 × 106K« km s−1 (3.10) and we have described the magnetic field in terms of the halo pressure using equation 3.2. This requirement is also independent of the halo density as we have related the magnetic field to the halo pressure, although it is dependent on the halo temperature. This requirement is nearly satisfied for HVCs if the magnetic field is near equipartition. 3.2. The Smith Cloud Arguably, the high-latitude HI cloud that we know most about is the Smith Cloud. Lockman et al. (2008) have recently published spectacular HI data for this HVC and deduce a remarkable amount about its past and future properties. The HVC has an estimated distance of 12.4 ± 1.3 kpc, a Galactocentric radius of R ≈ 8 kpc, a vertical height below the plane of -2.9 kpc, a mass of at least 106 M⊙ in a volume of order 3 kpc3 corresponding to nc ≈ 0.014 cm−3. The cloud has a prograde orbit that is inclined 30◦ to the plane and appears to have come through the disk 70 Myr ago at R ≈ 13 kpc moving from above to below the plane. longer than for the disk. It can be shown that In order to have punched through the disk, the shock crossing time for the cloud must be dc zd >r nd nc (3.11) where zd is the vertical thickness and nd is the mean density of the HI at the crossing point. This is essentially a statement that the surface density of the cloud must be higher than the disk. If we assume the cloud punched through the Galactic hydrogen density profile determined by Kalberla & Dedes (2008), equation 3.11 indicates that the cloud was substantially thicker than the disk when it came through and somewhat more massive than what is observed now. Warm gas accretion onto the Galaxy 11 Consistent with this picture, the observed wake may result from ablation processes induced by the impact. For cloudlets smaller than 100 pc, thermal conduction due to the halo corona (McKee & Cowie 1977) and the Galactic radiation field convert the ablated gas to a clumpy plasma. The kinetic energy of the Smith Cloud observed today is ∼ 1054 erg -- this is enough to punch through the disk if sufficiently concentrated. Impulsive shock signatures at UV to x-ray wavelengths will have largely faded away, and the HI "hole" at the crossing point will have been substantially stretched by differential shear†. Figs. 2(e) and (f) show that a cloud of several kpc can survive RT instabilities at these latitudes, but it is difficult to see how the Smith Cloud, like several other large HVCs, could have come in from, say, the distance of the Magellanic Stream. Lockman et al. (2008) use essentially the same Galactic potential as described here to determine the cloud's orbit parameters. We conjecture that either the cloud has been dislodged from the outer disk by a passing dwarf, or the cloud has been brought in by a confining dwarf potential. A cloud metallicity of [Fe/H]≈-1 is appropriate in either scenario. Interestingly, the impulse from the Galactic disk can cause the gas to become dislodged from the confining dark halo or to oscillate within it. The interloper must be on a prograde orbit which rules out some infalling dwarfs (e.g. ωCen; Bekki & Freeman 2007), but conceivably implicates disrupting dwarfs like Canis Major or Sagittarius, assuming these were still losing gas in the recent past. 3.3. Discussion In §2, we presented evidence for a shock cascade along the Magellanic Stream arising from the disruption of upstream clouds due to their interaction with the Galactic halo. Bland-Hawthorn et al. (2007) make firm predictions that can be tested in future observations. A possible improve- ment is to consider the entire Magellanic System, i.e. the influence of the LMC-SMC system that lies further upstream. Mastropietro et al. (2008) present evidence for a strong interaction along the leading edge of the LMC; for their quoted model parameters, it seems plausible that this results in a stand-off bowshock ahead of the galaxy. The cross wind over the face of the LMC could be confused for a starburst-driven wind from the LMC (cf. Lehner & Howk 2007). In all likelihood, the LMC-SMC system creates a turbulent wake behind it which may impact the development of instabilities in the trailing stream. The issue of cloud survival is highly complex. In §3, we did not consider the role of viscosity in quenching RT or KH instabilities. Simulations have shown that viscosity does lead to stabi- lization (Pavlovski et al. 2008) but we have not been able to estimate a lengthscale or a growth timescale appropriate for our setting. Kaiser et al. (2005) show that when the density ratio between the two media is large, KH instabilities fail to grow and the growth rate of RT insta- bilities depends only on the properties of higher density medium, in our case the cloud medium. However their result, taken in the limit of one density much larger than the other, ρ2 ≫ ρ1, will not apply if ν1ρ1 ≫ ν2ρ2. Here the subscripts refer to the fluids on either side of the boundary and ν is the kinematic viscosity. Because diffusivity coefficients are sensitive functions of temper- ature (∝ T −2.5), they could dampen fluid instabilities. Unfortunately the expected differences in temperatures between HVCs and the halo gas (corona) suggest that ν1ρ1 ≫ ν2ρ2 and thus we cannot apply the limit used by Kaiser et al. (2005). A proper treatment is required to cover the Galactic halo setting. Other studies have argued that the KH instability leads to a turbulent mixing layer on the surface and so is less destructive than the RT instability (e.g. De Young 2003). At the present time, there are no relevant astrophysical codes that are capable of handling mixing in a satis- factory manner. On the issue of magnetic stability, more sophisticated treatments using MHD have been attempted, but the main conclusions appear to be contradictory (Konz et al. 2002; Gregori et al. 1999). We are not aware of MHD codes that are sufficiently capable of answering this question at the present time. † It is sometimes claimed, this meeting notwithstanding, that outer disk HI "holes" are evi- dence of dark matter minihalos passing through the disk, but it can be shown that the gravita- tional impulse has negligible impact on the gaseous disk. 12 J. Bland-Hawthorn Without excessive erudition, which is inappropriate for a conference proceeding, it is difficult to mount a solid case for why hydro processes could ultimately save the day for HVCs. But the fact of the matter is that fast-moving gas clouds do survive their passage through the Galactic halo. These may be mostly shortlived entities on the road to destruction, suggesting that there is a largely hidden plasma component that we have yet to fully comprehend. This will require more extensive observations at difficult parts of the observational parameter space, matched by hydro codes that can properly treat instabilities and mixing in a multi-phase gas. Acknowledgements JBH is supported by a Federation Fellowship through the Australian Research Council. I thank Alice Quillen and Ralph Sutherland for their role in the work presented here. I acknowledge helpful discussions with Bob Benjamin, Chris Flynn, Bryan Gaensler, Greg Madsen and Mary Putman. References Agertz, O., et al. 2007, MNRAS, 380, 963 Bekki, K. & Freeman, K.C. 2003, MNRAS, 346, L11 Benjamin, R. & Danly, L. 1997, ApJ, 481, 764 Besla, G. et al. 2007, ApJ, 668, 949 Binney, J. & Dehnen, W. 1997, MNRAS Binney, J., Dehnen, W., & Bertelli, G. 2000, MNRAS, 318, 658 Bland-Hawthorn et al. 1998, MNRAS, 299, 611 Bland-Hawthorn, J., & Maloney, P.R. 1999, ApJL, 510, L33 Bland-Hawthorn, J., & Maloney, P.R. 2002, Extragalactic Gas at Low Redshift, 254, 267 Bland-Hawthorn, J., Sutherland, R., Agertz, O., & Moore, B. 2007, ApJ, 670, L109 Bregman, J. N. 2007, ARA&A, 45, 221 Bruns, C., et al. 2005, AAp, 432, 45 Canto, J., & Raga, A.C. 1991, ApJ, 372, 646 Chandrasekhar, S. 1961, Hydrodynamic and Hydromagnetic Stability, Clarendon, Oxford Chevalier, R.A., & Raymond, J.C. 1978, ApJL, 225, L27 Connors, T.W., Kawata, D., & Gibson, B.K. 2006, MNRAS, 371, 108 De Young, D. S. 2003, MNRAS, 343, 719 Ferrara, A. & Field, G.B. 1994, ApJ, 423, 665 Flynn, C. et al. 2006, MNRAS, 372, 1149 Gaensler, B. et al. 2008, PASA, submitted Gibson, B.K. et al. 2000, AJ, 120, 1830 Gregori, G. et al. 1999, ApJ, 527, L113 Heng, K., & McCray, R. 2007, ApJ, 654, 923 Hibbard, J.E., & van Gorkom, J.H. 1996, AJ, 111, 655 Hill, A. et al. 2008, ApJ, 686, 363 Jones, T.W. & De Young, D. 2005, ApJ, 624, 586 Kahn, F.D. 1980, AAp, 83, 303 Kaiser, C. R., Pavlovski, G., Pope, E. D. C., & Fangohr, H. 2005, MNRAS, 359, 493 Kalberla, P. & Dedes, L. 2008, A& A, in press (0804.4831) Kallivayalil, N., van der Marel, R.P. & Alcock, C., 2006, ApJ, 652, 1213 Keres, D., Katz, N., Weinberg, D. H., & Dave, R. 2005, MNRAS, 363, 2 Klein, R.I., McKee, C.F., & Colella, P. 1994, ApJ, 420, 213 Konz, C., Bruns, C. & Birk, G.T. 2002, A&A, 391, 713 Larson, R. B. 1969, MNRAS, 145, 405 Lehner, N. & Howk, C. 2007, MNRAS, 377, 687 Liska, R., & Wendroff, B. 1999, International Journal for Numerical Methods in Fluids, 30, 461 Lockman, F.J., Benjamin, R.A., Heroux, A.J. &, Langston, G.I. 2008, ApJ, 679, L21 Madsen, G.J., Haffner, L.M. & Reynolds, R.J. 2002, ASPC, 276, 96 Warm gas accretion onto the Galaxy 13 Maloney, P.R., & Bland-Hawthorn, J. 1999, ApJL, 522, L81 Maloney, P. 1993, ApJ, 414, 41 Mastropietro, C., Moore, B., Mayer, L., Wadsley, J., & Stadel, J. 2005, MNRAS, 363, 509 McKee, C.F. & Cowie, L.L. 1977, ApJ, 215, 213 Miyamoto, M. & Nagai, R. 1975, PASJ, 27, 533 Moore, B. & Davis, M. 1994, ApJ, 270, 209 Murray, S.D., White, S.D.M., Blondin, J.M., & Lin, D.N.C. 1993, ApJ, 407, 588 Murray, S.D. & Lin, D.C. 2004, ApJ, 615, 586 Nicastro, F., Mathur, S., & Elvis, M. 2008, Science, 319, 55 Oort, J. 1966, Bull. Astron. Inst. Neth., 18, 421 Pavlovski, G., Kaiser, C., Pope, E. C. D., & Fangohr, H. 2008, MNRAS, 384, 1377 Peek, J.E.G., Putman, M.E., & Sommer-Larsen, J. 2008, ApJ, 674, 227 Piatek, S., Pryor, C. & Olszewski, E.W. 2008, AJ, 135, 1024 Putman, M.E. et al. 2003, ApJ, 597, 948 Quilis, V. & Moore, B. 2001, ApJ, 555, L95 Rasmussen, A., Kahn, S. & Paerels, F. 2003, ASSL, 281, 109 Rosen, A., & Smith, M.D. 2004, MNRAS, 347, 1097 Ruszkowski, M., Enslin, T. A., Bruggen, M., Heinz, S., Pfrommer, C. 2007, MNRAS, 378, 662 Savage, B. D. et al. 2003, ApJS, 146, 125 Sembach, K.R., Howk, J.C., Savage, B.D., Shull, J.M., & Oegerle, W.R. 2001, ApJ, 561, 573 Sembach, K. R., et al. 2003, ApJS, 146, 165 Sembach, K. R. et al. 2004, ApJS, 150, 387 Slavin, J.D., Shull, J.M., & Begelman, M.C. 1993, ApJ, 407, 83 Smith, M. et al. 2007, MNRAS, 379, 755 Sternberg, A., McKee, C.F. & Wolfire, M. 2002, ApJS, 143, 419 Sternberg, A. & Soker, N., 2008, MNRAS, 389, L13 Sun, X. H., Reich, W., Waelkens, A., Enslin, T. A. 2008, A&A, 477, 573 Sutherland, R.S., 2008, ApJ, in preparation Tamm, A., Tempel, E. & Tenjes, P. 2007, astro-ph/0707.4375 Teyssier, R. 2002, AAp, 385, 337 Thom, C., et al. 2008, astro-ph Tisserand, P. et al. 2007, A&A, 469, 387 Tripp, T. M., et al. 2003, AJ, 125, 3122 Wakker, B.P. 2001, ApJS, 136, 463 Wakker, B. P et al. 2007, ApJ, 207, 670, L113 Weiner, B.J., & Williams, T.B. 1996, AJ, 111, 1156 Weiner, B.J., Vogel, S.N., & Williams, T.B. 2002, Extragalactic Gas at Low Redshift, 254, 256 Westmeier, T. & Koribalski, B.S. 2008, MNRAS, 388, L29 Wilkinson, M.I. & Evans, N.W. 1999, MNRAS, 310, 645 Williams, J.P. & McKee, C.F. 1997, ApJ, 476, 166 Wolfire, M. et al. 1995, ApJ, 453, 673 Wolfire, M. et al. 2003, ApJ, 587, 278 This figure "Bland-Hawthorn_fig3.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0811.2467v2 This figure "Bland-Hawthorn_fig4.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0811.2467v2 This figure "Bland-Hawthorn_fig5a.gif" is available in "gif"(cid:10) format from: http://arxiv.org/ps/0811.2467v2 This figure "Bland-Hawthorn_fig5b.gif" is available in "gif"(cid:10) format from: http://arxiv.org/ps/0811.2467v2
astro-ph/9404058
1
9404
1994-04-22T19:13:30
The Properties of an HI-Selected Galaxy Sample
[ "astro-ph" ]
We analyze the properties of a sample of galaxies identified in a 21-cm, HI-line survey of selected areas in the Perseus-Pisces supercluster and its foreground void. Twelve fields were observed in the supercluster, five of them (target fields) centered on optically bright galaxies, and the other seven (blank fields) selected to contain no bright galaxies. We detected nine previously uncataloged, gas-rich galaxies, six of them in the target fields. We also detected HI from seven previously cataloged galaxies in these fields. Observations in the void covered the same volume as the twelve supercluster fields at the same HI-mass sensitivity, but no objects were detected. Combining our HI data with optical broad-band and H-alpha imaging, we conclude that the properties of HI-selected galaxies do not differ substantially from those of late-type galaxies found in optical surveys. In particular, the galaxies in our sample do not appear to be unusually faint for their HI mass, or for their circular velocity. The previously cataloged, optically bright galaxies in our survey volume dominate the total HI mass density and cross-section; the uncataloged galaxies contribute only 19% of the mass and 12% of the cross-section. Thus, existing estimates of the density and cross-section of neutral hydrogen, most of which are based on optically-selected galaxy samples, are probably accurate.
astro-ph
astro-ph
The Properties of an Hi-Selected Galaxy Sample Arpad Szomoru , Puragra Guhathakurta , Jacqueline H. van Gorkom ,  ;  Johan H. Knapen , David H. Weinberg , and Andrew S. Fruchter   ; 4 9 r p A 2 2 8 5 0 4 0 4 9 / h p - o r t s a        Kapteyn Astronomical Institute, P.O. Box , NL  AV Groningen, The Netherlands Hubble fellow, Princeton University Obs., Peyton Hall, Princeton, NJ , USA Current address: STScI,  San Martin Drive, Baltimore, MD , USA Department of Astronomy, Columbia University, New York, NY , USA Instituto de Astrof(cid:19)(cid:16)sica de Canarias, E- La Laguna, Tenerife, Spain Institute for Advanced Study, Princeton, NJ , USA Hubble fellow, Astronomy Dept., University of California, Berkeley, CA , USA Abstract We analyze the properties of a sample of galaxies identi(cid:12)ed in a -cm, Hi-line survey of selected areas in the Perseus-Pisces supercluster and its foreground void. Twelve (cid:12)elds were observed in the supercluster, (cid:12)ve of them (target (cid:12)elds) centered on optically bright galaxies, and the other seven (blank (cid:12)elds) selected to contain no bright galaxies within  of their centers. We detected nine previously uncataloged, gas-rich galaxies, six of them in the target (cid:12)elds. We also detected Hi from seven previously cataloged galaxies in these (cid:12)elds. Observations in the void covered the same volume as the twelve supercluster (cid:12)elds at the same Hi-mass sensitivity, but no ob jects were detected. Combining our Hi data with optical broad-band and H(cid:11) imaging, we conclude that the properties of Hi-selected galaxies do not di(cid:11)er substantially from those of late-type galaxies found in optical surveys. In particular, the galaxies in our sample do not appear to be unusually faint for their Hi mass, or for their circular velocity. We (cid:12)nd tentative evidence for a connection between optical surface brightness and degree of isolation, in the sense that low surface brightness galaxies tend to be more isolated. The previously cataloged, optically bright galaxies in our survey volume dominate the total Hi mass density and cross-section; the uncataloged galaxies contribute only (cid:24)  % of the mass and (cid:24) % of the cross-section. Thus, existing estimates of the density and cross-section of neutral hydrogen, most of which are based on optically-selected galaxy samples, are probably accurate. Such estimates can be used to compare the nearby universe to the high-redshift universe probed by quasar absorption lines.  Introduction Most investigations of the properties or the spatial distribution of galaxies begin with a sample selected in the optical, or sometimes in the infrared (based on IRAS [Neugebauer  et al.  ]  (cid:22)m detections). Even -cm or radio continuum studies usually target a sample of galaxies taken from optical catalogs, like the Uppsala Galaxy Catalog (UGC; Nilson  ) or the Catalog of Galaxies and Clusters of Galaxies (Zwicky et al.  {  ). Thus, any population of galaxies or gas clouds with very low optical luminosity or surface brightness could have largely escaped detection and study. To circumvent this problem, we have conducted a direct search in the -cm line of neutral hydrogen, using the D-con(cid:12)guration of the Very Large Array (VLA ; Napier  et al.  ), covering (cid:12)elds in the Perseus-Pisces supercluster and in its foreground void (see Haynes & Giovanelli   , hereafter HG). In an earlier paper (Weinberg et al.  , hereafter Paper I), we summarized the Hi properties of our detected ob jects, and we discussed the implications of our results for the shape of the Hi mass function and for the spatial distribution of dwarf galaxies. In this paper, we describe the -cm observations and results in greater detail, and we also present the results of follow-up optical observations, both broad-band and H(cid:11) imaging. These data allow us to compare directly the properties of our Hi-selected ob jects to those of optically selected galaxies. For the VLA survey, we observed twelve (cid:12)elds in the Perseus-Pisces supercluster and thirty (cid:12)elds in the foreground void. The e(cid:11)ective survey volume varies with Hi mass, since more massive ob jects can be detected farther from the center of the primary beam and at greater distances from the earth. Paper I discussed the resulting \selection function" of the survey and compared it to that of previous -cm searches, particularly that of Fisher & Tully ( ), which at the time was the largest directly comparable survey. We will not repeat the details herethe salient points are that our survey covered nearly equal volumes in the void and the supercluster, and that its total e(cid:11)ective volume, roughly  Mpc for M >  (cid:2)  M and  Mpc for M >  M , was much larger than HI (cid:12) HI (cid:12)    that of previous surveys for Hi masses in the  { M range typical of gas-rich dwarf (cid:12)  galaxies. Henning ( ) has recently published results of a much larger -cm survey,  which covers regions that are obscured by the Galactic plane at optical wavelengths. The sensitivity of these observations varied from one line-of-sight to another because of solar and man-made interference, dependence on declination and frequency resolution (Henning  ); we are therefore unable to make a quantitative comparison of survey volumes. The survey detected  extragalactic ob jects,  of them previously cataloged. Follow-up optical analysis of the sample is still in progress. For a review of other -cm surveys, see Briggs ( ). Our survey detected nine previously uncataloged galaxies, along with seven brighter galaxies that are listed in the UGC or the Zwicky catalog. Our de(cid:12)nition of supercluster and void (cid:12)elds is based on HG's redshift surveys of optically selected galaxies. This strategy allows us to address some questions of large-scale structure even with a small survey volume, but it complicates the interpretation of our results to some extent. In particular, it is unclear whether the optically cataloged galaxies should be considered part of our Hi-selected \sample", since they played a role in the choice of survey (cid:12)elds. Our approach in this paper is to present results for our full sample but to focus on  The VLA of the National Radio Astronomy Observatory (NRAO) is operated by Associated Universities, Inc., under a cooperative agreement with the National Science Foundation.  (cid:0) (cid:0) Unless stated otherwise, we have adopted a Hubble constant H =  km s Mpc in this paper.  the previously uncataloged ob jects as a distinct subset. All of the uncataloged ob jects are fainter (in B , R, and I ) than all of the previously cataloged galaxies, and for convenience we will hereafter refer to them as \Hi-dwarfs". However, only three of the nine have absolute B -magnitudes fainter than (cid:0), which is the boundary often used in conventional, optical de(cid:12)nitions of dwarf galaxies. The results of our survey, combined with the follow-up optical data, allow us to examine both the properties and the spatial distribution of Hi-selected galaxies. We are limited by small numbers in both cases, but because we are sensitive to ob jects that could have been entirely missed by earlier studies, even results with large statistical uncertainties are important; they provide a rough guide to previously unexplored territory. Although we covered equal volumes in the void and supercluster, all our detections were in supercluster (cid:12)elds. As discussed in Paper I, this result implies that the Hi-dwarfs trace the structure outlined by optically bright galaxies, at least qualitatively, though we cannot rule out a modest \bias" between the Hi-dwarfs and the giants. Our other signi(cid:12)cant (cid:12)nding in Paper I was that the number of ob jects detected was no more than one would predict by combining a standard estimate of the optical luminosity function with the Hi-mass- to-optical-luminosity relation that describes optically selected dwarf galaxies. From this statistical evidence, we inferred that our -cm search had not turned up a new population of gas-rich, optically faint ob jects, but that it had rather detected the \normal" population of irregular galaxies. The optical data in this paper allow us to address this point more directly. We describe our observing procedures in the next section, present our results in x, and discuss their implications in x.  Observations and Data Analysis .  cm Spectral Line The VLA spectral line observations took place during November   . The parameters of the observations are listed in Table . All  antennae were employed in D-con(cid:12)guration resulting in an angular resolution of about  (full width at half maximum [FWHM] of the synthesized beam). The bandwidth of the -cm observations was . MHz, corresponding to a usable velocity range of about  km s . On-line Hannning (cid:0) smoothing was used after which every other channel was discarded, resulting in sets of  independent channels and a velocity resolution of (cid:24) km s . The total observing time, (cid:0) including time spent on phase and (cid:13)ux calibration, was  hr. All the VLA pointings were located within the strip < (cid:11) <    and   h m s h m s o o o o  < (cid:14) <  . The pointings are listed in Table . For each (cid:12)eld we made  (cid:2)    images, mapping out the full extent of the primary beam. Twelve (cid:12)elds were observed in the Perseus-Pisces supercluster at a redshifted velocity of  km s , and  (cid:12)elds in the (cid:0) foreground void at  km s . The volume surveyed in a single pointing scales as D , (cid:0)  where D is the distance corresponding to the mean velocity to which the VLA bandpass is tuned; the number of supercluster and void (cid:12)elds were chosen so that the total volume surveyed in each region was roughly the same. The integration times t were  min for int each of the void (cid:12)elds and  min for each of the supercluster (cid:12)elds, roughly in the ratio    of D to ensure equal Hi-mass sensitivity in the void and supercluster (M / (cid:27)D and HI (cid:0)= (cid:27) / t , where (cid:27) is the r.m.s. noise in the -cm data). int Five of the  supercluster (cid:12)elds were centered on Perseus-Pisces galaxies brighter than M = (cid:0): (target (cid:12)elds). These optically bright galaxies were chosen from the B redshift catalog of HG. The remaining seven supercluster (cid:12)elds were selected to have no bright galaxies within  of their pointing centers (blank (cid:12)elds). The UV data were calibrated and processed into ((cid:11); (cid:14); V ) datacubes using standard techniques in NRAO's Astronomical Image Processing System. To achieve the best possible signal-to-noise ratio, natural weighting was applied in making the cubes. Subsequent analysis of the VLA datacubes was carried out using the Groningen Image Processing System (GIPSY). We searched for Hi line emission by subtracting the -cm continuum emission from each cube; preliminary estimates of the continuum map were obtained by () averaging channels { of the datacube, and () averaging channels { and {. If line emission was detected, the continuum subtraction was redone using the average of the line-free channels on the low and high frequency side of the line emission. In a few cases, the continuum was subtracted by (cid:12)tting a linear baseline to the line-free channels. Images with Hi emission were CLEANed and restored with a gaussian beam. The r.m.s. noise (cid:27) =  mJy/beam in the (cid:12)nal images of the void (cid:12)elds and . mJy/beam in the supercluster (cid:12)elds. Total Hi images were constructed by blanking areas without line emission and summing the channel maps. Neutral hydrogen masses were calculated using the formula: M =M = : (cid:2)  D S dV HI (cid:12) Z   where D is the distance in Mpc, and the (cid:13)ux density S (in Jy) is integrated over the -cm line, with the velocity V expressed in km s . The mean heliocentric velocity V hel (cid:0) of each galaxy was determined from its global Hi line pro(cid:12)le by averaging the velocities at which the (cid:13)ux density falls to % and % of the peak value on each side of the -cm line. We compute the distance to each galaxy by correcting its heliocentric velocity for the motion of the sun with respect to the velocity centroid of the Local Group: D = [V +  sin (l) cos(b)]=H , where l and b are the Galactic longitude and latitude of the hel galaxy, and H is the Hubble constant. The observed line widths at % and % of the peak value, W and W , are listed   in Table . Also listed is log(W ), which is the logarithm of W , corrected for instrumental R  i broadening following Bottinelli et al. ( ), for turbulent broadening following Tully & Fouqu(cid:19)e ( ) and corrected to edge-on. The magnitude of the instrumental broadening correction, which is based on a statistical intercomparison of line pro(cid:12)les measured with various velocity resolutions, is (cid:0). km s . (cid:0) The ma jority of the uncataloged galaxies is barely resolved in Hi. An estimate of their angular size was obtained by (cid:12)tting a two-dimensional gaussian to the total Hi distribution; the sizes quoted for the Hi-dwarfs are the FWHM of these gaussians with the FWHM of the beam subtracted in quadrature. For the bright galaxies, the Hi column density contour at  (cid:2)  cm was used to de(cid:12)ne the angular size. It should  (cid:0) be kept in mind, however, that the size determinations for the smaller galaxies are only rough indications of the true sizes because of the marginal angular resolution. Intensity-  weighted (cid:12)rst moment maps were used to construct velocity (cid:12)elds for the large galaxies in which the Hi distribution was well resolved both spatially and in velocity. . Optical Broad-band Optical observations of the galaxies detected in Hi were carried out during the second week of October  using the -m Jacobus Kapteyn Telescope (JKT) on the island of La Palma. The faintest of the galaxies are uncataloged and are barely visible on the Palomar Optical Sky Survey plates. The camera on the JKT consisted of a GEC charge-coupled device (CCD) with  (cid:2)  pixels and a pixel scale of : . Broad-band optical images were obtained through B ( A), R ( A), and I ( A) (cid:12)lters. (cid:23) (cid:23) (cid:23) Integration times were  (cid:2) s in B ( (cid:2) s for the faintest ob jects), s in R, and  (cid:2) s in I . The seeing ranged from : to  , and the sky conditions were photometric during most of the observations. We obtained short exposures of bright standard stars (Landolt  ) at regular intervals throughout the photometric nights, and we used these to photometrically calibrate the JKT images. All the galaxies were observed close to transit so that airmass corrections were minimal. The estimated errors in the magnitude zero points are B = : mag, R = : mag and I = : mag. Twilight sky (cid:13)ats were used for (cid:13)at (cid:12)elding the CCD images. Sky subtraction was done by (cid:12)tting (cid:12)rst- and second-order polynomial surfaces to blank regions of sky around the image of each galaxy. Three of the galaxies, UGC  b, PPHI +, and PPHI +, are too faint to be detected in the JKT B images. These ob jects were reobserved in service time with the .-m William Herschel Telescope (WHT) in November  using the Taurus instrument in imaging mode. A  (cid:2)  EEV CCD was employed, with a pixel size of : . One  s exposure was obtained for each of the three galaxies through a B (cid:12)lter. Twilight sky (cid:13)ats were used for (cid:13)at (cid:12)elding the CCD frames. An artifact in the sky (cid:13)at image a(cid:11)ected the (cid:13)at (cid:12)elding of the UGC  b image; the (cid:13)ux determination of this galaxy is therefore uncertain. After sky subtraction, the WHT frames were calibrated by a bootstrap procedure: the (cid:13)uxes of some bright stars in the WHT images were compared with the photometrically calibrated (cid:13)uxes of the same stars in the JKT B images. The total magnitude of each galaxy was calculated by integrating over a circular aperture of radius  , except for UGC  for which an aperture of radius  was used. In a few cases, images of foreground stars superposed on the galaxy images were removed; imperfect star subtraction is an additional source of uncertainty in the (cid:13)ux measurement of these galaxies. Magnitudes were corrected for absorption by dust along the line-of-sight in the Galaxy following the precepts of Burstein & Heiles ( ). For comparison with existing samples, we have also applied standard corrections for internal absorption as a function of wavelength and disk inclination (Tully & Fouqu(cid:19)e  ). In order to derive surface photometry of the galaxies, we (cid:12)tted ellipses to a succession of isophotes of the I -band light distribution (of decreasing brightness and increasing radius). Radial brightness pro(cid:12)les in the B , R, and I bands were constructed by integrating the light in these images within the best-(cid:12)t ellipses. The brightness pro(cid:12)les of the galaxies are well approximated by exponential laws and these were used to obtain estimates of the central surface brightness (cid:22) , the disk scale length h, the diameter D   of the B =  mag arcsec isophote, and the B (cid:0) R and R (cid:0) I color as a function of (cid:0) radius. The best-(cid:12)t ellipses also yielded an axis ratio and a position angle for each galaxy. Disk inclinations were derived from the isophotal axis ratios by accounting for the (cid:12)nite thickness of stellar disks, assuming an intrinsic axis ratio of . and correcting for the  o measurement bias (Tully  ). . Narrow-band H(cid:11) The narrow-band H(cid:11) imaging observations were done on the Shane -m telescope at Lick Observatory using the UV Schmidt Spectrograph. Due to limited observing time, only four ob jects were observed in H(cid:11): UGC a, UGC b, UGC b, and UGC  a. Line (cid:13)ux was detected in all four galaxies. The Cassegrain spectrograph imaging system allowed one to switch between CCD imaging and spectroscopy through the interposition of a (cid:13)at mirror in the spectroscopic light path. One could therefore use the spectrograph to directly measure the transmission function of the imaging (cid:12)lter, although a small uncertainty remains due to possible spectral dependence of the re(cid:13)ectance of the (cid:13)at mirror. The H(cid:11)+Nii emission of all four galaxies observed fell within the passbands of two Lick narrow-band (cid:12)lters, which have e(cid:11)ective wavelengths of  A and   A and FWHM of  A and  A, respectively. Continuum (cid:23) (cid:23) (cid:23) (cid:23) (o(cid:11)-band) images of the galaxies were obtained with a  A-wide (cid:12)lter centered at about (cid:23)  A. Integration times ranged from  s to  s for the H(cid:11) exposures, and from (cid:23)  s to  s for the continuum exposures. Very light cirrus was present throughout the H(cid:11) observations, so that our photometry is no better than {%. As a consequence of these non-photometric conditions, we had to determine the appropriate scaling for continuum subtraction on an image-by-image basis by minimizing the residuals of bright stars. We estimate that this empirical method of continuum subtraction provides H(cid:11)+Nii (cid:13)uxes to an accuracy of about %. The measured H(cid:11)+Nii (cid:13)uxes were corrected for Galactic absorption (Burstein & Heiles  ; Rieke & Lebofsky  ) and used to estimate star formation rates (SFR) following Hunter & Gallagher ( ). The H(cid:11) luminosity of a galaxy L was converted H(cid:11) to the total number of ionizing photons, which was then compared to the integrated number of photons expected from massive stars for a Salpeter ( ) initial mass function. Extrapolating the Salpeter stellar mass function down to a lower mass limit of : M , (cid:12) the SFR of all stars (with masses in the range .{ M ) is given by: (cid:12) _ (cid:0) (cid:0) M = : (cid:2)  L M yr ; H(cid:11) (cid:12) where L is in units of erg s . Based on this SFR, we calculated the distance- H(cid:11) (cid:0) independent quantities M =L and the timescale (cid:28) over which the current gas supply B _ in a galaxy would last at the current rate of star formation. The gas depletion time (cid:28) is computed by multiplying the Hi mass by ., to account for the presence of He, and dividing it by the SFR.   Results Sixteen galaxies were detected in Hi in our VLA observations of  (cid:12)elds in the ridge of the Perseus-Pisces supercluster region. Of these,  Hi detections are located in `target' (cid:12)elds centered on optically bright galaxies, while the remaining three are in `blank' (cid:12)elds away from bright galaxies (see x.). Nine of the  galaxies detected at  cm do not have counterparts in optical catalogs; we refer to these galaxies as \Hi-dwarfs" regardless of their optical luminosity. Six of these Hi-dwarfs were found in target (cid:12)elds {, and the remaining three in blank (cid:12)elds. Two optically faint, but previously cataloged galaxies, Zw  . and Zw ., were detected in Hi in target (cid:12)elds  and , respectively. One of the targets, the Sa galaxy UGC  , was not detected in Hi, while the (cid:12)fth target (cid:12)eld contains two bright galaxies (UGC  and UGC ). We adopt the following convention: () uncataloged galaxies near bright galaxies (i.e. in target (cid:12)elds) are denoted by the UGC number of the bright galaxy with \a" or \b" appended, in order of increasing right ascension; and () uncataloged galaxies in blank (cid:12)elds are named \PPHI" followed by their   right ascension and declination. No Hi detections were made in the  void (cid:12)eld observations to a (cid:27) (cid:13)ux density limit of  mJy/beam, corresponding to an Hi mass per channel of (cid:24)  (cid:2)  M at the center (cid:12)  of the primary beam. The noise in our datacubes is somewhat non-gaussian because of imperfect subtraction of remote continuum sources; the short duration of the VLA void observations (and the resulting sparse UV coverage) makes it practically impossible to remove these sources exactly. The Hi and radio continuum properties of the galaxies detected in the supercluster are summarized in Table . Their optical magnitudes, colors, and morphological properties, based on BRI data, are listed in Table . Table  contains the results of the H(cid:11) observations, along with other parameters derived from the various datasets, including surface brightness, size, Hi-mass-to-light ratio, and gas depletion timescale. Some of the Hi parameters in Table  (e.g. velocity width and total Hi mass) di(cid:11)er slightly from those given in Paper I. These di(cid:11)erences are a result of redoing a large part of the analysis using the completely rewritten, UNIX-based version of GIPSY; this allowed us to take advantage of several convenient features in the new version and corrected a minor software error in the old package. In nearly every case, the values of the Hi parameters have remained equal to the previously published value to within the (cid:27) uncertainty. The velocity width of UGC  at % (W ) was listed incorrectly in Table  of Paper I, and this has been  corrected in Table  of this paper. We present global Hi pro(cid:12)les in Figure  of all  galaxies detected at  cm: four galaxies in (cid:12)eld  (top left panel), three in (cid:12)eld  (top right panel), two in (cid:12)eld  (center left), two in (cid:12)eld  (center right), and two in (cid:12)eld  (bottom left). The line pro(cid:12)les of the three Hi-dwarfs detected in blank (cid:12)elds are shown together in the bottom right panel of Figure . Figure  shows the velocity (cid:12)elds of six large galaxies, those in which the Hi is well resolved in position and velocity. These include the target galaxies of (cid:12)elds , , , and  (the target galaxy of (cid:12)eld , UGC  , was not detected in Hi), along with UGC  (the second bright galaxy in (cid:12)eld ) and Zw . ((cid:12)eld ). Contour plots of the Hi column density distributions in the supercluster target (cid:12)elds { are shown in  Figures {. Figure  (four panels) and Figures { show contour plots of the total Hi distribution in some of the individual galaxies superposed on gray-scale representations of their B -band images. The shaded ellipses in Figures { indicate the FWHM of the VLA synthesized beam. In the rest of this section, the galaxies detected in each of the supercluster (cid:12)elds are discussed in turn. . Field  The central galaxy of this (cid:12)eld is UGC , an SBb galaxy. In addition to UGC , three companions were detected around this galaxy in our -cm observations, one of which was found to be a cataloged galaxy, Zw  . (Figs.  and ). The four galaxies are at nearly identical redshifts, the largest velocity di(cid:11)erence being about  km s (top left (cid:0) panel of Fig. ). The distribution of neutral hydrogen in UGC  is symmetric (lower left panel of Fig. ) and its velocity (cid:12)eld regular (top left panel of Fig. ); its double-peaked global pro(cid:12)le indicates a disk in di(cid:11)erential rotation. The other three galaxies are barely resolved in Hi. Figure  shows the Hi and blue light distributions of all four galaxies in (cid:12)eld . The optical surface brightness pro(cid:12)les of UGC  deviate signi(cid:12)cantly from an exponential law due to the e(cid:11)ects of a central bar and prominent spiral arms (lower left panel of Fig. ). As a consequence, the extrapolated central surface brightness and the exponential scale length of the stellar disk are sub ject to systematic error. In the case of Zw  . , a bright foreground star is seen superposed on its disk, and this makes the measurement of its isophotal axis ratio and position angle uncertain. Spiral arms are discernible in the optical images of Zw  . and, as in the case of UGC , its disk parameters cannot be determined accurately. The distribution of starlight in the uncataloged galaxy UGC a has a very small scale length, and the galaxy appears to be featureless in both optical and Hi. Given its very blue B (cid:0) R color, it could be classi(cid:12)ed as a compact blue dwarf. The other Hi-dwarf in this (cid:12)eld, UGC b, is slightly asymmetric and has an exponential stellar disk without a clear bulge component. The Hi-dwarfs UGC a and UGC b were observed in H(cid:11). The morphology of the line emission (and hence of the ionized gas) in UGC a is compact and smooth. By contrast, the H(cid:11) distribution in UGC b is clumpy and extended, indicating that massive star formation is taking place throughout the disk of this galaxy. . Field  The VLA observations of this (cid:12)eld were centered on the Sb galaxy UGC . The target galaxy was detected along with two gas-rich galaxies around it (Fig. ). One of these, UGC a, is close to UGC  and may be interacting with it, while the other, UGC b, is about  from the target galaxy. The neutral hydrogen distributions of UGC  and UGC a overlap partially, with a hint of an Hi bridge between them and a faint extension on the opposite side of the larger galaxy (possibly a tidal feature). The velocity (cid:12)elds of both galaxies are shown in the top right panel of Figure . The iso-velocity contours of the main galaxy UGC  appear to be twisted, suggestive of an interaction. Kinematically,  the companion galaxy UGC a is distinct from UGC , with its ma jor axis at right angles to the ma jor axis of the latter. We were unable to derive optical surface photometry of the close companion UGC a because a bright foreground star happens to lie at the position of the peak of the Hi emission. The other Hi-dwarf UGC b was studied in BRI and in H(cid:11) (see Fig. ). Its optical surface brightness pro(cid:12)le is exponential with no sign of a bulge component. The H(cid:11) emission is smoothly distributed and somewhat extended, with the peak slightly o(cid:11)set from the optical center of the galaxy. . Field  The central galaxy of (cid:12)eld , the Sa galaxy UGC  , was not detected in Hi. Two gas-rich galaxies were detected in this (cid:12)eld, at large velocities relative to the central galaxy: V =  km s and  km s for the Hi-dwarfs UGC  a and UGC  b, hel (cid:0) (cid:0) respectively (see Fig. ), compared to  km s for UGC   (obtained from RC). The (cid:0) distribution of neutral hydrogen in UGC  a is practically unresolved, and is somewhat o(cid:11)set from the optical centroid (Fig. ). The Hi in the other Hi-dwarf UGC  b seems to consist of two components, separated by about  but at the same redshift (Figs.  and ). The smaller component of UGC  b, though faint, is seen in three independent velocity channels and is not a noise artifact. Both Hi-dwarfs in this (cid:12)eld have low central B -band surface brightnesses, UGC  b being the faintest of all the Hi-dwarfs in our sample in this respect. The optical morphology of UGC  a is slightly asymmetrical with a roughly exponential surface brightness pro(cid:12)le, and it has a smooth, centrally concentrated H(cid:11) distribution. Seen at optical wavelengths, UGC  b shows an irregular structure with no clearly de(cid:12)ned center. The optical emission coincides with the peak of the larger Hi component. The inclination, position angle, and radial color pro(cid:12)les of UGC  b, derived by (cid:12)tting ellipses to the optical images, are not very accurate due to the amorphous nature of this galaxy. The B -band image of UGC  b could not be (cid:13)at (cid:12)elded properly; its total blue magnitude is therefore not well determined. . Field  The target galaxy of this (cid:12)eld is UGC , an S galaxy. The only other -cm detection in this (cid:12)eld is a galaxy from the Zwicky catalog, Zw . (Fig. ). The central galaxy UGC  has an Hi disk in di(cid:11)erential rotation, as the double peaked-global Hi pro(cid:12)le in the center right panel of Figure  shows. The asymmetry of the line pro(cid:12)le and the curving of the kinematical ma jor axis (center left panel of Fig. ) could indicate some disturbance in UGC . The Hi distribution of Zw . is irregular in appearance and is slightly o(cid:11)set with respect to the optical image (Fig. ). In spite of this, both the global pro(cid:12)le (center right panel of Fig. ) and the velocity (cid:12)eld (center right panel of Fig. ) are regular and symmetrical, with the iso-velocity contours of the gas aligned with the optical minor axis. The optical image of Zw . shows a normal disk galaxy, with an exponential light pro(cid:12)le. No H(cid:11) observations were made of the galaxies in this (cid:12)eld. . Field  This (cid:12)eld contains two optically bright galaxies, UGC  and UGC . The galaxies are classi(cid:12)ed as Sc and SBa-b, respectively. Despite their proximity to each other in position and velocity space, both galaxies have regular Hi distributions (Fig. ) and kinematics (bottom left panel of Fig. ; bottom two panels of Fig. ) and show no obvious signs of interaction. Solar interference during the VLA observations of this (cid:12)eld made it necessary to discard a large fraction of the UV data. Consequently, the sensitivity in this (cid:12)eld is lower than in other (cid:12)elds, with an r.m.s. noise level of . mJy/beam. We did not observe these two galaxies in H(cid:11). . Fields , , and  Hi-dwarfs were found in each of these three blank supercluster (cid:12)elds. Two of the detected Hi-dwarfs, PPHI + ((cid:12)eld ) and PPHI + ((cid:12)eld ), have asymmetrical Hi distributions that are o(cid:11)set slightly with respect to the optical image (Figs.  and ). The northward extension of the gas in PPHI + is seen in only one channel and may not be real. Both PPHI + and PPHI + are low surface brightness systems and have irregular optical morphology. The Hi-dwarf detected in (cid:12)eld , PPHI +, is somewhat more regular than the other two Hi-dwarfs in both its Hi and light distributions (Fig. ). Its stellar disk has an exponential brightness pro(cid:12)le. None of the three Hi-dwarfs in the blank (cid:12)elds were observed in H(cid:11).  Discussion We discussed the spatial distribution of our Hi-dwarfs at some length in Paper I; we have little to add here on the sub ject, since we have no new data on the spatial distribution. We found nine Hi-dwarfs in the supercluster and none in the void, so the basic message is that Hi-dwarfs trace the structure outlined by optically bright galaxies. The correlation seems to continue down to small scales within the supercluster, since we found six of our nine Hi-dwarfs in the (cid:12)elds targeted on bright galaxies despite observing more blank (cid:12)elds than targeted (cid:12)elds. This is the sort of behavior we might expect based on the galaxy autocorrelation function, which continues as a power law down to very small scales. With only nine ob jects, we are unable to rule out a modest bias between the clustering of bright galaxies and the clustering of Hi-dwarfs (see Paper I for details), but it is clear that the most likely place to (cid:12)nd a gas-rich Hi-dwarf is near an optically bright giant. The optical and H(cid:11) data reported in x allow us to compare the properties of our Hi- selected sample to those of more traditional, optically selected samples. We provide such a comparison in x. below. We then describe possible environmental e(cid:11)ects within our sample and discuss the implications of our results for the Hi mass density and cross-section in the local universe.  . Comparison to Other Samples .. Luminosities and Star Formation Rates The distinguishing feature of our sample of Hi-dwarfs is that it is not pre-selected in the optical, so the most interesting question to ask about it is whether the ob jects detected in this way are anomalously faint. There are at least two sensible ways to pose this question: are the Hi-dwarfs unusually faint for their Hi masses, and are they unusually faint for their circular velocities? The (cid:12)rst of these questions can be rephrased in terms of the Hi-mass-to-light ratios M =L , which are listed in column  of Table . Values for the Hi-dwarfs range from HI B . to ., with a mean of . (we exclude UGC a, for which we did not obtain L ). B For their sample of  irregular galaxies in the Virgo cluster (excluding (cid:12)ve that were not detected in Hi), Gallagher & Hunter (  ) (cid:12)nd ratios ranging from . to ., with a mean of . . In Haynes & Giovanelli's ( ) sample of isolated galaxies, the M =L HI B ratios for the  galaxies of type later than Sc (speci(cid:12)cally, those assigned a numerical type { in Table I of their paper) range from . to ., with a mean of .. Two galaxies in the Haynes & Giovanelli sample have M =L larger than the extreme value in HI B our sample, that of PPHI +. It appears that the M =L ratios of our Hi-selected HI B galaxies are not extraordinarily high relative to those of optically selected samples of late- type galaxies. Certainly, -cm selection has not turned up any population resembling star-free gas clouds. There is some evidence for an increase in M =L in lower density HI B environments: our supercluster has a higher mean ratio than Gallagher & Hunter's Virgo cluster sample, Haynes & Giovanelli's sample of isolated galaxies has a still higher mean ratio, and within our sample, the blank-(cid:12)eld Hi-dwarfs have high M =L relative to HI B Hi-dwarfs in target (cid:12)elds. We can address the second of the questions posed above by plotting our sample of galaxies on the Tully-Fisher ( ) relation, as shown in Figure . The dashed line in this Figure represents the relation derived by Pierce & Tully ( , hereafter PT). For consistency, we have corrected the absolute magnitudes M listed in column  of Table  b;i (cid:0) (cid:0) B to the Hubble constant H =  km s Mpc derived by PT. Linewidths, corrected for inclination and internal motions, are from column  of Table  (see x.). The Hi-dwarfs in our sample follow the Tully-Fisher relation derived from optically selected galaxies fairly well; they are not unusually faint for their circular velocity. The three Hi-dwarfs detected in the blank (cid:12)elds all lie somewhat below the PT relation (the possible implications of this are discussed in x.). An unweighted least-squares (cid:12)t to all of the galaxies in our sample yields a slope of (cid:0): (cid:6) :, more or less consistent with PT's value of (cid:0): (cid:6) :. This (cid:12)t is plotted as a dotted line in Figure . The vertical scatter about the best-(cid:12)t relation is . magnitude. This scatter is higher than that found for optical samples, but it could arise largely from errors in the linewidths and inclinations of the small systems in our sample. Kunth ( ) and Tyson & Scalo ( ) suggest that the ma jority of dwarf galaxies may be in a quiescent state, not forming young stars, and that they might consequently be missed by optical selection methods. We have already seen that the Hi-dwarfs in our sample are not unusually faint, and we can address the star formation issue directly for the  four Hi-dwarfs that we observed in H(cid:11). The H(cid:11) equivalent widths (column  of Table ) are similar to those of typical late-type galaxies (Kennicutt & Kent  ), and the M =L B _ ratios and the timescales for gas depletion (columns  and  of Table ) are similar to the values derived by Gallagher & Hunter (  ) for their sample of Virgo irregulars. The gas depletion timescales are on the order of a Hubble time, so these Hi-dwarfs do not seem to be in the midst of star formation bursts. In Paper I, we argued on the basis of the number of detected ob jects that our Hi- selection procedure had not turned up a population of optically under-luminous galaxies. The optical data presented here con(cid:12)rm this conclusion; our Hi-dwarfs appear to be garden variety late-type galaxies. If there is a large population of optically faint galaxies lurking in the shadows, they must either have very little gas, or they must store that gas in ionized or molecular form. .. Colors and Ages To our knowledge, no published BRI magnitudes exist for large samples of dwarf galaxies. In Table , therefore, we compare the colors of our galaxies to those of the statistically complete, diameter-limited galaxy sample of de Jong and van der Kruit (in preparation). They have obtained B V RIHK surface photometry of  galaxies, ranging in morphological type from Sa to Irr. Table  lists the mean and dispersion of the B (cid:0) R and R (cid:0) I colors for their sample (the \control sample") and for a subset of this sample with morphological types Sd and later. We also list the mean colors and dispersions in Table  for those galaxies in our sample which have BRI photometry: eight of the nine Hi- dwarfs and three of the eight cataloged galaxies (see Table ). Our sample is somewhat bluer than the control sample in B (cid:0) R, but the di(cid:11)erence is not large compared to the dispersion within the control sample. Hunter & Gallagher ( ) (cid:12)nd that optically selected dwarf galaxies tend to be bluer than bright galaxies, and the same trend seems to hold for Hi-selected galaxies. Figure  plots the galaxies for which we have BRI photometry in a color-color diagram, with the Hi-dwarfs as open squares and the cataloged galaxies as (cid:12)lled triangles. For comparison, the open circles and connecting line show the galaxy evolution model of Mazzei et al. ( ). The model incorporates stellar evolution, chemical evolution of the stellar population, and absorption and re-emission of starlight by interstellar dust. The model does not (cid:12)t the observational data very well, and two colors are in any case not enough for accurate age determination. However, it does appear that the galaxies in our sample are fairly young, and that most of the Hi-selected galaxies are younger than the cataloged galaxies. . Environmental E(cid:11)ects In Figure , we plot the central B surface brightness (upper panel) and the logarithm of the total blue luminosity (lower panel) of all galaxies that were not (cid:12)eld targets against the logarithm of the pro jected distance to the (cid:12)eld's target galaxy (or, in the case of the blank- (cid:12)eld Hi-dwarfs, the pro jected distance to the nearest galaxy in HG's redshift catalog). Cataloged galaxies are shown by triangles, Hi-dwarfs by squares. Surface brightnesses  come from column  of Table , and luminosities from column  of Table  (see x. for details). In both cases, there appears to be a modest trend, with fainter galaxies having larger pro jected separations. We can make this trend a bit more convincing if we distinguish between Hi-dwarfs that are gravitationally bound to bright galaxies and Hi-dwarfs that are not. We can reasonably assume that the Hi-dwarfs in blank (cid:12)elds are not bound, but the question is trickier for Hi-dwarfs in target (cid:12)elds. We will try to make the distinction by computing dynamical mass-to-light ratios on the assumption that the Hi-dwarfs are bound, then asking whether the resulting values are \plausible." We use two di(cid:11)erent techniques to estimate mass-to-light ratios. The (cid:12)rst employs the \average" mass estimator of Heisler et al. ( ) for galaxy groups: M = (V (cid:0) V ) R : est zi zj ?;ij f Av X X  GN (N (cid:0) ) i j<i Here N is the number of galaxies in the group, V is the velocity along the line of sight, zi R is the pro jected distance between galaxies i and j , G is the gravitational constant, ?;ij f is a scaling factor, and the sums are carried out over all group members. Heisler Av et al. ( ) calibrate this and other, similar mass estimators using model groups with N =  and N = . For N = , they (cid:12)nd that % of the mass estimates lie within a factor  of the correct value. Applying the estimator to the nearby groups cataloged : by Huchra & Geller ( ), they (cid:12)nd a median M=L of about  M =L for groups (cid:12) (cid:12) : with three members. The mass estimates for our (cid:12)ve groups (i.e. our (cid:12)ve target (cid:12)elds) appear in column  of Table . The corresponding mass-to-light ratios appear in column . While four of the (cid:12)ve groups have what might be called reasonable M=L values, the high value in (cid:12)eld ,   M =L , makes it unlikely that the galaxies in this (cid:12)eld form a gravitationally bound (cid:12) (cid:12) group. Even for the bound groups, the mass estimates in Table  have large statistical uncertainties because of the small numbers of group members. To avoid these large uncertainties, we can take a more conservative approach and compute the minimum mass that would be required to bind each companion to its bright target galaxy. We obtain this mass by assuming that the line-of-sight velocity di(cid:11)erence (cid:1)V and pro jected radial distance (cid:1)R represent the true, three-dimensional values. These quantities are listed in columns  and  of Table . The minimum mass follows from the inequalities K E (cid:20) (cid:0)P E =) ((cid:1)V ) (cid:20) ; M M GM M       (M + M ) (cid:1)R   where M and M are the masses of the central galaxy and the companion, respectively.   This inequality implies a minimum value of (M + M ):     M = ((cid:1)V ) (cid:1)R : min G Column  of Table  lists this minimum pair mass for each companion galaxy (plus the central galaxy), and column  lists the mass-to-light ratio obtained by dividing this  minimum mass by the total blue luminosity of the pair. Again, we conclude that the galaxies in (cid:12)eld  are probably not gravitationally bound, while the remaining galaxies could plausibly be bound. This conclusion has an interesting implication for the low surface brightness systems in our sample. Three of the Hi-dwarfs have very low surface brightness ((cid:22) > ); two of these were found in blank (cid:12)elds, the third in (cid:12)eld . This point is illustrated in Figure , where we use open symbols to denote \unbound" systems, i.e. blank-(cid:12)eld Hi-dwarfs and the Hi-dwarfs in (cid:12)eld , and (cid:12)lled symbols to denote bound systems. The average (cid:22) of the unbound systems is ., with a dispersion of .; for the bound systems, the average is ., with a dispersion of .. The mean and dispersion in values of log(L ) B are : (cid:6) : for the unbound systems and : (cid:6) : for the bound systems. Although the statistics are poor, there is at least a hint in our data that the presence of a neighboring bright galaxy in(cid:13)uences surface brightness and luminosity. In particular, all of our low surface brightness systems are unbound, consistent with the (cid:12)ndings of Bothun et al. ( ), though not all of our isolated systems have low surface brightness. . Hi Cross Section and Mean Density in the Local Universe Our -cm survey provides us with a picture of the distribution of neutral gas in regions of the nearby universe. This picture can be compared to the distribution of neutral hydrogen inferred from quasar absorption lines at high redshifts. Most relevant to our survey are the damped Ly(cid:11) absorbers (Wolfe  ; Lanzetta et al.  ), which have neutral hydrogen column densities similar to those of the systems that we detect at  cm. The question of what constitutes the parent population of the damped Ly(cid:11) systems remains a matter of debate. It has been suggested that the absorption lines could arise in extended galactic disks (Wolfe  ), galactic halos (Bahcall & Spitzer   ), dwarf galaxies (York et al.  ; Tyson  ), or tidal features around interacting and merging galaxies (Carilli & van Gorkom  ). By most accounts (cf. Lanzetta et al.  ; Rao & Briggs  ), the neutral hydrogen contained in damped Ly(cid:11) systems greatly exceeds that in present-day galaxies, by factors of (cid:24) {. However, the estimate of the hydrogen content at the current epoch is based on observations of gas associated with optically cataloged galaxies. By comparing theoretical models to observations, Tyson & Scalo ( ) conclude that many dwarf galaxies might escape optical detection. Tyson ( ) argues that this population of faint dwarfs could produce the observed damped Ly(cid:11) absorption without strong evolution in the neutral gas fraction. Arguments like those of Tyson & Scalo ( ) provided much of the motivation for our -cm survey. However, we found in Paper I that the Tyson & Scalo ( ) model predicted far more dwarfs than we observed, unless we assumed that the ratio of Hi mass to optical luminosity was much smal ler for \quiescent" dwarfs than for bright Hi-dwarfs. In a similar vein, we have seen in x. that our Hi-selected galaxies are not anomalously faint, so estimates of the local neutral gas density based on optically selected samples, (cid:18)a la Lanzetta et al. ( ) and Rao & Briggs ( ), should be quite safe. Our survey is not sensitive to Hi masses much below  M , but other \blind" -cm surveys (cid:12)   have covered this regime. Although these surveys cover smaller volumes than our survey, they are probably large enough to rule out a steep rise in the Hi mass function between    M and (cid:24)  M (Fisher & Tully  ; Ho(cid:11)man et al.   ; Briggs  ; Ho(cid:11)man (cid:12) (cid:12) et al.  ). We could attempt to estimate the space density and cross-section of neutral hydrogen in the local universe directly from our sample. Unfortunately, our survey volume is small, and our (cid:12)elds are not randomly chosen, so such an estimate would have large statistical uncertainty and a systematic bias that would be di(cid:14)cult to evaluate. However, we can say that within our survey region, the Hi-dwarfs contribute only  % of the Hi mass and % of the Hi cross-section. Once again, the message is that the Hi content of the nearby universe is dominated by relatively bright galaxies. Lanzetta ( ) notes the suggestive coincidence that the mass seen in gas at high redshifts is equal to the luminous mass seen at z = , and he therefore proposes that the gas observed in the damped Ly(cid:11) systems has formed into the stars of present-day galaxies. Our results support this inference, albeit indirectly. . Summary Galaxies can \shine" at both optical and -cm wavelengths, and in principle one might (cid:12)nd rather di(cid:11)erent ob jects in these two regions of the spectrum. However, our Hi survey of the Perseus-Pisces supercluster and void has not turned up any remarkable population of unusual ob jects. Most of the neutral hydrogen gas is associated with big, optically bright galaxies. Sensitive -cm observations reveal a substantial number of smaller ob jects, but these appear to be normal irregular galaxies, and their spatial distribution is not radically di(cid:11)erent from that of the optically bright giants. The distribution of Hi in an individual galaxy or galaxy group may be quite di(cid:11)erent from the distribution of stars, but in its broad features, the universe seen at  cm is a familiar one. Acknow ledgments. We are grateful to Martha Haynes and Riccardo Giovanelli for helpful discussions and for allowing us to use their redshift data, in advance of publication, to select our (cid:12)elds. We also thank Richard McMahon for providing us with APM data that were used to determine star positions on some of our JKT frames, and Ronald Hes for excellent support during the early phase of this pro ject. We thank Hy Spinrad and Dave Cudaback for the loan of narrow-band (cid:12)lters used to make the H(cid:11) measurements. Finally, we thank the NRAO for generous allocation of observing time. This research was supported in part by NSF grants AST  - and AST - to Columbia University, under grant no. -- by the Netherlands Foundation for Research in Astronomy (NFRA), which receives its funds from the Netherlands Organization for Scienti(cid:12)c Research (NWO). ASF and PG were supported by NASA through grants HF-.- A and HF-.- A, respectively, from the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS-. DW acknowledges support from the W.M. Keck Foundation and from NSF grant PHY -. The Jacobus Kapteyn Telescope (JKT)  and the William Herschel Telescope (WHT) are operated on the island of La Palma by the Royal Greenwich Observatory in the Spanish Observatorio del Roque de los Muchachos of the Instituto de Astrof(cid:19)(cid:16)sica de Canarias.  References Bahcall, J.N., and Spitzer, L. Jr.   , ApJ, , L. Bothun, G.D., Schombert, J.M., Impey, C.D., Sprayberry, D., and McGaugh, S.S.  , AJ, , . Bottinelli, L., Gouguenheim, L., Fouqu(cid:19)e, P., Paturel, G.  , A&AS, ,  . Briggs, F.H.  , AJ, , . Burstein, D., and Heiles, C.  , ApJS, , . Carilli, C.L., and van Gorkom, J.H.  , ApJ,  , . Fisher, J.R., and Tully, R.B.  , ApJ, , L. Gallagher, J.S. and Hunter, D.A.   , AJ , . Haynes, M.P., and Giovanelli, R.  , AJ,  , . Haynes, M.P., and Giovanelli, R.   , in Pont. Acad. Sci. Study Week , Large-Scale Motions in the Universe, ed. V. Rubin and G. Coyne (Rome: Specola Vaticana),  (HG). Heisler, J., Tremaine, S., and Bahcall, J.N.  , ApJ,  , . Henning, P.A.  , ApJS, , . Ho(cid:11)man, G.L, Helou, G., Salpeter, E.E., and Williams, H.L.   , ApJS,  , . Ho(cid:11)man, G.L., Lu, N.Y., and Salpeter, E.E.  , AJ, , . Huchra, J.P., and Geller, M.J.  , ApJ, , . Hunter, D.A., and Gallagher, J.S.  , PASP, , . Kennicutt, R.C., and Kent, S.M.  , AJ, ,  . Kunth, D.  , in Evolutionary Phenomena in Galaxies, ed. J.E. Beckman and B.E.J. Pagel (Cambridge: Cambridge University Press), . Landolt, A.U.  , AJ, ,  . Lanzetta, K.M., Wolfe, A.M., Turnshek, D.A., Lu, L., McMahon, R.G., and Hazard, C.  , ApJS, , . Lanzetta, K.M.  , in The Evolution of Galaxies and Their Environment, Proceedings of the Third Tetons Summer Astrophysics Conference, ed. J.M. Shull and H.A. Thronson, Jr. (Dordrecht: Kluwer Academic Publishers), . Mazzei, P., Xu, C., and De Zotti, G.  , A&A, , . Napier, P.J., Thompson, A.R., and Ekers, R.D.  , Proc. Inst. Electr. Eng., ,  . Neugebauer, G. et al.  , ApJ, , L. Nilson, P.  , Uppsala General Catalogue of Galaxies, Uppsala: Uppsala Obs. Ann. (UGC). Pierce, M.J., and Tully, R.B.  , ApJ, ,  (PT). Rao, S., and Briggs, F.H.  , ApJ,  , in press. Rieke, G.H., and Lebofsky, M.J.  , ApJ, , . Salpeter, E.E.  , ApJ, , . Tyson, N.D.  , ApJ,  , L. Tyson, N.D., and Scalo, J.M.  , ApJ,  , . Tully, B.  , Nearby Galaxies Catalog, Cambridge University Press. Tully, R.B., and Fisher, J.R.  , A&A, , . Tully, B., and Fouqu(cid:19)e, P.  , ApJS, , .  de Vaucouleurs, G., de Vaucouleurs, A., Corwin, H.G. Jr., Buta, R.J., Paturel, G., and Fouqu(cid:19)e, P.  , Third Reference Catalogue of Bright Galaxies, New York: Springer- Verlag (RC). Weinberg, D.H., Szomoru, A., Guhathakurta, P., and van Gorkom, J.H.  , ApJ, , L (Paper I). Wolfe, A.M.  , in The Interstellar Medium in Galaxies, Proceedings of the Second Tetons Summer Astrophysics Conference, ed. H.A. Thronson, Jr. and J.M. Shull (Dordrecht: Kluwer Academic Publishers), . York, D.G., Dopita, M., Green, R., and Bechtold, J.  , ApJ, , . Zwicky, F., Herzog, E., Wild, P., Karpowicz, M., and Kowal, C.T.  { , Catalog of Galaxies and Clusters of Galaxies, Pasadena: California Inst. of Technology.  Figure captions Figure . Global Hi pro(cid:12)les of all detected galaxies. The (cid:13)ux densities have been corrected for primary beam attenuation. All galaxies detected in a given target (cid:12)eld are plotted together in the same panel ((cid:12)elds { appear in order starting from top left), as are the three Hi-dwarfs detected in blank (cid:12)elds (bottom right panel). The velocity resolution is about  km s . The uncertainties in the (cid:13)ux densities per channel range (cid:0) from . to . mJy for the Hi-dwarfs, depending on the distance to the (cid:12)eld center, and from . to . mJy for the bright galaxies. Figure . Intensity-weighted mean velocity (cid:12)elds of selected galaxies, those in which the Hi is well resolved, both spatially and kinematically. These include the target galaxies of (cid:12)elds , , , and  (UGC  , the target galaxy of (cid:12)eld , was not detected in Hi), along with Zw . ((cid:12)eld ) and UGC , the second bright galaxy in (cid:12)eld . The spacing between adjacent iso-velocity contours is  km s , except in the cases of UGC  and (cid:0) UGC  which have a spacing of  km s , and the velocities of a few contours are (cid:0) marked in km s . The FWHM of the synthesized VLA beam is indicated by the shaded (cid:0) ellipse in the lower left corner of each panel. In top right panel, the Hi-dwarf UGC a can be seen as a close companion,  south of the target galaxy UGC . The velocity (cid:12)eld shows two distinct rotating disks whose pro jected rotation axes are perpendicular. Figure . Spatial distribution of the total Hi emission in (cid:12)eld , centered on the target galaxy UGC . The map has not been corrected for primary beam attenuation. The FWHM of the beam is indicated by the shaded ellipse. The contour levels are  (equal to (cid:27) in the outer parts of UGC ), , , , , and  mJy beam km s . (cid:0) (cid:0) Figure . Same as Figure  for (cid:12)eld , centered on UGC . Note that the Hi distributions of UGC  and the Hi-dwarf UGC a overlap. The contour levels are  (equal to (cid:27) in the outer part of UGC ), , , , , , , , , and  mJy beam km s . (cid:0) (cid:0) Figure . Same as Figure  for (cid:12)eld , centered on the target galaxy UGC  . No Hi is detected in the target galaxy of this (cid:12)eld, UGC  ; its position is marked by a cross. The contour levels are  (equal to :(cid:27) in the outer parts of UGC  a), , , , and  mJy beam km s . (cid:0) (cid:0) Figure . Same as Figure  for (cid:12)eld , centered on the target galaxy UGC . The contour levels are  (equal to :(cid:27) in the outer parts of UGC ), , , , , , , , and  mJy beam km s . (cid:0) (cid:0) Figure . Same as Figure  for (cid:12)eld , centered on the target galaxy UGC . The contour levels are  (equal to :(cid:27) in the outer parts of UGC ),  , , , ,   , , and  mJy beam km s . (cid:0) (cid:0) Figure . Contour plots of the total Hi distribution of the galaxies in (cid:12)eld  superimposed on B -band (negative) gray-scale images. The Hi contours have been corrected for primary beam attenuation. The beam size (FWHM) is indicated by the shaded ellipse in each panel. The lowest gray level corresponds to B (cid:24)  mag arcsec . The Hi contour levels (cid:0) are: () Zw  . : . (:(cid:27) ), ., ., ., and : (cid:2)  cm ;  (cid:0) () UGC a: . (:(cid:27) ), ., ., ., ., and : (cid:2)  cm ;  (cid:0) () UGC : . (:(cid:27) ), ., ., ., , ., . and . (cid:2) cm ; and  (cid:0) () UGC b: . ((cid:27) ), ., ., ., . , ., and : (cid:2)  cm .  (cid:0) Figure . Same as Figure  for UGC b, an Hi-dwarf detected in (cid:12)eld . The Hi column density contours are . (:(cid:27) ), , .,  and : (cid:2)  cm .  (cid:0) Figure . Same as Figure  for UGC  a, an Hi-dwarf detected in (cid:12)eld . The Hi column density contours are . (:(cid:27) ), ., ., and : (cid:2)  cm .  (cid:0) Figure . Same as Figure  for UGC  b, an Hi-dwarf detected in (cid:12)eld . The Hi column density contours are . (:(cid:27) ), , ., , ., and (cid:2)  cm .  (cid:0) Figure . Same as Figure  for Zw ., a previously cataloged galaxy that was detected in Hi in (cid:12)eld . The Hi column density contours are . (:(cid:27) ), ., ., ., , ., and : (cid:2)  cm .  (cid:0) Figure . Same as Figure  for PPHI +, an Hi-dwarf detected in blank (cid:12)eld . The Hi column density contours are . (:(cid:27) ), ., ., ., ., ., and : (cid:2) cm .  (cid:0) Figure . Same as Figure  for PPHI +, an Hi-dwarf detected in blank (cid:12)eld . The Hi column density contours are . (:(cid:27) ), ., ., ., ., ., ., and : (cid:2)  cm .  (cid:0) Figure . Same as Figure  for PPHI +, an Hi-dwarf detected in blank (cid:12)eld . The Hi column density contours are . (:(cid:27) ), ., ., ., and : (cid:2)  cm .  (cid:0) Figure . The B -magnitude-vs-linewidth (Tully-Fisher) relation. The dashed line shows the relation derived for optically selected galaxies by Pierce and Tully ( ), the dotted line an unweighted least-squares (cid:12)t to our data. Triangles represent the previously cataloged galaxies in our Hi sample, squares the Hi-dwarfs. Absolute magnitudes are computed assuming H =  km s Mpc , for consistency with Pierce and Tully. Our (cid:0) (cid:0) Hi-dwarfs lie close to the relation derived for optically selected galaxies, indicating that  they are not anomalously faint for their circular velocities. Figure . Two-color diagram for galaxies in our sample. Squares represent Hi- dwarfs; triangles represent the three previously cataloged galaxies for which we have BRI photometry. The open circles and connecting line show the evolution model of Mazzei et al. ( ). Figure . Extrapolated central B -band surface brightness (upper panel) and logarithm of total blue luminosity (lower panel) plotted against logarithm of the pro jected angular distance to the nearest bright galaxy. Triangles represent previously cataloged galaxies (excluding (cid:12)eld target galaxies); we have a blue luminosity but not a central surface brightness for UGC . Squares represent Hi-dwarfs, which we have divided into bound ((cid:12)lled symbols) and unbound (open symbols) systems as described in x.. There is weak evidence for an environmental e(cid:11)ectin particular, the three low surface brightness systems are all unbound.  Table : Parameters of VLA observations. Con(cid:12)guration D Date of observations November   Number of antennae  Observing mode IF Total bandwidth . MHz Number of channels  Shortest spacing . km Longest spacing . km FWHM of synthesized beam (cid:24)  (cid:2)  FWHM of primary beam  Equivalent T for  mJy/beam . K b Region Void Supercluster Central heliocentric velocity  km s  km s (cid:0) (cid:0) Velocity resolution . km s . km s (cid:0) (cid:0) Velocity coverage . km s . km s (cid:0) (cid:0) Number of (cid:12)elds   Integration time per (cid:12)eld  min  min R.m.s. noise  mJy/beam . mJy/beam  Table : Pointings of VLA observations. () () () (cid:11) (cid:14) Region     (h m s) ( ) o .  V+SC  .   V+SC  .    V+SC  .  V+SC  .  V  .  V  .  V+SC  .  V  .  V  .  V+SC  .  V  .  V  .  V  .    SC  .  V  .  V  .  V  .    V+SC   .  V   .  V   .  V   .    V+SC   .  V+SC   .  V   .  V   .    V   .  V+SC   .  V   .  V   .  V+SC   .    V Note to table : Col. : V denotes the void (cid:12)elds, SC the supercluster (cid:12)elds.    Table . HI and radio continuum parameters () () () () () () () () ( ) () () () Field Name Type (cid:11) (cid:14) V D W W log W M F       hel HI cont R i (h m s) ( ) (km/s) (Mpc) (km/s) (km/s) ( M ) (mJy) (cid:12) o   Zw  . Sc .      (cid:6)  .   (cid:6)   (cid:6)  . : (cid:6) : UGC a  .      (cid:6)  .  (cid:6)   (cid:6)  . : (cid:6) : UGC  SBb  .      (cid:6)  .   (cid:6)   (cid:6)  . : (cid:6) : : (cid:6) : (cid:3) UGC b  .     (cid:6)  .  (cid:6)   (cid:6)  . : (cid:6) :  UGC  Sb  .     (cid:6)  .  (cid:6)   (cid:6)  . : (cid:6) : : (cid:6) : (cid:3) UGC a  .     (cid:6)  .  (cid:6)   (cid:6)  : (cid:6) : UGC b  .     (cid:6)  .  (cid:6)   (cid:6)  . : (cid:6) :  UGC  a  .     (cid:6)  .   (cid:6)   (cid:6)  .  : (cid:6) : UGC   Sa  .     . (cid:3) UGC  b  .     (cid:6)  .  (cid:6)   (cid:6)  . : (cid:6) :  UGC  S  .     (cid:6)  .  (cid:6)   (cid:6)  : (cid:6) : (cid:3) Zw .  .     (cid:6)  .  (cid:6)   (cid:6)  . : (cid:6) :  UGC  Sc   .    (cid:6)  .  (cid:6)   (cid:6)  . : (cid:6) : : (cid:6) : UGC  SBa-b   .     (cid:6)  .  (cid:6)   (cid:6)  . : (cid:6) : : (cid:6) : (cid:3)  PPHI +  .     (cid:6)  .  (cid:6)   (cid:6)  . : (cid:6) :  PPHI +  .      (cid:6)  .  (cid:6)   (cid:6)  . : (cid:6) :  PPHI +   .    (cid:6)  .  (cid:6)   (cid:6)  .  : (cid:6) : Col. : Target galaxies are denoted by asterisks. UGC  , the target galaxy of (cid:12)eld , was not detected in Hi: all data related to this Notes to table : galaxy and all optical data related to the cataloged galaxies which were not reobserved by us were taken from the Third Reference Catalogue of Bright Galaxies (RC; de Vaucouleurs et al.  ). Col. : Heliocentric velocity, computed by averaging the mean velocities at % and % of the peak emission. The uncertainties are derived from the uncertainties in the  cm (cid:13)uxes per channel, but estimated to be no less than  km s . (cid:0) Col. : Adopted distance, calculated from the heliocentric velocity corrected for solar motion with respect to the velocity centroid of the Local Group. Col.  and :  cm line widths at % and % of the peak emission, not corrected for instrumental broadening. The uncertainties are derived from the uncertainties in the  cm (cid:13)uxes per channel, but assumed to be no less than  km s . (cid:0) Col. : Logarithm of the  cm line width at % of the peak emission, corrected for instrumental broadening according to Bottinelli et al. ( ), for internal motions according to Tully & Fouqu(cid:19)e ( ) and corrected to edge-on. We do not list this parameter for UGC  as it is face-on. Col. : Total Hi mass. Errors are derived from the (cid:27) errors in the  cm (cid:13)ux per channel and the uncertainties in velocity width and distance. Col. : Integrated  cm continuum (cid:13)ux at the position of the Hi emission. The errors are the (cid:27) errors in the continuum maps. Table . Optical parameters () () () () () () () () ( ) () () Field Name B B (cid:0) R R (cid:0) I M (B (cid:0) R) (R (cid:0) I ) i P:A: L T B B b;i b;i b;i (mag) (mag) (mag) (mag) (mag) (mag) ( ) ( ) ( L ) B;(cid:12) o o   Zw  . : (cid:6) : : (cid:6) : : (cid:6) : (cid:0): . . : (cid:6) :  (cid:6)  : (cid:6) : UGC a : (cid:6) : : (cid:6) : : (cid:6) : (cid:0): . . : (cid:6) :  (cid:6)  : (cid:6) : UGC  : (cid:6) : :  (cid:6) : : (cid:6) : (cid:0) : . . : (cid:6) :  (cid:6)   (cid:6)  (cid:3) UGC b : (cid:6) : :  (cid:6) : : (cid:6) : (cid:0): . . : (cid:6) :  (cid:6)  : (cid:6) :  UGC  : (cid:6) : (cid:0): .   (cid:6)  (cid:3) UGC a UGC b : (cid:6) : : (cid:6) : : (cid:6) : (cid:0): . . : (cid:6) :  (cid:6)  : (cid:6) :  UGC  a : (cid:6) : :  (cid:6) : : (cid:6) : (cid:0): . . : (cid:6) :  (cid:6)  : (cid:6) : UGC   : (cid:6) : (cid:0) : .   (cid:6)  (cid:3) UGC  b : (cid:6) : : (cid:6) : : (cid:6) : (cid:0): . . .  : (cid:6) :    UGC  : (cid:6) : (cid:0):   (cid:6)  (cid:3) Zw . : (cid:6) : : (cid:6) : : (cid:6) : (cid:0): . . : (cid:6) :  (cid:6)  : (cid:6) :  UGC  : (cid:6) : (cid:0) :  .   (cid:6)  UGC  : (cid:6) : (cid:0): .   (cid:6)  (cid:3)  PPHI + : (cid:6) : : (cid:6) : : (cid:6) : (cid:0): . . : (cid:6) :  (cid:6)  : (cid:6) :  PPHI + : (cid:6) : : (cid:6) : : (cid:6) : (cid:0): . .  : (cid:6) :  (cid:6)  : (cid:6) :  PPHI + : (cid:6) : :  (cid:6) : : (cid:6) : (cid:0): . .  : (cid:6) :  (cid:6)  : (cid:6) : Col. : Target galaxies are denoted by asterisks. UGC  , the target galaxy of (cid:12)eld , was not detected in Hi: all data related to this galaxy Notes to table : and all optical data related to the cataloged galaxies which were not reobserved by us were taken from the RC catalog. Col. : Total B -band magnitude. Errors are a combination of the (cid:27) zero-point errors, (cid:13)at-(cid:12)elding errors and the estimated errors due to imperfect subtraction of adjacent and superposed stars. Col.  and : Uncorrected broadband colors. Errors are derived from the errors in the total magnitudes in the individual bands. Col. : Absolute B magnitude corrected for Galactic and internal absorption following Burstein & Heiles ( ) and Tully & Fouqu(cid:19)e ( ). Col.  and : Broadband colors corrected for Galactic and internal absorption following Burstein & Heiles ( ) and Tully & Fouqu(cid:19)e ( ). Col. and : Inclinations and position angles derived by (cid:12)tting ellipses to the I -band images, assuming an intrinsic axis ratio of . and correcting for the  e(cid:11)ect (Tully  ). These quantities are not well determined for Zw  . and UGC  b. The values for UGC , o UGC  , UGC , UGC  and UGC  are taken from the RC catalog. The errors are the r.m.s. variation of the (cid:12)ts. Col. : Total blue luminosity, derived from the blue magnitudes listed in column , assuming an absolute solar blue magnitude of .. Table . H(cid:11) and derived parameters () () () () () () () () ( ) () () () () Field Name (cid:22) (cid:22) h D D M =L S W M log ( M =L ) log ((cid:28) ) ;B  HI HI B H(cid:11) B eq b;i _ _ (cid:0) (cid:1) (cid:0) (cid:1) ;B mag mag  erg (cid:23) (cid:0) arcsec arcsec    cm s ( ) ( ) ( ) (M =L ) ( ) ( A) (M =yr) (M =L =yr) (yr) (cid:12) (cid:12) (cid:12) (cid:12) (cid:12)  Zw  . : (cid:6) : . : (cid:6) : : (cid:6) :  : (cid:6) : UGC a : (cid:6) : . : (cid:6) : : (cid:6) :  : (cid:6) : .  . (cid:0): .  UGC  : (cid:6) : .  (cid:6) : : (cid:6) :  : (cid:6) : (cid:3) UGC b : (cid:6) : . : (cid:6) : : (cid:6) :  : (cid:6) : .  . (cid:0): .  UGC  : (cid:6) :  : (cid:6) : (cid:3) UGC a  UGC b  : (cid:6) : . : (cid:6) :  : (cid:6) :  : (cid:6) : .  . (cid:0): .  UGC  a : (cid:6) : . : (cid:6) : : (cid:6) :  : (cid:6) : .  . (cid:0): . UGC    (cid:6) : (cid:3) UGC  b : (cid:6) : . : (cid:6) :  : (cid:6) :  UGC   : (cid:6) :  : (cid:6) : (cid:3) Zw . : (cid:6) : . : (cid:6) : : (cid:6) :  : (cid:6) :    UGC  : (cid:6) :  : (cid:6) : UGC  : (cid:6) :  : (cid:6) : (cid:3)  PPHI + : (cid:6) : . : (cid:6) : : (cid:6) :  : (cid:6) :  PPHI + : (cid:6) : . : (cid:6) : : (cid:6) :  : (cid:6) :  PPHI + : (cid:6) : . : (cid:6) : : (cid:6) :  : (cid:6) : Col. : Target galaxies are denoted by asterisks. UGC  , the target galaxy of (cid:12)eld , was not detected in Hi: all data related to this galaxy Notes to table : and all optical data related to the cataloged galaxies which were not reobserved by us were taken from the RC catalog. Col. : Extrapolated B -band central surface brightness, determined by an extrapolation of a linear (cid:12)t to the straight part of the radial luminosity pro(cid:12)le plotted on a magnitude scale. The errors are the (cid:27) errors in the y-estimate of the regression. Col. : Extrapolated B -band central surface brightness, corrected for Galactic and internal absorption following Burstein & Heiles ( ) and Tully & Fouqu(cid:19)e ( ) and corrected for line-of-sight integration. Col. : Scale length in seconds of arc, determined by a linear (cid:12)t to the straight part of the radial luminosity pro(cid:12)le plotted on a magnitude scale. At  Mpc,  corresponds to . kpc. The errors are computed from the (cid:27) errors in the estimate of the x-coe(cid:14)cient of the regression. Col. : Diameter at B =  mag arcsec , in seconds of arc. This was determined by a linear (cid:12)t to the straight part of the radial luminosity (cid:0) pro(cid:12)le plotted on a magnitude scale, except for UGC  b, in which case an elliptical (cid:12)t to the B =  mag arcsec isophote was used (cid:0) The errors are computed from the (cid:27) errors in the estimates of y and x-coe(cid:14)cients of the regression. Col. : Hi diameter, obtained by (cid:12)tting gaussians to the Hi distributions and deconvolving these with the beams. Col. : Total Hi mass divided by the blue luminosity. The errors are computed from the (cid:27) errors in Hi mass and B -band luminosity. Col. and : The measured H(cid:11) (cid:13)ux and equivalent width, respectively. Col. : The star formation rate (SFR) derived as described by Hunter & Gallagher ( ). The measured (cid:13)uxes were corrected for Galactic absorption. Col. : The logarithm of the ratio of the SFR and blue luminosity, both corrected for Galactic absorption only. Col. : The logarithm of the timescale for gas depletion (cid:28) (in years). This has been computed by dividing the Hi mass, multiplied by . to account for the presence of He, by the SFR. a Table : Average optical colors. Color All Types(cid:21) Sd All Hi-dwarfs Cataloged Control sample Our sample B (cid:0) R : (cid:6) : : (cid:6) : :  (cid:6) : : (cid:6) : : (cid:6) : R (cid:0) I : (cid:6) : : (cid:6) : : (cid:6) : : (cid:6) : : (cid:6) : Table : Group Mass Estimates () () () () () () () () Field Name (cid:1)V (cid:1)R M M M =L M =L est min est B min B (cid:0)   (km s ) (kpc) ( M ) ( M ) (M =L ) (M =L ) (cid:12) (cid:12) (cid:12) (cid:12) (cid:12) (cid:12)    Zw  . {. . . . UGC a . . . . UGC b {.  . . .    UGC a {. . . . UGC b . . . .     UGC  a . . . . UGC  b .  . . .     Zw . .  .  .  .    UGC  {. . . . Notes to table : Cols.  and : Line-of-sight velocity di(cid:11)erence and pro jected distance relative to the central galaxy of the (cid:12)eld. Col. : Total mass of the group, computed using the average mass estimator of Heisler et al. ( ). Col. : Minimum pair mass needed to bind a companion to the central galaxy of the (cid:12)eld. Col. : Total estimated mass divided by total blue luminosity of the group of galaxies. Col. : Minimum binding mass of each companion-central galaxy pair divided by the total blue luminosity of this pair. 
astro-ph/0012392
1
0012
2000-12-18T16:41:42
Properties of the Chandra Sources in M81
[ "astro-ph" ]
The Chandra X-ray Observatory obtained a 50-ks observation of the central region of M81 using the ACIS-S in imaging mode. The global properties of the 97 x-ray sources detected in the inner 8.3x8.3 arcmin field of M81 are examined. Roughly half the sources are concentrated within the central bulge. The remainder are distributed throughout the disk with the brightest disk sources lying preferentially along spiral arms. The average hardness ratios of both bulge and disk sources are consistent with power law spectra of index Gamma~1.6 indicative of a population of x-ray binaries. A group of much softer sources are also present. The background source-subtracted logN-logS distribution of the disk follows a power law of index ~ -0.5 with no change in slope over three decades in flux. The logN-logS distribution of the bulge follows a similar shape but with a steeper slope above ~4.0e+37 ergs/s. There is unresolved x-ray flux from the bulge with a radial profile similar to that of the bulge sources. This unresolved flux is softer than the average of the bulge sources and extrapolating the bulge logN-logS distribution towards weaker sources can only account for 20% of the unresolved flux. No strong time variability was observed for any source with the exception of one bright, soft source.
astro-ph
astro-ph
Submitted to Astrophysical Journal Letters. Preprint typeset using LATEX style emulateapj v. 04/03/99 0 0 0 2 c e D 8 1 1 v 2 9 3 2 1 0 0 / h p - o r t s a : v i X r a PROPERTIES OF THE Chandra SOURCES IN M81 Allyn F. Tennant1, Kinwah Wu2, Kajal K. Ghosh3, Jeffery J. Kolodziejczak1, and Douglas A. Swartz4 Submitted to Astrophysical Journal Letters. ABSTRACT The Chandra X-ray Observatory obtained a 50-ks observation of the central region of M81 using the ACIS-S in imaging mode. The global properties of the 97 x-ray sources detected in the inner 8′.3 × 8′.3 field of M81 are examined. Roughly half the sources are concentrated within the central bulge. The remainder are distributed throughout the disk with the brightest disk sources lying preferentially along spiral arms. The average hardness ratios of both bulge and disk sources are consistent with power law spectra of index Γ∼1.6 indicative of a population of x-ray binaries. A group of much softer sources are also present. The background source-subtracted logN -logS distribution of the disk follows a power law of index ∼-0.5 with no change in slope over three decades in flux. The logN -logS distribution of the bulge follows a similar shape but with a steeper slope above ∼ 4 × 1037 ergs s−1. There is unresolved x-ray flux from the bulge with a radial profile similar to that of the bulge sources. This unresolved flux is softer than the average of the bulge sources and extrapolating the bulge logN -logS distribution towards weaker sources can only account for 20% of the unresolved flux. No strong time variability was observed for any source with the exception of one bright, soft source. Subject headings: X rays: galaxies -- galaxies: individual (M81) -- galaxies: luminosity function -- galaxies: structure -- galaxies: fundamental parameters 1. INTRODUCTION deferred to a future paper. Among the best-studied galaxies beyond the Local Group is the nearby Sab spiral M81 (NGC 3031). With an inclination of 59◦, a dominant two-armed spiral pattern, and a well-defined central bulge, M81 is ideal for investi- gating formation and evolution of galaxies similar to our own. Most x-ray studies of M81 have concentrated on the dominant source in the field, a low-luminosity AGN at its core. The only extensive examination of the entire M81 field at x-ray energies is that of Fabbiano (1988, hereafter F88) based on Einstein observations. Einstein detected 9 sources in M81 including the nucleus, all with x-ray lu- minosities exceeding ∼ 2 × 1038 ergs s−1, with 5 of these located along spiral arms. We have identified an additional 9 sources within a ∼6′ radius of the nucleus from archival ROSAT observations. In this first examination of a Chandra X-ray Observa- tory observation of M81, we analyze the global properties of the 97 x-ray sources detected in the inner 8′.3×8′.3 field to a limiting 0.2-8 keV luminosity of L ∼ 3 × 1036 ergs- s−1. Details of the observation and the analysis methods are given in § 2. Section 3 examines the spatial distribu- tion of sources and identifies those sources spatially coinci- dent with previously classified objects. The hardness ratio distribution is presented in § 4 and the logN -logS distribu- tion in § 5. Section 6 discusses the luminosity evolution of sources common to the Einstein (c.1980), ROSAT (∼1990) and present Chandra observations. The results are sum- marized in § 7. Detailed analysis of individual sources are 2. OBSERVATIONS A 49926 second observation of a portion of the galaxy M81 was obtained with the Chandra Advanced CCD Imaging Spectrometer (ACIS, G. P. Garmire et al., in preparation) on 2000 May 7. The primary target, SN 1993J, was located near the nominal aimpoint on the back- illuminated CCD S3 of the spectroscopy array (operating in imaging mode). The nucleus of M81 lies 2.′79 from SN 1993J towards the center of S3 in this observation. The observation was taken in faint time exposure mode at 3.241 s-frame−1. Standard Chandra X-ray Center pro- cessing has applied aspect corrections and compensated for spacecraft dither. We selected the standard grade set and events in pulse invariant (PI) channels corresponding to ∼ 0.2 to 8.0 keV for analysis. The majority of sources identified are within the 8′.3 × 8′.3 field of S3. To avoid difficulties with the intrinsic dif- ferences between front- and back-illuminated devices we restrict our analysis to CCD S3. Due to poor cosmic ray rejection for the central two columns of S3, all events with CHIPX equal to 512 or 513 were excluded from analy- sis. Since these columns are dithered across the sky, their removal results in a slight loss of sensitivity (5%) in the central 16′′. Pixels with centers less than 1′′.25 from the readout trail of the nucleus were flagged and excluded from analysis as were pixels containing less than 50% sky expo- sure near the edge of the CCD. Sources were located using a minimum variance estima- tor that compares the data to the telescope Point Spread 1Space Science Department, NASA Marshall Space Flight Center, SD50, Huntsville, AL, USA 2School of Physics, University of Sydney, 2006, Australia; and MSSL, University College London, Holmbury St. Mary, Surrey, RH5 6NT, UK 3NAS/NRC Senior Resident Research Associate, NASA Marshall Space Flight Center, SD50, Huntsville, AL, USA 4Universities Space Research Association, NASA Marshall Space Flight Center, SD50, Huntsville, AL, USA 1 2 Function (PSF). This method is most suited to finding point sources that roughly match the telescope's PSF. The PSF was approximated by a circular gaussian with the same width as the measured PSF at the aimpoint and in- creasing quadratically with distance from the aimpoint to approximately match the core and off-axis broadening of the true PSF. We define the signal-to-noise ratio (S/N) to be the estimated source counts divided by the 1σ uncer- tainity in this number. To set a minimum threshold, we considered the number of sources found per S/N interval down to a minimum S/N of 1.0. This distribution shows a clear gaussian noise peak at low S/N that indicates a sin- gle false detection (on S3) is to be expected for a minimum S/N of 3.0 and less than 0.1 false detections at a threshold of 3.5. We include all sources with S/N>3.0 in our anal- ysis with the exception of the spectral analysis (§ 4) and logN -logS analysis (§ 5) for which only S/N>3.5 sources are considered. A second pass, designed to mitigate source confusion, was made through the data. For this purpose, all pixels within 2σ of each source position were flagged and the source-finding process repeated. In this way, the variance in the local background due to nearby sources is greatly reduced and otherwise confused sources isolated. An ad- ditional 6 sources, 4 with S/N>3.5, were detected. The final source list contained 81 sources with S/N≥3.5 with an additional 16 with 3.0≤S/N<3.5. These source positions are identified in the ACIS data shown in fig- ure 1a and superimposed on the second generation digital sky survery (DSS) image in figure 1b. Small spatial regions centered on each source were then selected for further analysis. The size of these spatial re- gions also varies quadratically with off-axis position and encompasses roughly the 95% encircled energy radii of the PSF at 1.5 keV. No correction for the low background, ∼0.04 cts-pixel−1, is made. For spectral analysis (§ 4), the total counts within each source region were binned into soft (0.2 -- 2.0 keV) and hard (2.0 -- 8.0 keV) bands. Higher res- olution spectra, at 16 PI channel binning out to ∼10 keV, were examined for spurious effects such as anomolously high background. No anomolies were encountered. Finally, we distinguish between the 'bulge' and 'disk' defined here as those regions internal and external, respec- tively, to a 4.′70×2.′35 ellipse centered on the nucleus with major axis position angle 149◦ (cf. figure 1). This ellipse corresponds to a circle in the plane of the galaxy of radius 2.′35 (∼2.5 kpc). There is extended, unresolved x-ray flux in the bulge that is above background and is not due to the PSF wings of the nucleus nor to other point sources. This component will be referred to as 'excess bulge emission.' 3. MORPHOLOGY The bulge contains 41 sources at a surface density of 5.4 sources-arcmin−2 and the disk contains 56 sources (on S3) at a density of ∼1.0 source-arcmin−2. Among the latter, 21 are located within ±400 pc of spiral arms (as defined by Matonick & Fesen (1997) by tracing the Hα emission peak, their figure 31b). The surface density of spiral arm sources in our sample is slightly less than, but statisti- cally consistent with, the surface density of disk sources not associated with spiral arms. The surface density distribution of bulge sources, inclination-corrected and centered on the nucleus, can be described by an exponential radial distribution, σ(r) ∝ e−r/ro, with ro ∼ 0.92 ± 0.26 kpc. After subtracting the background component and the contribution from the wings of the PSF of the nucleus, the surface brightness profile of the excess bulge emission can also be described by an exponential with scale length ro ∼ 0.83±0.20 kpc for 0.′8 ≤ r ≤ 5.′5. The excess bulge emission is much steeper at smaller radii though a residual contribution from the wings of the nuclear PSF cannot be definitively excluded. Beyond r ∼ 1 kpc, these profiles are roughly consistent with the V (Georgiev & Getov 1991), g (Frei et al. 1996), and B and I (Elmegreen & Elmegreen 1984) profiles sug- gesting the bulge x-ray emission follows the distribution of starlight. A search of the literature for known objects spatially coincident with Chandra sources found no matches within the bulge, with the exception of the nucleus. Among the disk sources, 5 are located within 3′′ of supernova rem- nants (SNR; Matonick & Fesen 1997) including SN 1993J and the bright Einstein source X-6 (F88). All 5 SNR can- didates lie on spiral arms as expected for young, massive star, Type II supernova progenitors. Source X-6 is a point source located within a 90-pc-diameter SNR (Matonick & Fesen 1997). The high luminosity and hard spectrum (§ 4) of X-6 is uncharacteristic of SNRs. Six x-ray sources, all located within spiral arms, are within 10′′ of known HII regions, half of which are giant radio HII regions (Kaufman et al. 1987). One of these x- ray sources, though weak, has the hardest spectrum of the entire source sample (§ 4). No x-ray sources were found to coincide with known globular clusters, novae, or stars. Roughly 10% of globu- lar clusters are expected to contain accreting neutron stars detectable above our threshold and galaxies like M81 typ- ically contain of order 200 globular clusters (Perelmuter & Racine 1995). However, there are only 25 confirmed glob- ular clusters in M81 (Perelmuter, Brodie & Huchra, 1995) and only 4 within the S3 field. The nova rate in M81 is probably of order 20 year−1 (A. W. Shafter, private communication). The x-ray sig- natures and light curves of novae are only sparsely doc- umented ( Ogelman, Krautte & Beuermann 1987) making novae difficult to identify in our source sample. There are two sources, Einstein source X-2 and a ROSAT-detected object, which are near stars or star clusters. X-2 is 1.′′8 distant from a star cluster with a spectrum consistent with G-type stars (Sholukhova et al. 1998). Zickgraf & Humphreys (1991) list the opti- cal object as a probable foreground F-type dwarf. The ROSAT source is coincident with an undocumented ∼18- magnitude object visible in the Digital Sky Survey image. Interestingly, both x-ray sources are transient and have been much brighter in the past (§ 6). No known background or foreground objects were found coincident with the S3 Chandra sources. Up to ∼25 background objects are expected in the S3 field above S/N = 3.5 based upon our analysis of the calibration field CRSS J0030.5+2618 (see also Brandt et al. 2000). This estimate ignores any obscurtion by M81. Based on results of the Einstein stellar survey (Topka et al. 1982), we es- timate roughly 0.4 F- and G-stars could be detectable in our field but these should all be extremely soft (Maggio et al. 1987, Hodgkin & Pye 1994, Schmitt 1997). A soft spectrum was observed from 3 catalogued stars detected on CCD S2. 4. HARDNESS RATIO The total source counts in the hard x-ray band (2.0 -- 8.0 keV) is shown in figure 2 against the soft (0.2 -- 2.0 keV) band for the subset of sources with S/N≥3.5 excluding the nucleus. There is no adjustment for pileup which tends to harden spectra of the brightest sources. Also shown are the hardness ratios of the sum of all the bulge sources, and of all the disk sources, the hardness ra- tio of the excess bulge emission (background and nuclear PSF subtracted), and of the typical background (scaled to a 2′ × 2′ area). The nucleus and the bright soft source are omitted from the summed bulge point, and X-6 is omitted from the summed disk point as these sources are clearly spectroscopically atypical. The line shown corresponds to an absorbed power law of spectral index Γ = 1.6, typical of x-ray binaries, with the Galactic absorbing column density of NH = 4 × 1020 cm−2 (Stark et al. 1992). Though indicative and useful to guide the eye, this canonical spectral shape is not a model fit result. Most sources fall near this curve. Notable ex- ceptions are the soft sources at the lower right of figure 2. The softest bulge sources have no known associations but the 3 softest disk sources all lie on spiral arms. The softest source is also time-variable (§ 6) and is the third brightest source in the entire sample. The hardness ratios for the summed bulge and summed disk spectra are also typical of x-ray binaries. The excess bulge emission is softer in- dicating some of the excess emission is from truly diffuse, hot, interstellar gas and not entirely from unresolved point sources. 5. LOGN -LOGS The number of sources above a limiting brightness, N (>S), is shown against the 0.2 -- 8.0 keV count rate, S, in figure 3. The nucleus is excluded as are sources with S/N<3.5. Background sources were taken from the cali- bration field CRSS J0030.5+2618 (§ 3), by applying our same analysis methods, and have been subtracted after scaling to the fractional areas of the bulge and disk. The curve shown in figure 3 represents the best power law fit to the disk distribution, N = 0.43S−0.50, over the entire range of S. Similarly, N = 0.44S−0.57 for the bulge source distribution for S<0.004 count-s−1. There is a break in the slope of the distribution of bulge sources above this point but no such break in the disk source dis- tribution. Of the 10 disk sources with S>0.004 count-s−1, 7 are coincident with spiral arms. The luminosity of each source can be estimated by as- suming the Γ = 1.6 power law spectral model of § 4 and a distance of 3.6 Mpc to M81 (Freedman et al. 1994). For our 50-ks observation, therefore, S = 0.004 count- s−1 in the 0.2 -- 8.0 keV band corresponds to a luminosity LX = 3.7 × 1037 ergs s−1 and the faintest source in the field corresponds to a luminosity of ∼ 3 × 1036 ergs s−1. Excluding the nucleus, the total bulge luminosity is LX ∼ 2.4 × 1039 erg s−1 of which 36% is excess (unre- solved) emission. The total luminosity of the disk sources is LX ∼ 3.9 × 1039 erg s−1. The luminosity of the nucleus 3 is LX ∼ 4 × 1040 erg s−1 based on spectral fits to the trailed image. This is within the ASCA-observed range of luminosities (Ishisaki et al. 1996) and comparable to the BeppoSAX observed luminosity (Pellegrini et al. 2000). The nucleus contributes approximately 86% of the total luminosity in the 8′.3 × 8′.3 S3 field of view. 6. TIME EVOLUTION A simple test for time variability was made by binning the light curves of all sources on 1000, 2000, and 4000 second intervals. Applying a χ2 test for the hypothesis of consistency with a mean value found only one clearly significant deviation on all three timescales. This source is also the softest source on S3 and the third brightest. It is present in at least 3 of 6 ROSAT HRI observations spanning 1993 - 1998 but is too close to the nucleus to be identified in any other previous x-ray observation. M81 has been observed in x-rays at moderate spatial res- olution by Einstein in 1979 and by ROSAT over the period 1991 -- 1998. There are 6 Einstein sources (F88) within the S3 field of view and another 4 ROSAT-detected sources according to our analysis. Chandra resolves six ROSAT source regions into two or more sources including Einstein sources X-7 and X-10, and, of course, the nucleus. Long-term variability has been detected in the Chandra variable source, the nucleus, X-2 and the ROSAT-detected source coincident with an undocumented star-like object (§ 3). Both of the latter sources are moderately weak in the present observations (∼3 and ∼ 6 × 1037 ergs-s−1, re- spectively) but have been much brighter in the past. The location of another Einstein source, X-12, places it on the eastern edge of S3. There are no Chandra sources within the portion of the 45′′ error circle of this IPC source that falls on S3 and it is not detectable in the ROSAT ob- servations. 7. SUMMARY The global properties of the Chandra sources identified in the central 8′.3 × 8′.3 region of M81 have been exam- ined. There is a high density of sources located in the bulge of the galaxy. These sources, and the excess bulge x-ray emission, follow an exponential radial distribution about the nucleus with an ∼1 kpc scale length. This is similar to that of optical light and suggests the excess bulge emis- sion is from unresolved point sources and that the point sources trace the old stellar bulge population. However, extrapolating the bulge logN -logS distribution towards weaker sources predicts ∼0.022 source counts-s−1 from sources <0.001 counts-s−1-source−1. Thus, weak sources can only account for 20% of the total excess bulge emis- sion (∼0.092 counts-s−1). There remains a significant flux in the bulge unaccounted for by unresolved sources, par- ticularly at lower energies, indicative of a truly diffuse, hot interstellar gas. If all the excess bulge emission arises from a diffuse thermal bremsstrahlung at kT ∼ 0.2 keV, uniformly distributed throughout the bulge, then there is ∼ 4 × 106 M⊙ of hot gas. The average hardness ratio of the Chandra sources can be described by a power law of index Γ = 1.6 with a Galac- tic absorbing column density of NH = 4 × 1020 cm−2. This is typical of x-ray binaries. Assuming thie canoni- cal spectrum applies to all sources, the background source 4 subtracted logN -logS distribution of the disk sources fol- lows an N ∝ S−0.50 profile extending to the most lumi- nous sources at LX > 1039 ergs-s−1. The distribution of bulge sources, however, shows a break at LX ∼ 4 × 1037 ergs-s−1. Thus the bulge sources are predominantly low- mass x-ray binaries, consistent with an old bulge stellar population, while the disk has an additional population of much brighter sources. Seven of the 10 brightest Chandra disk sources lie within spiral arms, consistent with previ- ous Einstein observations (F88), but the entire sample of disk sources, down to the limiting Chandra luminosity of ∼ 3 × 1036 ergs-s−1, are distributed throughout the disk without preference to spiral arms. Some of these brighter sources may be black hole candidates with high-mass com- panions. We thank Martin Weisskopf for deriving the expression for the uncertainty in the source flux used in the source- finding algorithm. The digital sky survey image used in Figure 1b is from the "Palomar Observatory - Space Tele- scope Science Institute Digital Sky Survey" of the northern sky, based on scans of the Second Palomar Sky Survey, and was produced under NASA Contract NAS5-2555. REFERENCES Brandt, W. N., et al. 2000, AJ, 119, 2349 Elmegreen, D. M., & Elmegreen, B. G. 1984, ApJS, 54, 127 Fabbiano, G. 1988, ApJ 325, 544 Freedman, W. L., et al. 1994, ApJ, 427, 628 Frei, Z., Guhathakurta, P., Gunn, J. E., & Tyson, J. A. 1996, AJ, 111, 174 Georgiev, Ts. B., & Getov, R. G. 1991, SvAL, 17, 168 Hodgkin, S. T., & Pye, J. P. 1994, MNRAS, 267, 840 Ishisaki, Y., et al. 1996, PASJ, 48, 237 Kaufman, M., Bash, F. N., Kennicutt, R. C., Jr., & Hodge, P. W. 1987, ApJ, 319, 61 Maggio, A., et al. 1987, ApJ, 315, 687 Matonick, D. M., & Fesen, R. A. 1997, ApJS, 112, 49 Ogelman, H., Krautter, J., & Beuermann, K. 1987, A&A 177, 110 Pellegrini, S., Cappi, M., Bassani, L., Malaguti, G., Palumbo, G. G. C., & Persic, M. 2000, A&A, 353, 447 Perelmuter, J-. M., Brodie, J. P., & Huchra, J. P. 1995, AJ, 110, 620 Perelmuter, J-. M., & Racine, R. 1995, AJ, 109, 1055 Schmitt, J. H. M. M. 1997, A&A, 318, 215 Sholukhova, O. N., Fabrika, S. N., Vlasyuk, V. V., & Dodonov, S. N. 1998, AZh, 24, 591 Stark, A. A., Gammie, C. F., Wilson, R. W., Bally, J., Linke, R. A., Heiles, C., & Hurwitz, M. 1992, ApJS, 79, 77 Topka, K., Avni, Y., Golub, L., Gorenstein, P., Harnden, F. R., Rosner, R., & Vaiana, G. S. 1982, ApJ, 259, 677 Zickgraf, F-. J., & Humphreys, R. M. 1991, AJ, 102, 113 5 FIGURE CAPTION Fig. 1. -- Left: Chandra ACIS-S image of the inner region of M81 in J2000 coordinates. North is up and east is to the right. Each sky pixel is 0.492′′. The aimpoint is near SN 1993J. The readout trail, caused by photons from the bright nucleus hitting the detector during readout, has been removed. Chandra sources are marked as light blue (S/N>3.5) or magenta (3<S/N<3.5) circles. Einstein sources (Fabbiano 1988) are marked by red circles encompassing the approximate Einstein positional uncertainties. The 4′.7 × 2′.35 ellipse denotes the boundary between 'bulge' and 'disk' as used in the text. Right: B-band digital sky survey image of M81 with the locations of sources identified as in left panel. Fig. 2. -- Hardness ratio of the 80 sources with S/N>3.5, excluding the nucleus, with √N errors. Also included are the sum of all bulge sources (B) excluding the bright soft source and the nucleus; the sum of all disk sources (D) excluding Einstein source X-6; the excess (unresolved) bulge emission (⋄); and the average instrument background (∗) over a 2′ × 2′ region. The line denotes the hardness ratio of a typical x-ray binary; a Galactically-absorbed power law of index Γ = 1.6. Fig. 3. -- Background-source-subtracted logN -logS for bulge and disk sources. The background logN -logS distribution was taken from Chandra calibration observation of CRSS J0030.5+2618 (see also Brandt et al. 2000) using the same analysis procedures used in this study of M81. The line represents the power law fit to the disk distribution, N = 0.43S−0.50. This figure "fig01.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/astro-ph/0012392v1
astro-ph/9911212
1
9911
1999-11-11T18:46:27
Accretion Discs Around Black Holes: Developement of Theory
[ "astro-ph" ]
Standard accretion disk theory is formulated which is based on the local heat balance. The energy produced by a turbulent viscous heating is supposed to be emitted to the sides of the disc. Sources of turbulence in the accretion disc are connected with nonlinear hydrodynamic instability, convection, and magnetic field. In standard theory there are two branches of solution, optically thick, and optically thin. Advection in accretion disks is described by the differential equations what makes the theory nonlocal. Low-luminous optically thin accretion disc model with advection at some suggestions may become advectively dominated, carrying almost all the energy inside the black hole. The proper account of magnetic filed in the process of accretion limits the energy advected into a black hole, efficiency of accretion should exceed $\sim 1/4$ of the standard accretion disk model efficiency.
astro-ph
astro-ph
ACCRETION DISCS AROUND BLACK HOLES: DEVELOPEMENT OF THEORY G.S. Bisnovatyi-Kogan1 1 Space Research Institute, Russian Academy of Sciences Profsoyuznaya 84/32, Moscow 117810 Russia, [email protected] ABSTRACT. Standard accretion disk theory is for- mulated which is based on the local heat balance. The energy produced by a turbulent viscous heating is supposed to be emitted to the sides of the disc. Sources of turbulence in the accretion disc are con- nected with nonlinear hydrodynamic instability, con- vection, and magnetic field. In standard theory there are two branches of solution, optically thick, and opti- cally thin. Advection in accretion disks is described by the differential equations what makes the theory nonlo- cal. Low-luminous optically thin accretion disc model with advection at some suggestions may become advec- tively dominated, carrying almost all the energy inside the black hole. The proper account of magnetic filed in the process of accretion limits the energy advected into a black hole, efficiency of accretion should exceed ∼ 1/4 of the standard accretion disk model efficiency. Key words: Stars: accretion discs; black holes. 1. Introduction Accretion is a main source of energy in binary X-ray sources, quasars and active galactic nuclei (AGN). The intensive development of accretion theory began after discovery of X-ray sources and quasars. Accretion into stars is ended by a collision with an inner boundary, which may be a stellar surface, or outer boundary of a magnetosphere for strongly magnetized stars. All grav- itational energy of the falling matter is transformed into heat and radiated outward. In black holes matter is falling to the horizon, from where no radiation arrives. All luminosity is formed on the way to it. The efficiency of accretion is not known from the beginning, and depends on angular momen- tum of the falling matter, and magnetic field embedded into it. It was first shown by Schwartsman (1971) that during spherical accretion of nonmagnetized gas the efficiency may be as small as 10−8 for sufficiently low mass fluxes. He had shown that presence of magnetic field in the accreting matter increase the efficiency up to about 10%, and account of heating of matter due to magnetic field annihilation rises the efficiency up to about 30% (Bisnovatyi-Kogan and Ruzmaikin, 1974) In the case of a thin disc accretion, when matter has large angular momentum, the efficiency is about 1/2 of the efficiency of accretion into a star with a radius equal to the radius of the last stable orbit. In the case of geometrically thick and optically thin accretion discs the situation is approaching the case of spherical symmetry, where presence of a magnetic field plays a critical role. Advection dominated accretion flow (ADAF) was suggested by Narayan and Yu (1995), and used as a solution for some astrophysical problems. The sugges- tions underlying ADAF: ignorance of the magnetic field annihilation in heating of a plasma flow, electron heat- ing only due to binary collisions with protons (ions) had been critically analyzed in papers of Bisnovatyi- Kogan and Lovelace (1997, 1999), Bisnovatyi-Kogan (1999), Quataert (1997). There are contradictions be- tween ADAF model and observational data in radioe- mission of elliptical galaxies (Di Matteo et al., 1999), and X-ray emission of Seyfert galaxy NGC4258 (Can- nizzo, 1998). Account of processes connected with a small-scale magnetic field in accretion flow, strongly restricts solution. Namely, the efficiency of the accre- tion flow cannot become less then about 1/4 of the standard accretion disc value. 2. Development of the standard model Matter falling into a black hole forms a disc when its angular momentum is sufficiently high. It happens when matter comes from the star companion in the binary, or from a tidal disruption of the star which tra- jectory approaches close to the black hole. The first situation is observed in many galactic X-ray sources (Cherepashchuk, 1996). A tidal disruption happens in quasars and active galactic nuclei (AGN) in the model of supermassive black hole surrounded by a dense stel- lar cluster. Equations of a standard accretion disk theory had been first formulated by (Shakura, 1972); some cor- rections and generalization to general relativity (GR) 1 2 Odessa Astronomical Publications, vol. 12 (1999) had been done by Novikov and Thorne (1973). Ob- servational aspects of accretion disks have been ana- lyzed by Shakura and Sunyaev (1973). Note, that all authors of the accretion disc theory from USSR were students (N.I.Shakura) or collaborators (I.D.Novikov and R.A.Sunyaev) of academician Ya.B.Zeldovich, who was not among the authors, but whose influence on them hardly could be overestimated. The main idea of this theory is to describe a geometrically thin non-self- gravitating disc of a mass Md, much smaller then the mass of a black hole M , by hydrodynamic equations averaged over the disc thickness 2h. 2.1. Equations locity equals to the keplerian one Ω = ΩK = (cid:0) GM The small thickness of the disc in comparison with its radius h ≪ r means small importance of the pres- sure gradient ∇P in comparison with gravity and in- ertia forces. Radial equilibrium equation in a disc is a balance between the last two forces with an angular ve- . Note, that just before a last stable orbit around a black hole this suggestion fails, but in the "standard" accre- tion disc model this relation is supposed to hold all over the disc, with an inner boundary at the last sta- ble orbit. The equilibrium equation in the vertical z- direction is determined by a balance between the gravi- tational force and pressure gradient dP . For a thin disc this differential equation is substituted by an algebraic one, determining the half-thickness of the disc in the form r3 (cid:1)1/2 dz = −ρ GMz r3 1 ΩK (cid:18)2 P ρ (cid:19)1/2 . h ≈ (1) The balance of angular momentum, related to the φ component of the Euler equation has an integral in a stationary case, written as M (j − jin) = −2πr2 2htrφ, trφ = ηr dΩ dr . (2) Here j = vφr = Ωr2 is a specific angular momen- tum, trφ is a component of the viscous stress tensor, M > 0 is a mass flux per unit time into a black hole, jin is equal to the specific angular momentum of mat- ter falling into a black hole. In the standard theory the value of jin is determined from physical considerations. For accretion into a black hole it is suggested, that on the last stable orbit the gradient of the angular veloc- ity is zero, corresponding to zero viscous momentum flux. In that case jin = ΩKr2 in, corresponding to the Keplerian angular momentum of the matter on the last stable orbit. at low microscopic (atomic or plasma) viscosity the sta- tionary accretion disc must be very massive and thick, and before its formation the matter is collected by disc leading to a small flux inside. It contradicts to obser- vations of X-ray binaries, where a considerable matter flux along the accretion disc may be explained only when viscosity coefficient is much larger. It was sug- gested by Shakura (1972), that matter in the disc is turbulent, what determines a turbulent viscous stress tensor, parametrized by a pressure trφ = −αρv2 s = −αP, (3) where vs is a sound speed in the matter. This simple presentation comes out from a relation for a turbulent viscosity coefficient ηt ≈ ρvtl with an average turbulent velocity vt and mean free path of the turbulent element l. It follows from the definition of trφ in (2), when we take l ≈ h from (1) trφ = ρvthr dΩ dr ≈ ρvtvs = −αρv2 s , (4) where a coefficient α < 1 is connecting the turbulent and sound speeds vt = αvs. Presentations of trφ in (3) and (4) are equivalent, and only when the angu- lar velocity differs considerably from the Keplerian one the first relation to the right in (4) is more preferable. That does not appear in the standard theory, but may happen when advective terms are included. Development of a turbulence in the accretion disc cannot be justified simply, because a Keplerian disc is stable in linear approximation to the development of perturbations. It was suggested by Ya.B.Zeldovich, that in presence of very large Reynolds number Re = ρvl the amplitude of perturbations at which nonlin- η ear effects become important is very low, so a turbu- lence may be developed due to nonlinear instability even when the disc is stable in linear approximation. Viscous stresses may arise from a magnetic field, it was suggested by (Shakura, 1972), that magnetic stresses cannot exceed the turbulent ones. It was shown by Bisnovatyi-Kogan and Blinnikov (1976), that inner re- gions of a highly luminous accretion discs where pres- sure is dominated by radiation, are unstable to vertical convection. Development of this convection produce a turbulence, needed for a high viscosity. With alpha- prescription of viscosity the equation of angular momentum conservation is written in the plane of a disc as M (j − jin) = 4πr2αP0h. (5) When angular velocity is far from Keplerian one the relation (2) is valid with a coefficient of a turbulent viscosity η = αρ0vs0h, where values with the index "0" denote the plane of the disc. The choice of the viscosity coefficient is the most speculative problem of the theory. In the laminar case In the standard theory a heat balance is local, all heat produced by viscosity in the ring between r and Odessa Astronomical Publications, vol. 12 (1999) 3 r + dr is radiated through the sides of disc at the same r. The heat production rate Q+ related to the surface unit of the disc is written as Q+ = h trφr dΩ dr = 3 8π M GM r3 (cid:18)1 − jin j (cid:19) . (6) Heat losses by a disc depend on its optical depth. The standard disc model (Shakura, 1972) considered a ge- ometrically thin disc as an optically thick in a vertical direction. That implies energy losses Q− from the disc due to a radiative conductivity, after a substitution of the differential equation of a heat transfer by an alge- braic relation Q− ≈ 4 3 acT 4 κΣ . (7) Here a is a constant of a radiation energy density, c is a speed of light, T is a temperature in the disc plane, κ is a matter opacity, and a surface density Σ = 2ρh. Here and below ρ, T, P without the index "0" are related to the disc plane. The heat balance equation is repre- sented by a relation Q+ = Q−. A continuity equation in the standard model is used for finding of a radial velocity vr vr = M 4πrhρ = M 2πrΣ . (8) Completing these equations by an equation of state P (ρ, T ) and relation for the opacity κ = κ(ρ, T ) we get a full set of equations for a standard disc model. For power low equations of state of an ideal gas P = Pg = ρRT (R is a gas constant), or radiation pressure P = Pr = aT 4 3 , and opacity in the form of electron scattering κe, or Karammers formulae κk, the solution of a standard disc accretion theory is obtained analyti- cally. Checking the suggestion of a large optical thick- ness confirms a self-consistency of the model. Note that solutions for different regions of the disc with dif- ferent equation of states and opacities are not matched to each other. 2.2. Optically thin solution It was found by Shapiro et al. (1976) that there is another branch of the solution for a disc structure M , α which is also with the same input parameters M, self-consistent but has a small optical thickness. That implies another equation of energy losses, determined by a volume emission Q− ≈ q ρ h. The emissivity of the unit of a volume q is connected with a Planckian averaged opacity κp by a relation q ≈ acT 4 0 κp. In the optically thin limit the pressure is determined by a gas P = Pg. In the optically thin solution the thickness of the disc is larger then in the optically thick one, and density is lower. While heating by viscosity is determined mainly by heavy ions, and cooling is determined by electrons, the rate of the energy exchange between them is important for a structure of the disc. The energy balance equa- tions are written separately for ions and electrons. For small accretion rates and lower matter density the rate of energy exchange due to binary collisions is so slow, that in the thermal balance the ions are much hot- ter then the electrons. That also implies a high disc thickness. In the highly turbulent plasma the energy exchange between ions and electrons may be strongly enhanced due to presence of fluctuating electrical fields, where electrons and ions gain the same energy. In such conditions difference of temperatures between ions and electrons may be negligible. The theory of relaxation in the turbulent plasma is not completed, but there are indications to a large enhancement of the relaxation in presence of plasma turbulence, in comparison with the binary collisions (Galeev and Sagdeev, 1983; Quataert, 1997). 2.3. Accretion disc structure from equations describ- ing continuously optically thin and thick regions Equations of the disc structure smoothly describing transition between optically thick and optically thin disc, had been obtained using Eddington approxima- tion. The expressions had been obtained (Artemova et al., 1996) for the vertical energy flux from the disk F0, and radiation pressure in the symmetry plane F0 = 2acT 4 0 3τ0Φ , Prad,0 = aT 4 0 3Φ (cid:18)1 + 4 3τ0(cid:19) , (9) where τα0 = κpρh = 1 1 + 4 3τ0 In the optically thin limit τ∗ ≪ τ0 ≪ 1 we get 2 κpΣ0, τ∗ = (τ0τα0)1/2, Φ = . At τ0 ≫ τ∗ ≫ 1 we have (7) from (9). + 2 3τ 2 ∗ F0 = acT 4 0 τα0, Prad,0 = 2 3 acT 4 0 τα0. (10) Using F0 instead of Q− and equation of state P = ρRT + Prad,0, the equations of accretion disc struc- ture together with equation Q+ = F0, with Q+ from (6), have been solved numerically by Artemova et al. (1996). Two solutions, optically thick and thin, exist separately when luminosity is not very large. They in- tersect at m = mb and there is no global solution for ac- cretion disc at m > mb. It was concluded by Artemova et al. (1996) that in order to obtain a global physically meaningful solution at m > mb, account of advectiv- ion is needed. For the calculated case MBH = 108 M⊙, α = 1.0 at m = mb luminosity of the accretion disk is less than the critical Eddington one. 3. Accretion discs with advection 4 Odessa Astronomical Publications, vol. 12 (1999) Standard model gives somewhat nonphysical behav- ior near the inner edge of the accretion disc around a black hole, with a zero heat production at the in- ner edge of the disk. It is clear from physical ground, that in this case the heat brought by radial motion of matter along the accretion disc becomes important. In presence of this advective heating (or cooling) term, depending on the radial entropy S gradient, written as Qadv = dr , the equation of a heat balance is modified to M 2πr T dS Q+ + Qadv = Q−. (11) In order to describe self-consistently the structure of the accretion disc we should also modify the radial disc equilibrium, including pressure and inertia terms r(Ω2 − Ω2 K) = 1 ρ dP dr − vr dvr dr . (12) Appearance of inertia term leads to transonic radial flow with a singular point. Conditions of a contin- uous passing of the solution through a critical point choose a unique value of the integration constant jin. First approximate solution for the advective disc struc- ture have been obtained by Paczy´nski and Bisnovatyi- Kogan (1981). Attempts to find a solution for advec- tive disc had been done by Abramovicz et al. (1988), Matsumoto et al. (1984). For moderate values of M a unique continuous transonic solution was found, passing through singular points, and corresponding to a unique value of jin. The number of critical point in the radial flow with the gravitational potential φg (Paczy´nski and Wiita, 1980) φg = GM rg = 2GM . c2 r−rg may exceed unity. Appearance of two critical points for a radial flow in this potential was analyzed by Chakrabarti and Molteni (1993). Using of equations averaged over a thickness of the disc changes a struc- ture of hydrodynamic equations, leading to a position of singular points not coinciding with a unit Mach num- ber point. , Kogan and Ruzmaikin (1974). For a spherical accre- tion with v = (vr, 0, 0) the equations describing a field amplification in the ideally conducting plasma reduce to (Bisnovatyi-Kogan, Ruzmaikin, 1974) d(r2Br) dt = 0, d(rvrBθ) dt = 0, d(rvrBφ) dt = 0, (13) ∂t + vr where d dt = ∂ ∂ ∂r is a full Lagrangian deriva- tive. Consider a free fall case with vr = −q 2GM . The initial condition problem is solved separately for poloidal and toroidal fields. For initially uniform field Br0 = B0 cos θ, Bθ0 = −B0 sin θ we get the solution (Bisnovatyi-Kogan, Ruzmaikin, 1974) r Br = B0 cos θ r2 Φ4/3 1 , Bθ = − B0 sin θ √r Φ1/3 1 , (14) 2 t√2GM . The radial component of where Φ1 = r3/2 + 3 the field is growing most rapidly. It is ∼ r−2 for large times, ∼ t4/3 at given small radius, and is growing with time everywhere. For initially dipole magnetic field Br0 = B0 cos θ r3 , Bθ0 = − B0 sin θ 2r3 we obtain the following solution Br = B0 cos θ r2 Φ−2/3 1 , Bθ = − B0 sin θ 2√r Φ−5/3 1 . (15) Here the magnetic field is decreasing everywhere with time, tending to zero. That describes a pressing of a dipole magnetic field to a stellar surface. The az- imuthal stellar magnetic field if confined inside the star. When outer layers of the star are compressing with a free-fall speed, then for initial field distribution Bφ0 = B0rnf (θ) the change of Bφ with time is de- scribed by a relation Bφ = − B0 f (θ) √r Φn+1/3 1 . 5. Two-temperature advective discs 4. Amplification of the magnetic field at a spherical accretion A matter flowing into a black hole is usually mag- netized. Due to more rapid increase of a magnetic en- ergy the role of the magnetic field increases when mat- ter flows inside. It was shown by Schwartsman (1971), that magnetic energy density EM approaches a density of a kinetic energy Ek, and he proposed a hypothesis of equipartition EM ≈ Ek, supported by continuous annihilation of the magnetic field in a region of the main energy production. This hypothesis is usually accepted in the modern picture of accretion (Narayan and Yu, 1995). Account of the heating of matter by magnetic field annihilation was done by Bisnovatyi- In the optically thin accretion discs at low mass fluxes the density of the matter is low and energy ex- change between electrons and ions due to binary colli- sions is slow. In this situation, due to different mech- anisms of heating and cooling for electrons and ions, they may have different temperatures. First it was real- ized by Shapiro et al. (1976), where advection was not included. It was noticed by Narayan and Yu (1995), that advection in this case is becoming extremely im- portant. It may carry the main energy flux into a black hole, leaving rather low efficiency of the accretion up to 10−3 − 10−4 (advective dominated accretion flows - ADAF). This conclusion is valid only when the ef- fects, connected with heating of matter by magnetic field annihilation are neglected. Odessa Astronomical Publications, vol. 12 (1999) 5 In the ADAF solution the ion temperature is about a virial one kTi ∼ GM mi/r, what means that even at high initial angular momentum the disc becomes thick, forming a quasi-spherical accretion flow. When energy losses by ions are low, some kind of a "thermo- viscous" instability is developed, because heating in- creases a viscosity, and viscosity increases a heating. Development of this instability leads to formation of ADAF. A full account of the processes, connected with a presence of magnetic field in the flow, is changing considerably the picture of ADAF. It was shown by Schwarzman (1971), that in the region of the main en- ergy production, the condition of equipartition takes place, and efficiency of a radiation increase enor- mously from ∼ 10−8 up to ∼ 0.1 due to magneto- bremstrahlung. To support the condition of equiparti- tion a continuous magnetic field reconnection and heat- ing of matter due to Ohmic dissipation takes place. It was shown by Bisnovatyi-Kogan and Ruzmaikin (1974), that due to Ohmic heating the efficiency of a radial accretion into a black hole may become as high as ∼ 30%. The rate of the Ohmic heating in the con- dition of equipartition was obtained in the form T dS dr σE 2 ρvr ≈ −σ v2 EB2 ρvrc2 = − = − B2v2 E 4πραmvrvslt , (18) what coincides with (16) when αm = 4rv2 E , or lt = 3vr vslt 4rv2 E 3vr vsαm . Here a local electrical field strength in a highly conducting plasma is of the order of E ∼ vE B , vE ∼ vt ∼ αvs for a radial accretion. Equations for a radial temperature dependence in the accretion disc, separately for the ions and electrons are written as c dEi dt − dρ dt Pi ρ2 dρ dt = Hηi + HBi − Qie , (19) dEe dt − Pe ρ2 = Hηe + HBe + Qie − Cf f − Ccyc , (20) A rate of a viscous heating of ions Hηi is obtained from (6) as Hηi = 2πr M Q+ = 3 2 α vKv2 s r , Hηe ≤ r me mi Hηi. (21) Combining (1),(8),(5), we get T dS dr 3 2 B2 8πρr . = − (16) vr = α r r 2 = ρGM 8π ≈ ρv2 In the supersonic flow of the radial accretion equipar- tition was suggested in a form (Schwartsman, 1971) B2 . For the disc accretion equiparti- tion between magnetic and turbulent energy was sug- gested by Shakura (1972), what reduces with account of "alpha" prescription of viscosity to a relation B2 8π ∼ ρv2 2 ≈ α2 mP , where αm characterizes a magnetic viscos- t ity in a way similar to the turbulent α viscosity. It was suggested by Bisnovatyi-Kogan and Ruzmaikin (1976) the similarity between viscous and magnetic Reynolds numbers, or between turbulent and magnetic viscosity coefficients Re = ρvl , where the turbu- lent magnetic viscosity ηm is connected with a turbu- lent conductivity σ = ρc2 α η, we get 4πηm a turbulent conductivity η , Rem = ρvl . Taking ηm = αm ηm σ = c2 4παmhvs , v2 s = Pg ρ (17) in the optically thin discs. For the radial accretion the turbulent conductivity may contain mean free path of a turbulent element lt in (17) instead of h. In ADAF solutions, where ionic temperature is of the order of the virial one two above suggestions for magnetic equipar- tition almost coincide at αm ∼ 1. The heating of the matter due to an Ohmic dissi- pation may be obtained from the Ohm's law in radial accretion v2 s vKJ h = √2 , vs vK r, ρ = M 4πα√2 v2 KJ r2v3 s , (22) where vK = rΩK , J = 1 − jin j . The rate of the energy exchange between ions and electrons due to the binary collisions was obtained by Landau (1937). Neglecting pair formation for a low density accretion disc, we may write an exact expression for a pres- sure Pg = Pi + Pe = nikTe + nekTp = nek(Te + Ti), and an approximate expression for an energy, con- taining a smooth interpolation between nonrelativis- tic and relativistic electrons. The bremstrahlung Cf f and magneto-bremstrahlung Ccyc cooling of maxwellian semi-relativistic electrons, with account of free-bound radiation in nonrelativistic limit, may be written as interpolation of limiting cases (Bisnovatyi-Kogan and Ruzmaikin, 1976). K, v2 B=v2 r , and v2 In the case of a disk accretion there are several char- acteristic velocities, vK, vr, vs, and vt = αvs, all of which may be used for determining "equipartition" magnetic energy, and one characteristic length h. Con- sider three possible choices of v2 t for scaling B2 = 4πρv2 B. The expression for an Ohmic heating in the turbulent accretion disc also may be written in different ways, using different velocities vE in the expression for an effective electrical field E = vE B . A self-consistency of the model requires, that expres- sions for a magnetic heating of the matter HB, ob- tained from the condition of stationarity of the flow (16), and from the Ohm's law (18), should be iden- 4 . That implies tical. It happens at = 3√2 α c v2 E v2 r J αm 6 Odessa Astronomical Publications, vol. 12 (1999) s < vt. B2 rρ vr = 1 vK√J ≃ vtvs vK√J vE ∼ vr ∼ αv2 In the advec- tive models J is substituted by a function which is not zero at the inner edge of the disc. The heating due to magnetic field reconnection HB in the equa- tions (19), (20) may be written with account of (21) as HB = 3 . At vB = vK the ex- pressions for viscous and magnetic heating are almost identical. Observations of the magnetic field recon- nection in the solar flares show (Tsuneta, 1996), that electronic heating prevails over the ionc one. Transfor- mation of the magnetic energy into a heat is connected with the change of the magnetic flux, generation of the vortex electrical field, accelerating the particles. vK(cid:17)2 2J Hηi(cid:16) vB 16π The equations (19), (20) have been solved by Bisnovatyi-Kogan and Lovelace (1997) for nonrela- tivistic electrons, at vB=vK. The combined heat- ing of the electrons and ions were taken as He = (2 − g)Hηi, He = gHηi. The results of calculations for g = 0.5 ÷ 1 show that almost all energy of the electrons is radiated, so the relative efficiency of the two-temperature, optically thin disc accretion cannot become lower then 0.25. Increase of the term Qie due to plasma turbulence may restore the relative efficiency to a value, corresponding to the optically thick disc. 6. Discussion Observational evidences for existence of black holes inside our Galaxy and in the active galactic nuclei (Cherepashchuk, 1996; Ho, 1999) make necessary to revise theoretical models of the disc accretion. The improvements of a model are connected with account of advective terms and more accurate treatment of the magnetic field effects. Account of the effects connected with magnetic field annihilation does not permit to make a relative efficiency of the accretion lower then ∼ 0.25 from the standard value. Strong relaxation con- nected with the plasma turbulence may increase the efficiency, making it close to unity. For explanation of underluminous galactic nuclei two possible ways may be suggested. One is based on a more accurate esti- mations of the accretion mass flow into the black hole, which could be overestimated. The second is based on existence of another mechanisms of the energy losses in the form of accelerated particles, like in the radio- pulsars, where these losses exceed strongly a radiation. This is very probable to happen in a presence of a large scale magnetic field which may be also respon- sible for a formation of the observed jets (Bisnovatyi- Kogan, 1999; Blandford and Begelman, 1999). We may suggest, that underlumilnous AGN loose main part of their energy to the formation of jets, like in SS 433. The search of the correlation between existence of jets and lack of the luminosity could be very informative. Acknowledgements. The author is grateful for par- tial support to RFBR, grant 99-02-18180, and GPNT "Astronomy", grant 1.2.6.5. References Abramovicz M.A., Czerny B., Lasota J.P., Szuszkiewicz E.: 1988, ApJ, 332, 646. Artemova I.V., Bisnovatyi-Kogan G.S., Bjornsson G., Novikov I.D.: 1996, ApJ, 456, 119. Bisnovatyi-Kogan G.S.: black evidences ed. S.Chakrabarti, Kluwer, p.1. 1999, holes for in "Observational in the universe", Bisnovatyi-Kogan G.S., Blinnikov S.I.: 1976, Pisma Astron. Zh., 2, 489. Bisnovatyi-Kogan G.S., Lovelace R.V.L.: 1997, ApJL, 486, L43. Bisnovatyi-Kogan G.S., Lovelace R.V.L.: 1999, ApJ (accepted), astro-ph/9902344. Bisnovatyi-Kogan G.S., Ruzmaikin A.A.: 1974, Ap. and Space Sci., 28, 45. Bisnovatyi-Kogan, G.S., Ruzmaikin, A.A., 1976, Ap. and Space Sci., 42, 401. Blandford R.D., Begelman M.C.: 1999, Month. Not. R.A.S., 303, 1. Cannizzo J.K. et al.: 1998, AAS Abstracts, 192, 4103C. Chakrabarti S.K., Molteni D.: 1993, ApJ, 417, 671. Cherepashchuk A.M.: 1996, Uspekhi Fiz. Nauk., 166, 809. Di Matteo T., Fabian A.C., Rees, M.J. et al.: 1999, Month. Not. R.A.S., 305, 492. Galeev A.A., Sagdeev R.Z.: 1983, Chap. Handbook of Plasma Physics, Vol. Rosenbluth M.N., Sagdeev R.Z. North-Holland Pub.), Ch. 6.1 6, in eds. (Amsterdam: 1, Ho L.: 1999, in "Observ. evidences for b. h. in the Universe", ed. S.Chakrabarti, Kluwer, p.157. Landau L.D.: 1937, Zh. Exp. Theor. Phys. 7, 203. Matsumoto R., Sato Sh., Fukue J., Okazaki A.T.: 1984, Publ. Astron. Soc. Japan, 36, 71. Narayan R., Yu I.: 1995, ApJ, 452, 710. Novikov I.D., Thorne K.S.: in Black Holes eds. C.DeWitt, B.DeWitt (New York: Gordon & Breach), p.345 1973, Paczy´nski B., Bisnovatyi-Kogan G.S.: 1981, Acta Astron., 31, 283. Paczy´nski B., Wiita P.J.: 1980, A&A, 88, 23. Pringle J.E., Rees, M.J.: 1972, A&A, 21, 1. Quataert E.: 1997, astro-ph/9710127. Schwartsman V.F.: 1971, Soviet Astron., 15, 377. Shakura N.I.: 1972, Astron. Zh., 49, 921; (1973, Sov. Astron., 16, 756) Shakura N.I., Sunyaev R.A.: 1973, A&A, 24, 337. Shapiro S.L., Lightman A.P., Eardley D.M.: 1976, ApJ, 204, 187. Tsuneta S.: 1996, ApJ, 456, 840.
astro-ph/0409628
1
0409
2004-09-27T11:01:16
The cosmological significance of Low Surface Brightness galaxies found in a deep blind neutral-hydrogen survey
[ "astro-ph" ]
We have placed limits on the cosmological significance of gas-rich low surface-brightness (LSB) galaxies as a proportion of the total population of gas-rich galaxies by carrying out a very deep survey (HIDEEP) for neutral hydrogen (HI) with the Parkes multibeam system. Such a survey avoids the surface-brightness selection effects that limit the usefulness of optical surveys for finding LSB galaxies. To complement the HIDEEP survey we have digitally stacked eight 1-hour R-band Tech Pan films from the UK Schmidt Telescope covering 36 square degrees of the survey area to reach a very deep isophotal limit of 26.5 R mag/sq. arcsec. At this level, we find that all of the 129 HI sources within this area have optical counterparts and that 107 of them can be identified with individual galaxies. We have used the properties of the galaxies identified as the optical counterparts of the HI sources to estimate the significance of LSB galaxies (defined to be those at least 1.5 magnitudes dimmer in effective surface-brightness than the peak in the observed distribution seen in optical surveys). We calculate the contribution of LSB galaxies to the total number, neutral hydrogen density, luminosity density, baryonic mass density, dynamical mass density and cross-sectional area of gas-rich galaxies. We do not find any `Crouching Giant' LSB galaxies such as Malin 1, nor do we find a population of extremely low surface-brightness galaxies not previously found by optical surveys. Such objects must either be rare, gas-poor or outside the survey detection limits.
astro-ph
astro-ph
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000) Printed 14 November 2018 (MN LATEX style file v2.2) The cosmological significance of Low Surface Brightness galaxies found in a deep blind neutral-hydrogen survey R. F. Minchin,1⋆ M. J. Disney,1 Q. A. Parker,2 P. J. Boyce,3 W. J. G. de Blok,1 G. D. Banks,1† R. D. Ekers 4 K. C. Freeman,5 D. A. Garcia,1 B. K. Gibson, 6 M. Grossi,1 R. F. Haynes,7 P. M. Knezek,8 R. H. Lang,1 D. F. Malin,9 R. M. Price,10 M. Putman,11 I. M. Stewart,12 A. E. Wright4 1 School of Physics and Astronomy, Cardiff University, 5 The Parade, Cardiff, CF24 3YB 2 Department of Physics, Macquarie University, Sydney, NSW 2109, Australia 3 Planning Division, Cardiff University, Park Place, Cardiff, CF10 3UA 4 Australia Telescope National Facility, PO Box 76, Epping, NSW 1710, Australia 5 Research School of Astronomy & Astrophysics, Mount Stromlo Observatory, Cotter Road, Weston ACT 2611, Australia 6 Centre for Astrophysics and Supercomputing, Swinburne University of Technology, PO Box 218, Hawthorn, Victoria 3122, Australia 7 School of Mathematics & Physics, University of Tasmania, Hobart, Tasmania 7001, Australia 8 WIYN Consortium Inc., 950 North Cherry Avenue, Tucson, AZ 85719, United States 9 Anglo-Australian Observatory, P.O. Box 296, Epping, NSW 1710, Australia 10 Department of Physics and Astronomy, University of New Mexico, 800 Yale Boulevard NE, Albuquerque, NM, United States 11 Center for Astrophysics and Space Astronomy, University of Colorado, Campus Box 389, Boulder, CO, United States 12 Department of Physics and Astronomy, University of Leicester, University Road, Leicester LE1 7RH 14 November 2018 ABSTRACT We have placed limits on the cosmological significance of gas-rich low surface- brightness (LSB) galaxies as a proportion of the total population of gas-rich galaxies by carrying out a very deep survey (HIDEEP; Minchin et al. 2003) for neutral hydrogen (Hi) with the Parkes multibeam system. Such a survey avoids the surface-brightness selection effects that limit the usefulness of optical surveys for finding LSB galaxies. To complement the HIDEEP survey we have digitally stacked eight 1-hour R-band Tech Pan films from the UK Schmidt Telescope covering 36 square degrees of the sur- vey area to reach a very deep isophotal limit of 26.5 R mag arcsec−2. At this level, we find that all of the 129 Hi sources within this area have optical counterparts and that 107 of them can be identified with individual galaxies. We have used the properties of the galaxies identified as the optical counterparts of the Hi sources to estimate the significance of LSB galaxies (defined to be those at least 1.5 magnitudes dimmer in effective surface-brightness than the peak in the observed distribution seen in optical surveys). Two different methods of correcting for ease-of-detection do not yield signif- icantly different results: LSB galaxies make up 62 ± 37 per cent of gas-rich galaxies by number according to our first method (weighting by Hi mass function), which includes a correction for large scale structure, or 51±20 per cent when calculated by our second method (1/Vmax correction). We also find that LSB galaxies provide 30 ± 10 per cent of the contribution of gas-rich galaxies to the neutral hydrogen density of the Universe, 7 ± 3 per cent of their contribution to the luminosity density of the Universe; 9 ± 4 of their contribution to the baryonic mass density of the Universe, 20 ± 10 per cent of their contribution to the dynamical mass density of the Universe and 40 ± 20 per cent of their cross-sectional area. We do not find any 'Crouching Giant' LSB galaxies such as Malin 1, nor do we find a population of extremely low surface-brightness galaxies not previously found by optical surveys. Such objects must either be rare, gas-poor or outside the survey detection limits. Key words: surveys -- galaxies: luminosity function -- radio lines: galaxies -- galaxies: fundamental parameters 2 R. F. Minchin et al. 1 INTRODUCTION Surveys in the optical are known to be affected by strong selection effects (e.g. Disney 1976; Disney & Phillipps 1983; McGaugh 1996) which bias our understanding of the local galaxy population. Corrections for these selection effects can be made, but these are large and controversial and must be applied to small numbers of sources, leading to large uncer- tainties (Impey & Bothun 1997; Disney 1999). Low Surface- Brightness (LSB) galaxies could make a significant contri- bution to the Universe that would not be recognised in an optical survey. They may dominate the luminosity, baryon or mass density of the galaxies in the Universe (e.g. Fukugita, Hogan, & Peebles 1998) and could contribute significantly to QSO absorption line spectra (e.g. Churchill & Le Brun 1998). Zwaan et al. (2003) derived the Hi mass function for the 1000 galaxies in the Hi Parkes All Sky Survey (HIPASS) Bright Galaxy Catalogue (Koribalski et al. 2004) and put a value of 15 per cent on the contribution of LSB galaxies to the neutral hydrogen density of the Universe. However, this number depends crucially on their assumption that optical catalogues are as (in)complete for LSB galaxies as for high surface-brightness (HSB) galaxies. This is obviously not the case as LSB galaxies are much harder to detect than HSB galaxies. Indeed, most of the new galaxies, outside of the zone-of-avoidance, found in the HIPASS Bright Galaxy Cat- alogue are LSB galaxies (Ryan-Weber et al. 2002). Thus the value of Zwaan et al. should be considered only a lower limit to the contribution of LSB galaxies to the neutral hydrogen density. The idea of searching for galaxies via the 21-cm neutral hydrogen line has long been considered as an alternative to optical surveys (e.g. Disney 1976). However, until recent ad- vances in technology such as multibeam receivers and pow- erful correlators, such a survey has not been practical. The HIDEEP survey (Minchin et al. 2003; Paper 1 hereafter) was motivated by the desire to reach previously inaccessible surface-brightness levels. If optical surface- brightness (e.g. luminosity per square arcsec) were to corre- late with hydrogen column-density (e.g. Hi flux per square arcsec), then to reach low surface-brightnesses it would be necessary to reach low column-densities. With an integra- tion time of 9000 seconds per beam, HIDEEP is significantly more sensitive to low column-density gas than any previous survey and thus potentially more sensitive to low surface- brightness galaxies. HIDEEP followed the same survey strategy and data- reduction path as HIPASS (Barnes et al. 1998), but with twenty times the integration time. The Parkes multibeam system was actively scanned in declination to cover a 6◦×10◦ region, with full sensitivity being reached over the central 4◦ × 8◦. Scans were interleaved to give Nyquist sampling and repeated to reach the required depth. The data were treated in the same way as HIPASS data: they were re- duced in AIPS++ using the LiveData and Gridzilla pro- grams and continuum sources were removed by template- fitting with the Luther program. In order to minimise the noise, all the observations were carried out at night. The full sensitivity reached was 3.2 mJy beam−1, which is consistent with a √t improvement on HIPASS. To complement the deep Hi data, we obtained eight 1- hour exposures on Tech Pan films at the UK Schmidt Tele- scope (UKST). These covered a sky area of 6◦ × 6◦, with the same centre as the Hi survey data. The films were digi- tised using the SuperCOSMOS machine at the Royal Ob- servatory, Edinburgh (Hambly et al. 2001). Digital stacking of these films brings a √t improvement in signal to noise (Bland-Hawthorn, Shopbell & Malin 1993), thus gaining us a little over a magnitude in depth. The use of Tech Pan films gives a further magnitude improvement over the III-aF R- band plates previously used at the UKST (Parker & Malin 1999). For galaxies with ordinary colours, our 1σ surface- brightness limit of 26.5 R mag arcsec−2 is equivalent to ∼ 27.5 B mag arcsec−2. The bandpass limit of the multibeam system means that only objects with a heliocentric velocity less than 12,700 km s−1 can be detected (173 Mpc for H0 = 75 km s−1 Mpc−1, after correction to CMB rest frame velocities), thus the vol- ume of the 36 square degree optical overlap region is 19,000 Mpc3. Within this volume we find a total of 129 sources, of which 107 can be identified with individual galaxies. 2 ANALYSIS OF THE OPTICAL DATA In Paper 1, we identified sources found in the Hi data with optical counterparts on the Tech Pan films. A comparison of the offsets of the optical positions from the Hi positions for the sources with secure identifications (those with cat- alogued velocities matching the Hi velocity and those con- firmed by Hi interferometry or optical spectroscopy follow- up observations) with the offsets for less certain counterparts (those without catalogued velocities or follow-up observa- tions) showed no significant difference. Of the 107 galaxies in this paper, 87 are secure identifications; this includes 18 confirmed with our interferometric follow-up and 2 with our spectroscopic follow-up. Of the remaining 20, 18 are previ- ously catalogued galaxies that do not have redshifts in the literature and two are new detections. An Hi-selected sample is expected to appear different from an optically-selected one. In particular low surface- brightness galaxies -- which are seen over much smaller vol- umes than high surface-brightness galaxies with the same to- tal luminosity in optical surveys (Disney & Phillipps 1983) -- are expected to be seen much more easily. However, gas-poor galaxies, such as ellipticals and dwarf spheroidals, which are found by optical surveys will be invisible at 21-cm. We find that this is indeed the case with our sample. As expected, our surface-brightness distribution (Figure 1) is higher to- wards lower surface-brightnesses than one finds in optically selected samples such as the ESO-LV (Lauberts & Valentijn 1989), which is also shown in the figure. A Kolmogorov- Smirnov test on our observed surface-brightness distribu- tion and the surface-brightness distribution of the ESO-LV, which also uses R-band effective surface-brightness, shows that the hypothesis that both are drawn from the same par- ent population has a significance of less than 1 per cent, due to the larger number of LSB galaxies seen in the HIDEEP sample. This confirms that Hi surveys do, as expected, avoid some of the surface-brightness selection effects present in op- tical surveys. Cosmological Significance of LSB galaxies 3 Figure 1. Number of galaxies found in each surface-brightness bin. The solid line shows the distribution of observed surface- brightnesses, the dashed line shows the distribution of surface- brightnesses after correction for galactic absorption, cosmological dimming, and inclination. The dotted line shows the observed surface-brightness distribution of ESO-LV galaxies (right hand scale). 2.1 Tabulated Optical Properties Optical photometry of HIDEEP galaxies was carried out us- ing a combination of sextractor (Bertins & Arnouts 1996) and the stsdas package in iraf, where the ellipse task was used to obtain radial surface-brightness profiles. The mea- sured optical parameters are displayed in Table 1 (a sample is shown here, the full table is available in the online edi- tion) where column (1) gives the source name and (2) the catalogued identification -- if there is one. Column (3) gives the type of the identification graded into 3 classes: "ID" de- notes objects where both positions and velocities coincide, "PCG" (for Previously Catalogued Galaxy) where there is a positional coincidence with a previously catalogued galaxy in the NASA Extragalactic Database (NED) and "PUG" (for Previously Uncatalogued Galaxy) where there is no possible counterpart in NED, and thus no previous posi- tional or redshift data. Three further sub-classes, "ID-VLA", "ID-ATCA", and "ID-Spec" denote where a previously cat- alogued galaxy without a redshift has been confirmed as the optical counterpart through our follow-up observations (by interferometry with the VLA or the ATCA or by spec- troscopy at the ANU 2.3-m telescope at Siding Spring re- spectively). The tabulated optical parameters are as follows: column (4) gives the apparent magnitude (mR) as measured by the mag best parameter in sextractor. This takes the form of either an adaptive aperture magnitude (Kron 1980), where the size of the aperture is 2.5 times the first-moment radius of the galaxy, or, for crowded fields, an isophotal magni- tude (to the 26 R mag arcsec−2 isophote) with a correction made for the part of the galaxy beyond the isophotal limit 4 R. F. Minchin et al. Tabe 1: i a eie f DEE gaaxie. Thi i a ae f he abe { he f abe i avaiabe i he ie edii D Caag e Ca   (cid:22) (cid:15) A  R (cid:22)   R e e R R e e a e  R(cid:22) k  R(cid:22) 1 2 3 4 5 6 7 8 9 10 11 DEE 1326 3024 C 4247 D 13.8 16.7 22.4 0.57 0.17 15.0 0.98 22.7 DEE 1326 3209 ES 444 G033 D VA 14.0 23.9 22.4 0.74 0.14 19.5 4.2 23.0 DEE 1327 2713 ES 509 G014 CG 16.5 4.3 22.4 0.40 0.15 16.5 0.72 22.5 DEE 1327 2935 GC 5150 D 12.4 11.7 20.9 0.24 0.15 21.6 3.1 20.8 DEE 1327 3006 UGCA 358 D 14.5 21.7 24.4 0.56 0.16 18.4 3.1 24.7 DEE 1328 2735 ES 509 G026 D 16.5 7.3 24.1 0.32 0.14 16.1 1.0 24.1 DEE 1328 2819 [BZZ2000℄ 132823.03 2811814.7 D 16.6 8.0 24.0 0.34 0.14 16.4 1.3 24.1 DEE 1328 3152 ES 444 G047 D 13.7 12.7 22.1 0.65 0.14 20.8 3.8 22.5 DEE 1329 2714 ES 509 G031 CG 17.3 9.7 25.3 0.10 0.15 16.4 2.4 25.2 DEE 1329 3144 ABE 3558:[G94℄ 2773 D VA 18.1 5.2 24.1 0.20 0.14 16.2 1.6 24.0 DEE 1329 3203 ES 444 G056 D 16.3 7.3 23.4 0.12 0.13 18.0 2.3 23.2 DEE 1330 2755 C 4264 D 14.2 10.7 22.2 0.60 0.14 19.8 2.6 22.5 DEE 1330 2809 GC 5182 D 12.0 30.0 22.0 0.53 0.15 22.5 9.0 22.2 DEE 1330 2936 2ASX 13303473 2934210 D VA 14.6 11.6 23.1 0.52 0.15 19.6 3.1 23.3 DEE 1330 3212 ES 444 G059 CG 16.8 12.1 25.1 0.11 0.14 17.2 3.3 25.0 DEE 1330 3233 ES 444 G061 CG 14.8 10.7 23.2 0.07 0.14 21.5 8.4 22.9 DEE 1330 3259 UG 14.9 26.7 24.8 0.45 0.15 19.1 7.9 24.9 DEE 1331 2804 [ R95℄1325 27 50 D 13.6 8.0 21.4 0.24 0.15 22.4 5.4 21.6 DEE 1331 2944 C 4275 D 13.1 11.4 21.3 0.39 0.15 21.2 3.4 21.3 DEE 1331 3155 ES 444 G066 D 14.9 11.6 23.7 0.09 0.14 19.3 3.6 23.6 DEE 1331 3205 Abe 3558:[G94℄3325 CG 16.9 6.6 23.9 0.35 0.13 15.7 0.90 24.0 DEE 1331 3258 GSC 7269 01680 D 14.4 7.3 22.0 0.34 0.15 19.5 1.8 22.0 DEE 1332 2727 ES 509 G048 D 14.0 9.3 22.5 0.10 0.17 22.3 6.6 22.4 DEE 1332 3141 Abe 3558:[G94℄3535 D 15.7 9.3 23.5 0.09 0.14 20.4 6.6 23.3 DEE 1333 2727 RAS F13304 2714 CG 16.3 4.5 22.9 0.35 0.18 19.9 3.1 22.8 DEE 1333 2743 C 4286 CG 15.0 5.2 21.9 0.11 0.17 20.9 3.4 21.7 DEE 1333 2901 [D99℄ 13300.7 284540 D 15.6 10.7 22.8 0.62 0.15 19.5 4.1 23.0 DEE 1333 3158 ES 444 G070 D 14.0 8.4 21.9 0.20 0.14 22.2 6.2 21.7 DEE 1333 3243 ES 444 G071 D 13.4 15.9 21.6 0.68 0.15 21.1 4.6 22.0 (Maddox et al. 1990), the best algorithm being selected auto- matically by the program. Simulations by Bertin & Arnouts (1996) of an R-band CCD image with an isophotal limit of 26 R mag arcsec−2 (i.e. very similar to our image) show that, to the magnitude level of the faintest galaxy in our survey (18.6 R mag), the true magnitude is recovered to better than 0.1 mag. Column (5) gives the effective radius, re -- the radius enclosing half the R-band light. Column (6) gives the effec- tive surface-brightness µe -- the R-band surface-brightness at the effective radius. This gives a model-independent measure of the surface brightness which can be applied to all types of galaxies. Column (7) gives the sextractor ellipticity (ǫ = 1 − b/a) and (8) the estimated R-band absorption at the position of the object from Schlegel et al. (1998) sup- plied by NED. Column (9) gives the absolute magnitude MR, including corrections for galactic absorption, cosmo- logical dimming and internal absorption (estimated using Ai,R = 0.95 log(a/b), where a/b is calculated from ǫ). The k-correction out to the bandpass limit of 12,700 km s−1 is negligible. We use the distances from Paper 1 except for galaxies in the Centaurus A group (taken to be those within 1000 km s−1), where we use the distance to the M83 sub- group of 4.5 Mpc from Thim et al. (2003). Column (10) gives the physical effective radius Re in kpc. Finally col- umn (13) gives the surface brightness µe(c) corrected for (1 + z)4 cosmological dimming, inclination (estimated using Ci,R = −1.25 log(a/b) from Graham & de Blok (2001)) and galactic absorption. Table 2 gives the positions of the optical counterparts of the PUGs. Where these have been confirmed by interfer- ometric follow up, this is indicated by a code in column (4) giving the run in which they were detected. The PUGs that were not detected in earlier runs were followed up again in later runs, with the result that there are no galaxies in the table that have been followed-up but not detected, although there are two sources which have not been followed-up in- terferometrically. 3 VOLUMETRIC AND LARGE SCALE STRUCTURE CORRECTED DATA No conclusions can be drawn as to the cosmic significance of the LSB galaxies found in HIDEEP until the raw numbers have been corrected for ease-of-Hi-detection and for the large scale structure in the region. The peak flux (Speak) of a galaxy determines the distance to which it may be seen in an Hi survey and is related to its Hi mass (MHI) and its velocity width (∆V ) as MHI ∝ d2∆V Speak. However, the velocity width is not independent of the mass of a galaxy -- these are related as ∆V ∝ M 1/3 HI (Briggs & Rao 1993), thus MHI ∝ d3S3/2 peak. For a fixed limiting peak flux, we therefore get MHI,lim ∝ d3. While this is an approximate empirical relationship originally defined using an optically- Cosmological Significance of LSB galaxies 5 Table 2. Positions of newly-discovered galaxies in the HIDEEP optical area. For those detected interferometrically, Column (4) gives the code for the observing run: A1 - ATCA, 11/1999; A2 - ATCA, 01/2000; V - VLA, 01-03/2003; A3 - ATCA, 04/2003. Those without a code have not been observed with an interfer- ometer. HIDEEP ID R.A. (1) (2) Decl. (3) Run (4) J1330-3259 J1336-2932 J1337-3118 J1338-3035 J1339-3022 J1342-2859 J1342-3021 J1344-3202 J1345-2908 J1345-3041 J1345-3104 J1347-2735 J1347-2810 J1348-2856 13:30:32 13:36:08 13:36:59 13:37:56 13:39:08 13:42:05 13:42:21 13:44:03 13:45:30 13:45:20 13:45:12 13:47:37 13:47:45 13:48:38 -32:57:07 V -29:34:12 A1 -31:19:05 V -30:35:12 A2 -30:22:10 V -29:01:21 A2 -30:23:14 V -32:02:11 - -29:06:47 V -30:43:21 - -30:56:52 V -27:35:22 V -28:11:56 A3 -29:00:22 V selected sample of galaxies, the general result that higher Hi-mass galaxies can be seen to greater distances is seen to hold true in blind Hi surveys (e.g. Zwaan et al. 1997 or Koribalski et al. 2004). The exact form of this dependence of limiting Hi mass on distance is not used in our analysis and is thus unimportant for the purposes of this study. We find that parameters such as absolute magnitude and surface brightness are correlated with hydrogen mass (Figures 2 and 3). Thus more luminous galaxies may be seen over larger volumes and, in a reversal of the situation in the optical, low surface-brightness galaxies may be seen over larger volumes than higher surface-brightness galaxies of the same luminosity due to the anti-correlation between surface-brightness and the Hi-mass to light ratio (Figure 4) seen both in our data and in previous studies using optically- selected samples (e.g.de Blok, McGaugh & van der Hulst, 1996). In order to make the volumetric corrections for Hi se- lection effects, we use two different methods, which are de- scribed below. We have chosen to remove dwarf galaxies (MHI < 108M⊙) from this analysis for two main reasons: that the number of low-mass galaxies in our sample is low and that the volume in which we can see these galaxies in dominated by the Cen A group. For these reasons, we fo- cus solely on the more Hi-massive galaxies (MHI > 108M⊙) where we have reasonable statistics and probe a variety of environments. 3.1 1/Vmax Weighting Weighting detections by 1/Vmax, where Vmax is the total volume in which they could have been found, is well estab- lished as a means of correcting for ease-of-detection. We use this here with a completeness limit of 18 mJy in peak flux (Paper 1) and a maximum distance, due to the bandpass limit of the Hi survey, of 173 Mpc. This gives us a complete sample of 67 galaxies with hV /Vmaxi = 0.49 ± 0.04. Once the dwarfs are removed, we are left with 62 galaxies with hV /Vmaxi = 0.52 ± 0.04. While 1/Vmax is well understood, Figure 2. Relationship between Hi mass and absolute magnitude Figure 3. Relationship between Hi mass and effective surface- brightness. Cen A group galaxies are shown by solid circles. The relationship is weaker than for luminosity, but there is a def- inite trend for lower Hi mass galaxies to have lower surface- brightnesses: 14 of the 23 galaxies with MHI < 109M⊙ are LSB galaxies (µe(R) > 23.3 R mag arcsec−2), while, at the high- mass end, there is only one LSB galaxy out of the sixteen with MHI > 1010M⊙. This dependence of surface-brightness on Hi mass means that corrections must be made for Hi selection effects before the surface-brightness distribution can be determined. 6 R. F. Minchin et al. Table 3. Hi Mass Functions used in calculating the weighting (corrected to H0 = 75 km s−1 Mpc−1). α -1.30 -1.53 -1.52 -1.51 -1.2 φ⋆ 0.0086 0.005 0.0032 0.006 0.006 HI log M ⋆ 9.79 9.88 10.1 9.7 9.8 Reference Zwaan et al. 2003 (BGC) Rosenberg & Schneider 2002 Kilborn 2001 Henning et al. 2000 Zwaan et al. 1997 Figure 4. Correlation of effective surface-brightness with Hi mass to light ratio. Cen A group galaxies are shown by solid circles. The correlation between µe and log(MHI /LR), shown as a solid line, has a slope of 1.51 ± 0.16 and a scatter of 0.9 magnitudes. As the survey detects galaxies by their Hi content, galaxies with low Hi mass-to-light ratios may be missed, however this survey goes deep enough to see galaxies down to MHI /LR ≃ 0.05M⊙/L⊙ so it is unlikely that this will be a significant effect. it cannot make any correction for the large scale structure, which could lead to distortions in the results. However the value of hV /Vmaxi implies that the overall effect of large scale structure on the sample is probably not large. 3.2 Hi Mass Function Weighting Our second method is to use an Hi Mass Function (HIMF) to correct for both ease-of-detection and large scale struc- ture, as described in Minchin (1999). This method makes the assumption that our galaxies are drawn from the same population as a general HIMF, such as that found for the HIPASS Bright Galaxy Catalogue (BGC) by Zwaan et al. (2003). Then the volumetric correction that needs to be ap- plied in order to match the distribution of Hi masses in HIDEEP with the HIMF can be calculated and applied to find other quantities, such as the luminosity function and the surface-brightness distribution. After dwarf galaxies are excluded, this gives us a sample of 101 galaxies. This HIMF-weighting cannot be used to construct an Hi mass function for the HIDEEP galaxies, as it would clearly just give the same answer as the input HIMF. However, we can use it to construct the bivariate brightness distribution, luminosity function and surface-brightness distribution of galaxies and to investigate how various parameters (e.g. lu- minosity density) change with surface-brightness. It would be possible to make a HIMF from the HIDEEP data using 1/Vmax, however we have chosen to use the HIMF of Zwaan et al. (2003), based on the 1000 galaxies in the HIPASS BGC. This contains an order of magnitude more galaxies Figure 5. The surface-brightness distribution formed by correct- ing with various different Hi Mass Functions. Solid line: HIPASS Bright Galaxy Catalogue (BGC; Zwaan et al. 2003) as used else- where in this paper; dotted line: Rosenberg & Schneider (2002); short-dashed line: Zwaan et al. (1997); long-dashed line: Kilborn (2001); dot-dashed line: Henning et al. (2000). Error bars are given for the distribution formed using the BGC HIMF. than HIDEEP and thus gives a much more accurate mass function. HIMF-weighting contains more sources of error than 1/Vmax. The main sources of additional errors are the width of the bins used in the HIMF and the numbers of galaxies in each bin. The resulting formal uncertainties are therefore normally larger than for 1/Vmax, despite the larger sam- ple size. However, it should be remembered that HIMF- weighting includes a correction for the large scale structure which is not factored into either the values or the errors of the 1/Vmax weighting. Systematic errors may be introduced by the choice of the HIMF. In order to gauge the size of these errors, we have looked at the contribution of LSB galaxies to the total number density of gas-rich galaxies when calculated with the HIMFs of Rosenberg & Schneider (2002), Kilborn (2001), Henning et al. (2000) and Zwaan et al. (1997) (see Table 3 for a summary). The resulting surface-brightness distributions are shown in Figure 5. It can be seen from Figure 5 that there is little dif- ference in the overall shape of the Surface Brightness Dis- tributions derived using the different HIMFs, although the overall density of gas-rich galaxies is significantly reduced if the HIMF of Zwaan et al. (1997) is used. The contribution of LSB galaxies as a proportion of all galaxies is therefore only weakly dependant on the HIMF chosen, with a slight rise in the proportion for steeper slopes. Within the range of HIMFs here the proportion of LSB galaxies varies between 59 and 67 per cent of all gas-rich galaxies, with the BGC HIMF giving 62 per cent. As the statistical error on this proportion is 37 per cent, it can be seen that the systematic error due to the HIMF is comparatively small. These sys- tematic errors are stated in Table 4 as a second error column to the percentage LSB contribution. Masters, Haynes, & Giovanelli (2004) have shown that if a peculiar velocity field model is used to assign distances to galaxies in the BGC rather than the pure Hubble law used by Zwaan et al. (2003) then the faint-end slope of the BGC HIMF steepens from α = −1.3 to −1.4, which is consistent with the steeper determinations. The HIMF of Zwaan et al. (1997) is also inconsistent with the other determinations; this cannot be explained by the method used for assigning distances but may be due to problems with correctly deter- mining the survey sensitivity (Schneider, Spitzak & Rosen- berg 1998). It is most likely, therefore, that if a systematic error has been introduced by the use of the HIMF, it is that the faint-end slope we have used is too shallow. The result of this would be for this paper to slightly underestimate the contribution of LSB galaxies. 3.3 The Bivariate Brightness Distribution As luminosity and surface-brightness appear to be correlated to some degree, the distributions of luminosity and surface- brightness on their own are less interesting than the bivari- ate brightness distribution (BBD) -- the joint distribution in the (MR,µR e ) plane. The BBD describes the contribution of galaxies of different luminosities and surface-brightnesses to the cosmos but is virtually impossible to obtain accu- rately from an optically-selected sample (Boyce & Phillipps 1995). For instance the largest sample of galaxies, complete in both surface brightness and luminosity, that can be as- sembled from the classical optical catalogues numbers less than 50 (Disney 1999). Figure 6 shows the BBDs drawn from the uncor- rected samples, showing the correlation between luminos- ity and surface-brightness and illustrating the conventional emphasis on the overwhelming importance of high-surface- brightness giant galaxies. However these galaxies tend to have much larger Hi masses and can be seen over much greater volumes. The corrected BBDs, showing the true space-density of galaxies, are given in Figure 7. The HIMF and 1/Vmax corrections give very similar results. Over a range of seven magnitudes in luminosity and five in surface-brightness, the galaxies appear to lie in an evenly populated strip four mag- nitudes wide in luminosity and two in surface-brightness that stretches from high-luminosity, high-surface-brightness to low-luminosity, low-surface-brightness. The HIMF-weighted BBD is obviously dependent on the HIMF used to construct it. However all of the HIMFs in Table 3 give results fairly similar to those shown here -- if a steeper HIMF is used then the density at low-luminosity, low surface-brightness is marginally higher while if the shallower Cosmological Significance of LSB galaxies 7 Figure 6. Uncorrected bivariate brightness distribution of HIDEEP galaxies. The upper panel shows the 101 galaxies in the HIMF-weighted sample, the lower panel the 62 galaxies in the 1/Vmax-weighted sample. The blank areas around the images indicate where there is no data. (and lower-density) HIMF of Zwaan et al. (1997) is used then the density decreases all round, with the density of low- luminosity, low surface-brightness galaxies falling slightly more than that of high-luminosity, high surface-brightness galaxies. We define LSB galaxies to be those with effective R- band surface-brightnesses more than 1.5 magnitudes lower than the peak in the uncorrected R-band effective surface- brightness distribution of ESO-LV galaxies, i.e. µe(R) > 23.3 mag arcsec−2. This is slightly dimmer than the defi- nitions used by Impey & Bothun (1997) (µ0(B) > 23) and by McGaugh (1996) (µ0(B) > 22.75), which are respectively 1.35 and 1.1 magnitudes dimmer than the peak in the uncor- rected B-band central surface-brightness found by Freeman (1970). Using our definition, we can set limits on the popu- lation of giant LSB galaxies as follows. There is a clear deficiency of high-luminosity, LSB 'Crouching Giant' galaxies (LSB galaxies with LR > 1010L⊙): of the 47 galaxies with LR > 1010L⊙, not one is an LSB galaxy. By applying the binomial theorem, we can calculate that the probability of finding no LSB galaxies in a sample of 47 is less than 0.05 if LSB galaxies make up more than 6 per cent of the population. We can therefore rule out that LSB galaxies make up more than 6 per cent 8 R. F. Minchin et al. Figure 7. The bivariate brightness distribution (BBD) of Hi-selected galaxies. The panels to the left show the BBD formed using HIMF-weighting, as described above, while the panels to the right show the BBD formed using 1/Vmax weighting. The central panels give the best-estimate of the BBD while the top panels show the BBD + 1σ in each bin and the bottom panels show the BBD - 1σ in each bin. of the high-luminosity, gas-rich population with 95 per cent confidence. Hi massive galaxies which are LSB galaxies of 26 per cent at the 95 per cent confidence level. Of the sixteen Hi-massive (MHI > 1010M⊙) galaxies found in HIDEEP, one (ESO 383-G059), is an LSB galaxy with µR e = 23.8 Rµ -- this galaxy has a high Hi mass despite not having a high optical (R-band) luminosity. By the same method as above, we can place a limit on the proportion of There is little evidence that the unpopulated regions at high-luminosity, low surface-brightness and low-luminosity, high surface-brightness are due to the Hi mass limits of the survey (imposed by the down-turn in the Hi mass function at high MHI and by small volumes, and the cut at 108M⊙, at low MHI ). Only one of the galaxies with MHI > 1010M⊙ is Cosmological Significance of LSB galaxies 9 Figure 8. Weighted luminosity function of HIDEEP galaxies. Points weighted by the HIMF are shown as open circles (offset for clarity by 0.1 mag dimmer), points weighted by 1/Vmax are shown as solid triangles (offset for clarity by 0.1 mag brighter). The LF of Blanton et al. (2001, α = −1.20 ± 0.03) is shown as a dashed line. Figure 9. Corrected surface-brightness distribution of HIDEEP galaxies. This is consistent with a flat or slowly rising surface- brightness distribution at lower surface-brightnesses and a down- turn at the bright end. The HIMF-weighted points are indicated by open circles (offset for clarity by 0.1 mag dimmer) and the 1/Vmax-weighted points by solid triangles (offset for clarity by 0.1 mag brighter). a low surface-brightness galaxy, as previously noted, and the high-mass galaxies are not spread along the high-luminosity edge of the populated region as would be expected if this edge were due to an Hi-mass selection effect. Similarly, only one of the galaxies in the lowest mass range included in the sample (108.5M⊙ > MHI > 108M⊙) is a high surface- brightness galaxy (NGC 5253); these galaxies are not spread along the low-luminosity edge of the populated region. The dwarf galaxies that were removed from the sample (MHI < 108M⊙) fall in the surface-brightness range 22.5 -- 24.5 R mag arcsec−2 and in the R magnitude range -13 -- -18, generally at a lower luminosity than is 'normal' for their surface-brightness. All but one of these galaxies (all in the 1/Vmax sample) lie in the Centaurus A group and are not, therefore, representative of a variety of environments -- the one dwarf that is not in the Cen A group (HIDEEP J1337-3118) lies well within the normal range of luminos- ity for its surface-brightness, while the higher-mass galaxy NGC 5253, which also lies in the Cen A group, is the galaxy most under-luminous for its surface-brightness in the sam- ple. Including these dwarf galaxies in the BBD would not populate in the low-luminosity, high surface-brightness re- gion, but would extend the BBD to lower luminosities at intermediate and low surface-brightnesses. By collapsing the BBD along the surface-brightness axis, we can form the optical Luminosity Function (LF) which is shown in Figure 8. This is compared to the LF obtained by Blanton et al. (2001) from Sloan Digital Sky Survey (SDSS) commissioning data. It can be seen that both the HIMF and 1/Vmax determinations are fairly consistent with the SDSS luminosity function, although M⋆ appears to be marginally brighter in our data - this could be due to uncertainties in the conversion between r⋆, which Blanton et al. use, and R. The 1/Vmax weighting gives a faint-end (MR > −21.5) slope of α = −1.40 ± 0.14 and the HIMF- weighting gives α = −1.27 ± 0.16. Figure 9 displays the surface-brightness distribution (SBD), formed by collapsing the BBD along the luminosity axis. This is consistent with optical determinations such as Davies (1990) and de Jong (1996a). The effect of large scale structure on 1/Vmax can be seen in the brightest bin. Both of the galaxies in this bin lie in the close by Cen A group and their contribution to the density is therefore over-estimated by this method. However, in general the 1/Vmax and HIMF- weighted points are in good agreement, giving confidence that large scale structure is not overly affecting the 1/Vmax data. Overall, it can be seen that optical surveys give very similar results to this survey. This implies that 21-cm sur- veys do not uncover any 'hidden' population of extremely low surface-brightness galaxies that is missed by optical sur- veys. If such a population does exist in significant numbers, it must be composed primarily of galaxies with neutral gas masses lower than 108M⊙. This is not to say, however, that LSB galaxies do not make up a significant population. The SBD derived from weighting by the BGC HIMF implies that LSB galaxies make up 62 ± 37 per cent of the total population of galaxies with MHI > 108M⊙, or 51 ± 20 per cent for 1/Vmax weight- ing. Even with the large errors on these estimates, it is clear that a large number of galaxies fall into our definition of low surface-brightness. In the next section we investigate what contribution these LSB galaxies make to the Universe. 10 R. F. Minchin et al. Figure 10. Luminosity density -- surface-brightness distribution for HIDEEP galaxies. The luminosity density can be seen to fall sharply either side of a peak near the Freeman-law value. The HIMF-weighted points are indicated by open circles (offset for clarity by 0.1 mag dimmer) and the 1/Vmax-weighted points by solid triangles (offset for clarity by 0.1 mag brighter). 4 THE COSMOLOGICAL SIGNIFICANCE OF LOW SURFACE-BRIGHTNESS GALAXIES As LSB galaxies have been proposed as repositories for some of the missing baryons (e.g. Impey & Bothun 1997) and may also contain large quantities of dark matter, it is important to make an estimate of how much they actually contribute to the Universe. It is possible to do this from our data, subject to the provisos that Hi-poor galaxies would not be found at 21-cm (e.g. elliptical galaxies) and that we do not detect sufficient numbers of dwarf galaxies (MHI < 108M⊙) to say anything about their contribution. We compare the contri- bution of the LSB galaxies in our sample to the total contri- bution of Hi-rich galaxies using the weightings described in Section 3 to correct the numbers in each surface-brightness bin. To compare the contribution to the luminosity-density made by galaxies of different surface-brightnesses we need to weight the surface-brightness distribution in Figure 9 with the luminosities of the galaxies. When this is done we obtain Figure 10 which shows that the luminosity density is sharply peaked near the Freeman-law value. Gas-rich LSB galaxies do not appear to emit much light: when analysed with the HIMF-weighting, they contribute 6.7 ± 2.8 per cent of the total luminosity-density of all gas-rich galaxies, or 6.5 ± 2.4 per cent according to the 1/Vmax analysis. This is very sim- ilar to the 7.3 ± 3.6 per cent contribution for LSB galaxies found by Driver (1999) for local (0.3 < z < 0.5) galaxies in the Hubble Deep Field. The main source of errors in our determination is Poisson noise in the surface-brightness bins but there is also a contribution due to the uncertainty on the luminosity of the sources and, for the Hi mass function Figure 11. Neutral hydrogen density -- surface-brightness dis- tribution for HIDEEP galaxies. It can be seen that this is con- sistent with a slowly falling distribution towards lower surface- brightnesses and a much steeper fall-off towards higher surface- brightnesses, similar in shape to the surface-brightness distribu- tion. The HIMF-weighted points are indicated by open circles (offset for clarity by 0.1 mag dimmer) and the 1/Vmax-weighted points by solid triangles (offset for clarity by 0.1 mag brighter). weighting, due to the width of the mass bins and the number of galaxies in each bin. Similarly the contribution of LSB galaxies to the Hi content as a whole can be calculated and is shown in Figure 11. This amounts to 32 ± 11 per cent of all Hi with HIMF- weighting, or 27 ± 7 per cent with 1/Vmax weighting. This is inconsistent, at the 1σ level, with the determi- nation of Zwaan et al. (2003) that LSB galaxies contribute only 15 per cent to the total Hi mass density. That deter- mination should, however, be treated as a lower limit. The sample of Zwaan et al., unlike our sample, did not have com- plete optical data -- surface-brightnesses were only available for 600 out of the 1000 galaxies. When corrections for this were made, Zwaan et al. assumed that the incompleteness in the optical data was unrelated to surface-brightness. It is far more likely that the LSB galaxies' data were more incomplete than the HSB galaxies', which could raise the contribution of LSB galaxies considerably. The baryonic content of the galaxies was calculated fol- lowing the method of McGaugh et al. (2000), adding the mass of the stars and the gas together to get a total bary- onic mass for the galaxy: Mbary = 1.4MHI + ΥX ⋆ LX (1) where ΥX ⋆ is the stellar mass to light ratio in the band used, and LX is the luminosity in that band. For our R-band data, ΥR ⋆ has been estimated, as in McGaugh et al., using the model of de Jong (1996b) for a 12 Gyr old, solar metallicity stellar population with a constant star formation rate and a Salpeter initial mass function, corrected to R-band using Cosmological Significance of LSB galaxies 11 Figure 12. Baryon density -- surface-brightness distribution for HIDEEP galaxies. It can be seen that the greatest contribution to the baryon density is made by Freeman-law galaxies around µ⋆ e , with the density falling of towards lower and higher surface- brightnesses. The HIMF-weighted points are indicated by open circles (offset for clarity by 0.1 mag dimmer) and the 1/Vmax- weighted points by solid triangles (offset for clarity by 0.1 mag brighter). Figure 13. Mass density -- surface-brightness distribution for HIDEEP galaxies. The greatest contribution is made by Freeman- law galaxies µ⋆ e , however the distribution seems fairly flat towards lower surface-brightnesses while it falls off towards higher surface- brightnesses. The HIMF-weighted points are indicated by open circles (offset for clarity by 0.1 mag dimmer) and the 1/Vmax- weighted points by solid triangles (offset for clarity by 0.1 mag brighter). the average colours in that paper (also used by McGaugh et al. for their correction to H-band). This gives a value of ΥR ⋆ ≈ 1.4, which we have used in our calculations. The relative contribution of LSB galaxies to the total baryon density is then calculated to be 9.3± 3.6 per cent us- ing HIMF-weighting or 8.7±2.9 per cent using 1/Vmax. This is only marginally higher than the contribution to the lumi- nosity, reflecting that the contribution to Hi mass density is basically flat and so only slightly affects the shape of the density distribution when the two are added together. LSB galaxies do have more of their baryons in the form of gas, as shown in the relationship between MHI /L and surface- brightness, but this is outweighed by their much lower lumi- nosities. If, in the usual way, we estimate the dynamical masses of HIDEEP galaxies using Mdyn = RHI × (∆V0)2 G (2) where we follow Paper 1 in assuming rHI ≃ 5re and ∆V0 is the inclination-corrected velocity width (∆V0 = ∆V / sin i) then we arrive at the relative contribution of galaxies of var- ious surface-brightnesses shown in Figure 13. For this calcu- lation, only those galaxies with reliable inclinations in the range 45◦ < i < 80◦ were used. This leaves 57 galaxies in the HIMF-weighted sample and 38 in the 1/Vmax-weighted sample. The share due to LSB galaxies is, although uncer- tain, relatively high at 22± 10 per cent for HIMF-weighting or 21 ± 12 per cent for 1/Vmax-weighting. This is in keeping with other studies that have shown that LSB galaxies con- tain proportionally more dark matter than normal galaxies. Since the cross-sections of large, luminous galaxies can by no means explain the prevalence of quasar absorption line systems (QSOALS) it has been suggested that LSB galax- ies might be responsible (e.g. Phillipps et al. 1993, Linder 1998; 2000) though this has proved controversial (e.g. Chen et al. 1998). If we assume that the absorption cross-sections of our galaxies are proportional to their effective areas (e.g. the area enclosed by the effective radius) then we can make an estimate of the contribution of LSB galaxies to this cross- section. Figure 14 shows the cross-sectional area distribution of our galaxies with surface-brightness (corrected to an in- clination of 60◦), formed in the same way as the above plots This gives the contribution of LSB galaxies to the cross- section to be 39 ± 15 per cent using the HIMF-weighting and 42 ± 22 per cent using 1/Vmax-weighting. Their high cross-section suggests that LSB galaxies are likely to form a significant fraction of the absorbers where galaxies themselves are responsible for QSOALS, as they might be in the case of damped Lyman-α systems. This is born out by recent observations where LSB galaxies, rather than HSB galaxies with immense halos, appear to be the more likely absorbers (Turnshek et al. 2001; Bowen, Tripp & Jenkins 2001). Table 4 summarises the findings of this section. It can be seen that LSB galaxies make up over half of all gas-rich galaxies, yet have less than 10 per cent of the luminosity den- sity. The relationship between dynamical M/L and surface- brightness means that luminosity is a biased indicator of the cosmological significance of LSB galaxies. Similarly, the higher Hi mass-to-light ratios of LSB galaxies means that 12 R. F. Minchin et al. Table 4. Summary of the measured contribution of HSB and LSB galaxies to the density of various quantities in gas-rich galaxies. These are the actual measured densities in the HIDEEP survey, for the SBDs constructed using weighting by the HIMF of Zwaan et al. (2003) and using 1/Vmax weighting. Possibly systematic errors due to the selection of the HIMF, calculated by comparing the results of the Zwaan et al. (2003) HIMF with other published HIMFs as described earlier, are given as a second error to the percentage LSB contribution. Quantity Weighting Ngal HSB contribution LSB contribution Percentage LSB contribution HIMF 1/Vmax Number density (10−2 galaxies Mpc−3) Neutral Hydrogen density HIMF (107M⊙ Mpc−3) Luminosity density (107L⊙ Mpc−3) Baryon density (107M⊙ Mpc−3) Mass density (107M⊙ Mpc−3) Area density (10−1 kpc2 Mpc−3) 1/Vmax HIMF 1/Vmax HIMF 1/Vmax HIMF 1/Vmax HIMF 1/Vmax 101 62 101 62 101 62 101 62 57 38 101 62 1.8 ± 0.7 2.9 ± 1.1 3.7 ± 0.7 5.7 ± 1.2 45 ± 13 57 ± 13 68 ± 17 88 ± 20 280 ± 80 570 ± 170 5.4 ± 1.2 7.4 ± 1.7 3.0 ± 1.5 3.0 ± 0.9 1.8 ± 0.5 2.1 ± 0.5 3.2 ± 1.0 3.9 ± 1.2 6.9 ± 2.1 8 ± 2 77 ± 31 150 ± 78 3.4 ± 1.2 5.4 ± 2.5 62 ± 37 +5 −3 51 ± 20 32 ± 11 +10 −3 27 ± 7 7 ± 3 +1 −1 7 ± 2 9 ± 4 +2 −1 9 ± 3 22 ± 10 +4 −2 21 ± 12 39 ± 15 +8 −3 42 ± 22 than this, while elliptical galaxies are too gas-poor to be de- tected. Most dwarf galaxies have low surface-brightnesses, therefore if these were included it is likely that the total contributions from LSB galaxies would rise. These numbers should therefore be seen as lower estimates for the total con- tribution of LSB galaxies. 5 CONCLUSIONS This survey does not find any LSB 'Crouching Giant' galax- ies, such as Malin 1. From this, we can rule out LSB galaxies making up more that 6 per cent of the population of lumi- nous (LR > 1010L⊙) galaxies. Indeed, there does seem to be a minimum surface-brightness at every magnitude level given approximately by µe,min = 45 + MR (Figure 7). To higher surface-brightnesses, galaxies seem to populate fairly evenly a band approximately 4 magnitudes wide in luminos- ity and 2 in surface-brightness. Beyond this, high surface- brightness, low-luminosity galaxies also appear to be rare in the gas-rich population. The populated band broadens slightly towards lower luminosities, giving it an approximate slope of Σ ∝ L0.7 (where Σ is surface brightness in luminos- ity per unit area and L is the total luminosity). From this, it can be calculated that the radius, R, is related to the surface-brightness as R ∝ Σ0.2, i.e. the radius only changes very slightly with surface-brightness. Once volumetric corrections are made, the number of galaxies per unit volume is flat as we go to lower surface-brightnesses. Even within the surface-brightness limit reached by this survey, LSB galaxies contribute over half of the number density of galaxies. Furthermore, we find that LSB galaxies contribute approximately 30 per cent of the neutral-hydrogen density, twice the lower limit set by Zwaan et al. (2003). LSB galaxies may also contribute around 20 per cent of the total mass density (again with large errors), but only 7 per cent of the luminosity-density. Overall, Hi-rich LSB galaxies do not appear to con- tribute much to the Universe. However, the cross-section to QSOALS made by LSB galaxies, which is around 40 per cent of the total cross-section, is disproportionate to their luminosity or baryonic content. This implies that LSB galax- Figure 14. Cross-sectional area density against surface- brightness for HIDEEP galaxies. This is fairly flat, with a fall off towards higher surface-brightnesses (the brightest point of the 1/Vmax distribution is, as noted earlier, affected by large scale structure). This implies that LSB galaxies will make a significant contribution to QSO absorption lines in so far as these are caused by galaxies. The HIMF-weighted points are indicated by open circles (offset for clarity by 0.1 mag dimmer) and the 1/Vmax- weighted points by solid triangles (offset for clarity by 0.1 mag brighter). they have more gas than would be indicated by their light on a straight extrapolation of MHI /L from Freeman-law galaxies. The relatively larger sizes of LSB galaxies means that their contribution to cross-sectional area, around 40 per cent, is much higher than would be expected from their luminosity or even from their Hi mass. The totals given here are only for galaxies with MHI > 108M⊙. These are almost entirely spiral galaxies. Dwarf galaxies, even gas-rich ones, tend to have lower gas masses ies could contribute substantially to those QSOALS where galaxies are the absorbers. Future work in progress will see an order of magnitude larger sample assembled from the overlap between HIPASS and the SDSS. This will give not only an increase in the significance of the results but will also give 5-colour op- tical data, thus greatly extending our knowledge of Hi- selected galaxies. A larger sample of Hi-selected samples from the SDSS may be assembled in the medium term by the proposed ALFALFA drift-scan survey with the Arecibo L-band Feed Array or in the longer term by surveys carried out with the Square Kilometre Array. However, as neither HIPASS nor ALFALFA reach column-density levels as low as HIDEEP it is unlikely that they will turn up any sig- nificant population of extremely LSB galaxies that we have missed. This is not to say that extremely LSB galaxies do not exist, just that if they do make up a significant population then the amount of Hi they contain must be small. In order to search for these galaxies, techniques other than Hi sur- veys will be necessary. The optical channel available to us on the ground is fundamentally unsuited to looking for LSB galaxies, but a quite modest switch in wavelengths to either side of the solar peak, which is possible from space, could yield dramatic results. O'Connell (1987) has discussed the idea of using UV wavelengths, while in the H-band at 1.6 µm the contrast between the Sun's scattered zodiacal spectrum and the light from red stars in LSB galaxies could be two or- ders of magnitude higher than the contrast from the ground. Wide-field near-IR cameras, such as the WFC-3 instrument currently awaiting shipment to the HST, will be needed to exploit this window which is, in principle, capable of reveal- ing galaxies 7 magnitudes lower in surface-brightness than we can detect from the ground. The authors would like to thank Lister Staveley-Smith, Tony Fairall, David Barnes, Jon Davies and Suzanne Linder for useful discussions. RFM, WJGdeB and PJB acknowledge the support of PPARC. BGK acknowledges the support of the Australian Research Council. The authors would also like to thank the staff of the CSIRO Parkes and Narrabri observatories, of the NRAO Ar- ray Operations Center and of the ANU Siding Spring Observatory for their help with observations. We ac- knowledge PPARC grant GR/K/28237 to MJD towards the construction of the multibeam system and PPARC grants PPA/G/S/1998/00620, PPA/G/S/2002/00053 and PPA/G/S/2003/00085 to MJD towards its operation. We would particularly like to thank the anonymous referee for an extremely helpful report. PPA/G/S/1998/00543, This research has made use of the NASA/IPAC Ex- tragalactic Database (NED) which is operated by the Jet Propulsion Laboratory, Caltech, under agreement with the National Aeronautics and Space Administration. This re- search has also made use of the Digitised Sky Survey, pro- duced at the Space Telescope Science Institute under US Government Grant NAG W-2166 and of NASA's Astro- physics Data System Bibliographic Services REFERENCES Bertins E., Arnouts S., 1996, A&AS, 117, 393 Cosmological Significance of LSB galaxies 13 Bland-Hawthorn J., Shopbell P. L., Malin D. F., 1993, AJ, 106, 2154 Blanton M. R. et al. 2001, AJ, 121, 2358 Bowen D. V., Tripp T. M., Jenkins E. B., 2001, AJ, 121, 1456 Boyce P. J., Phillipps S., 1995,A&A, 296, 26 Briggs F. H., Rao S., 1993, ApJ, 417, 494 Chen H.-W., Lanzetta K. M., Webb J. K., Barcons X., 1998, ApJ, 498, 77 Churchill C. W., Le Brun V., 1998, ApJ, 499, 677 Davies J. I., 1990, MNRAS, 244, 8 de Blok W. J. G., McGaugh S. S., van der Hulst J. M., 1996, MNRAS, 283, 18 de Jong R. S., 1996a, A&A, 313, 45 de Jong R. S., 1996b, A&A, 313, 377 Disney M. J., 1976, Nature, 263, 573 Disney M. J., 1999, in Davies J. I., Impey C., Phillipps S., eds, ASP Conf. Ser. 170, The Low Surface Brightness Universe, ASP, San Francisco, p. 9 Disney M. J., Phillipps, S., 1983, MNRAS, 205, 1253 Driver S. P., 1999, ApJ, 576, L69 Freeman K. C., 1970, AJ, 160, 811 Fukugita M., Hogan C. J., Peebles P. J. E., 1998, ApJ, 503, 518 Graham A. W., de Blok W. J. G., 2001, ApJ, 556, 177 Hambly N. C., MacGillivray H. T., Read M. A., Tritton S. B., Thomson E. B., Kelly B. D., Morgan D. H., Smith R. E., Driver S. P., Williamson J., Parker Q. A., Hawkins M. R. S., Williams P. M., Lawrence A., 2001, MNRAS, 326, 1279 Henning P. A., Staveley-Smith L., Ekers R. D., Green A. J., Haynes R. F., Juraszek S., Kesteven M. J., Koribalski B., Kraan-Korteweg R. C., Price R. M., Sadler E. M., Schroder A., 2000, AJ, 119, 2686 Impey C. D., Bothun G. D., 1997, ARA&A, 35, 367 Kilborn V. A., 2001, PhD Thesis, University of Melbourne Koribalski, B. S. et al., 2004, AJ, 128, 16 Kron R. G., 1980, ApJS, 43, 305 Lauberts A., Valentijn E.A., 1989, The Surface Photometry Cat- alogue of the ESO-Uppsala Galaxies, ESO, Garching Linder S. M., 1998, ApJ, 495, 637 Linder S. M., 2000, ApJ, 529, 644 McGaugh S. S., 1996, MNRAS, 280, 337 McGaugh S. S., Schombert J. M., Bothun G. D., de Blok W. J. G., 2000, ApJ, 533, 99 Maddox S. J., Efstathiou G., Sutherland W. J., 1990, MNRAS, 246, 433 Masters K .L., Haynes M. P., Giovanelli, R., 2004, ApJL, accepted Minchin R. F., 1999, PASA, 16, 12 Minchin R. F. et al. 2003, MNRAS, 346, 787 (Paper 1) O'Connell R., 1987, AJ, 94, 867 Parker Q. A., Malin D. F., 1999, PASA, 16, 288 Phillipps S., Disney M. J., Davies J. I. 1993, MNRAS, 260, 453 Rosenberg J. L., Schneider S. E., 2002, ApJ, 567, 247 Schlegel D. J., Finkbeiner D. P., Davis M., 1998, ApJ, 500, 525 Schneider S. E., Spitzak J. G., Rosenberg J. L., 1998, ApJ, 507, 9 Thim F., Tammann G. A., Saha A., Dolphin A., Sandage A., Tolstoy E., Labhardt L., 2003, ApJ, 590, 256 Turnshek D. A., Rao S., Nestor D., Lane W., Monier E., Bergeron J., Smette A., 2001, ApJ, 553, 288 Zwaan M. A., van der Hulst J. M., de Blok W. J. G., McGaugh S. S., 1995, MNRAS, 273, L35 Zwaan M. A., Briggs F. H., Sprayberry D., Sorar E., 1997, ApJ, 490, 173 Zwaan M. A. et al. 2003, MNRAS, 125, 2842
0804.1363
2
0804
2008-07-17T20:23:05
The Accuracy of Morphological Decomposition of Active Galactic Nucleus Host Galaxies
[ "astro-ph" ]
In order to assess the accuracy with which we can determine the morphologies of AGN host galaxies, we have simulated more than 50,000 ACS images of galaxies with z < 1.25, using image and noise properties appropriate for the GOODS survey. We test the effect of central point-source brightness on host galaxy parameter recovery with a set of simulated AGN host galaxies made by adding point sources to the centers of normal galaxies. We extend this analysis and also quantify the recovery of intrinsic morphological parameters of AGN host galaxies with a set of fully simulated inactive and AGN host galaxies. We can reliably separate good from poor fit results using a combination of reasonable error cuts, in the regime where L_{host}:L_{PS} > 1:4. We give quantitative estimates of parameter errors as a function of host-to-point-source ratio. In general, we separate host and point-source magnitudes reliably at all redshifts; point sources are well recovered more than 90% of the time, although spurious detection of central point sources can be as high as 25% for bulge-dominated sources. We find a general correlation between Sersic index and intrinsic bulge-to-total ratio, such that a host galaxy with Sersic n < 1.5 generally has at least 80% of its light from a disk component. Likewise, "bulge-dominated" galaxies with n > 4 typically derive at least 70% of their total host galaxy light from a bulge, but this number can be as low as 55%. Single-component Sersic fits to an AGN host galaxy are statistically very reliable to z < 1.25 (for ACS survey data like ours). In contrast, two-component fits involving separate bulge and disk components tend to over-estimate the bulge fraction by ~10%, with uncertainty of order 50%.
astro-ph
astro-ph
Accepted for publication in Ap. J. The Accuracy of Morphological Decomposition of AGN Host Galaxies B. D. Simmons and C. M. Urry Astronomy Department, Yale University, New Haven, CT USA [email protected] ABSTRACT In order to assess the accuracy with which we can determine the morpholo- gies of AGN host galaxies, we have simulated more than 50,000 ACS images of galaxies with z < 1.25, using image and noise properties appropriate for the GOODS survey. We test the effect of central point-source brightness on host galaxy parameter recovery with a set of simulated AGN host galaxies made by adding point sources to the centers of normal galaxies. We extend this analysis and also quantify the recovery of intrinsic morphological parameters of AGN host galaxies with a set of fully simulated inactive and AGN host galaxies. We can reliably separate good from poor fit results using a combination of reasonable error cuts, in the regime where (Lhost/LPS) > 1 : 4. We give quan- titative estimates of parameter errors as a function of host-to-point-source ratio. In general, we separate host and point-source magnitudes reliably at all redshifts; point sources are well recovered more than 90% of the time, although spurious detection of central point sources can be as high as 25% for bulge-dominated sources. We find a general correlation between S´ersic index and intrinsic bulge- to-total ratio, such that a host galaxy with S´ersic n < 1.5 generally has at least 80% of its light from a disk component. Likewise, "bulge-dominated" galaxies with n > 4 typically derive at least 70% of their total host galaxy light from a bulge, but this number can be as low as 55%. Single-component S´ersic fits to an AGN host galaxy are statistically very reliable to z < 1.25 (for ACS survey data like ours). In contrast, two-component fits involving separate bulge and disk components tend to over-estimate the bulge fraction by ∼ 10%, with uncertainty of order 50%. Subject headings: methods: data analysis -- galaxies: active -- galaxies: nuclei -- galaxies: bulges -- galaxies: fundamental parameters -- 2 -- 1. Introduction Morphological analysis of two-dimensional light profiles of galaxies in large data surveys provides detailed information about galaxy populations and their evolution (e.g., Simard et al. 1999; Ravindranath et al. 2004; Jogee et al. 2004; Sheth et al. 2008). Simulations of both large and small galaxy samples show that two-dimensional parametric and non-parametric morphological analysis of normal galaxies is extremely robust (Marleau & Simard 1998; Graham 2001; Conselice 2003; Trujillo & Aguerri 2004; Haussler et al. 2007). Studies of active galactic nucleus (AGN) host galaxies give us the opportunity to study not just the galaxies themselves, but also the well-established connection between galaxies and central supermassive black holes (Kormendy & Richstone 1995; Magorrian et al. 1998; Ferrarese & Merritt 2000; Gebhardt et al. 2000; McLure & Dunlop 2002; Marconi & Hunt 2003). In cases where central black hole masses are independently determined, accurate decompositions of AGN host galaxies from their central point-source light contributions allow for the direct study of how galaxy light distribution relates to black hole properties. Alternatively, when time-intensive observations of black hole mass are not available, we can use established bulge-black hole relations to determine the black hole masses from bulge luminosities. However, two-dimensional morphological fitting of AGN host galaxies is more compli- cated than that of "normal" (or "inactive") galaxies because of the presence of a central point source, which is often quite bright. In order to extract host galaxy properties from a source comprised of a host plus a central AGN, spatial resolution is critical. Thus the Advanced Camera for Surveys (ACS) on the Hubble Space Telescope (HST ) is the instrument of choice for many AGN host galaxy studies to date (S´anchez et al. 2004; Alonso-Herrero et al. 2008; Ballo et al. 2007). Of the large multi-wavelength surveys currently available, the Great Observatories Ori- gins Deep Survey (GOODS; Giavalisco et al. 2004) provides some of the deepest multicolor ACS data. AGN identification is possible because of deep X-ray imaging with Chandra, as well as ground-based optical and infrared spectroscopic follow up. Spitzer data provide additional information on total light but insufficient spatial resolution to separate the galaxy from the active nucleus. We performed detailed morphological analysis on the GOODS ACS data set using GALFIT (Peng et al. 2002), the results of which are presented in a forthcom- ing paper (B. D. Simmons et al. 2008, in preparation, hereafter S08). In order to understand the accuracy of those results -- in particular, to probe the well-known effects of surface bright- ness dimming and dependence of physical resolution on redshift in the presence of a central point source whose size does not change with redshift -- extensive simulations of host galaxy morphology are required. -- 3 -- Previously, S´anchez et al. (2004) simulated 1880 single-component host galaxies with point sources fainter than the host galaxy. Here we present a full treatment of over 50,000 simulated AGN host galaxies in the redshift range 0.1 < z < 1.1, with both single-component and two-component bulge-plus-disk morphologies, and central point sources that are both brighter and fainter than the host galaxy. Results of these simulations are intended to inform data analysis of AGN host morphologies, to better infer intrinsic host galaxy shapes from fitted morphological parameters in the presence of a central point source. We discuss the data from which we draw our noise properties and simulated sample in Section 2. The detailed fit procedures and the simulation parameter space are presented in Section 3, and we assess the ability of our fitting procedure to recover accurate host galaxy and point-source parameters from GOODS-like images of AGN host galaxies in Section 4. 2. Data 2.1. HST ACS Data The GOODS fields each cover an area of approximately 10′ × 16′ with a total of 398 HST orbits in both fields. The ACS has a resolution of 0′′.05 pixel−1, and observations were taken in the F435W, F606W, F775W, and F850LP ACS bands, hereafter referred to as B, V , I, and z ′, respectively. Data acquisition and reduction are detailed in Giavalisco et al. (2004) and Koekemoer et al. (2002). Each of the five epochs of data was processed via the basic ACS pipeline and a further processing task called multidrizzle, which finds an astrometric solution for all 5 epochs in order to correct for geometric distortion, and at the same time removes cosmic rays from the image (Koekemoer et al. 2002). The final images have resolution of 0′′.03 pixel−1, and the magnitude limits for extended sources are B < 28.4 mag, V < 28.4 mag, I < 27.7 mag, and z < 27.3 mag. From the reduced data, we randomly selected 450 inactive galaxies with stellarity class less than 0.8 (i.e., objects that are not point sources) and magnitude z ′ < 24.0 mag for use in those of our simulations that include real data (§3.3). Since the majority of AGN host galaxies selected for morphological analysis in current surveys using HST have redshifts such that their I-band data lie in the rest-frame B band (0.575 < z < 0.9), we performed simulations using the I-band images and noise properties. -- 4 -- 2.2. Noise Properties For fully simulated galaxies (described in §3.4), noise appropriate to the GOODS-ACS fields was added. The task of simulating the noise in the GOODS-ACS fields is complicated by the dithering process, which correlates the noise among nearby pixels. We therefore use actual noise pixels (i.e., source-free pixels) from the final GOODS-N and GOODS-S images rather than statistical models. We sampled 150 sub-sections of the GOODS images that contained no sources, 75 in the North field and 75 in the South, for a total of over 2.5 million noise pixels. Each of these sub-sections was cut into tiles of size 50 × 50 pixels; this resulted in 1042 tiles. These were then used to create 500 noise images of the same size as our data images by arranging random mosaics of the 50 × 50 noise tiles. We have verified that these new noise mosaics have the same overall noise properties (distribution of pixel intensities) as the original noise-only subsections of the GOODS images (Figure 1). We also tested noise mosaics created from 10×10 tiles of background noise but found that the distribution of pixel intensities was shifted compared to the correlated noise properties of the dithered ACS images. We thus used only the 50 × 50 pixel noise images to create our final noise mosaics. The noise value distributions of the 50 × 50 and 10 × 10 noise mosaics are compared to the actual GOODS noise distribution in Figure 1. 3. Creation and Morphological Fitting of Simulated Samples We performed two distinct sets of simulations to test the accuracy of derived morpho- logical parameters. First, we used a sample of 450 real galaxies from the GOODS fields and added point sources to their centers. This allowed us to test the ability of GALFIT to extract point sources from the centers of real galaxies. It also informed our choice of parameters and galaxy types for the second, more extensive, set of simulations. In the latter set of simulations, we created a sample of 12,592 completely simulated GOODS galaxies, in order to additionally test the recovery of a wide range of host galaxy parameters. We fitted all of our simulated galaxies with our batch-fitting algorithm (described below). We also simulated the redshifting of these galaxies to quantify redshift-dependent effects, and we added noise taken directly from real GOODS data. 3.1. Fit Procedure We performed morphological analysis in two dimensions using GALFIT, which can simultaneously fit an arbitrary number of components to an image (Peng et al. 2002). The -- 5 -- Fig. 1. -- Histograms of pixel intensities for GOODS noise (solid), noise mosaics using a 50 × 50-pixel sampling of the GOODS noise (dashed), and using a 10 × 10-pixel sampling of the GOODS noise (blue, dot-dashed). The histogram for the mosaic using 10 × 10-pixel tiles is significantly different from the intrinsic noise distribution. The distribution of pixel intensity of the 50 × 50 noise mosaic is nearly indistinguishable from that of the intrinsic GOODS noise distribution. -- 6 -- program uses a χ2 minimization method to determine the best-fit parameters. We used the S´ersic profile, which models the light distribution of a galaxy as an exponential function with a variable half-light radius, re, and an exponential parameter, n, called the S´ersic index (S´ersic 1968). When n is fixed at a value of 1, the S´ersic profile is equivalent to an exponential disk; when n = 4, the S´ersic profile is equivalent to a de Vaucouleurs bulge. Each of our simulated galaxies was fitted twice: once using a single S´ersic function with a variable index, and once with two S´ersic functions with fixed n = 4 and n = 1 (bulge + disk). We also simultaneously fitted point-source (PS) components using the point-spread function (PSF) of the data, determination of which is based on GOODS field stars and is described in S08. We developed a batch-fitting algorithm to fit AGN host galaxies in GALFIT (described in detail in S08), which we use here on our simulated galaxies. The batch-fitting algorithm uses a series of initial guesses to execute a first-pass fit of the central region of each AGN host galaxy, fitting the individual centroid positions of each component (S´ersic plus point source, or de Vaucouleurs bulge plus exponential disk plus point source). The second fit iteration zooms out to include the full extended galaxy and fits the central point-source magnitude and AGN host galaxy parameters (S´ersic index or bulge/disk decomposition, half-light radius, axial ratio, etc.). Finally, a third iteration is performed where all parameters are allowed to vary (except n in cases of a bulge + disk fit). It is important to note that, while these simulations include thousands of AGN host galaxies, even the largest (current) galaxy surveys using ACS data provide no more than hundreds of AGN host galaxies that can be reasonably fitted with GALFIT. We can therefore (as in S08) follow up the batch fits by hand, using the results of the batch-mode fitting as initial guesses to further constrain the results for each AGN host galaxy. Due to the size of the simulated galaxy sample in this paper, we did not do this final hand-fitting step. Thus our conclusions about the fraction of cases for which accurate morphological parameters are recovered should be considered conservative when compared to a sample of galaxies whose fits are adjusted and verified individually. 3.2. Determination of Initial Parameter Guesses For our real host galaxy sample, we use magnitude, flux radius, and position angle from the GOODS catalogs, which were created using SExtractor (Bertin & Arnouts 1996). Our simulated galaxies do not have a SExtractor catalog, but we simulated initial parameter guesses by introducing random errors into the true values. Specifically, when determining initial guesses for parameters that can be measured from -- 7 -- a data image, we assume that the guess value is accurate to within the following: ±(2.0, 2.0) pixels in position, ±0.5 in total magnitude, 10% in re and b/a, and 10◦ in position angle. We assumed no a priori knowledge of the S´ersic index n of each simulated galaxy, nor of whether each simulated galaxy was a two-component bulge plus disk system, or a system with either a pure bulge or disk. Our initial guess for the S´ersic index (or indices) of each fit is n = 2.5. In the case of a multi-component fit, we assume that the total magnitude is split evenly among the components (i.e., that we have no prior knowledge of host:PS contrast ratio or bulge-to-total ratio), and we additionally assume that the measured position angle and axial ratio are the average of the actual values for each separate component after adding the random fluctuation to the sum. 3.3. Real Galaxies with Added Point Sources In order to test whether we can recover central point sources from galaxies, we selected 450 I-band images of normal galaxies (i.e., not X-ray detected) from the GOODS-North and GOODS-South fields. The selection was performed so the sample has the same magnitude distribution as the full sample of galaxies in each of the GOODS fields, but was otherwise random. We assumed no knowledge of redshifts within the sample. We fitted each of these galaxies with a single S´ersic profile in order to determine the baseline set of morphological parameters for each galaxy. These fit parameters (S´ersic n, magnitude, half-light radius re, etc.) were taken to be the "actual" galaxy parameters. We then made nine copies of each galaxy, adding one point source to each, centered on the central pixel of the galaxy. The added point sources ranged in I-band magnitude from 21 to 29, in increments of 1 mag. This gave us a total of 4050 simulated AGN from the original 450 galaxies, with a large range of host galaxy-to-point-source contrast ratios, from ∼1000:1 to 1:100. These provide a straightforward means of assessing the effect of a central point source on the recovered parameters of a real galaxy. Our convergence rate for the fitting routine depends on the magnitude of the added central point source. Of the models with IP S = 21, 88% converge, whereas 64% of models with Ips = 29 (significantly below the flux limit of our I-band sample) converged. This is due to the fitting program being unable to converge to a value for the central point-source magnitude. Examples of fit results for typical galaxies in our sample are shown in Figure 2. -- 8 -- Fig. 2. -- Example fit results for two AGN hosts simulated from real galaxies. For each galaxy, the top row shows the original GOODS galaxy (from left to right: galaxy, fit, residual). Each row below shows the host+fit+residual for added central point sources of magnitudes IAB = 22, 24, 26, and 28, with the faintest point source in the bottom row. Note: These cutouts are zoomed-in to show the galaxies; the actual fitting regions are significantly larger. -- 9 -- 3.4. Simulated Galaxies The simulated hosts described in §3.3 provide information on how the presence of a central point source changes the fitted host galaxy parameters, but they do not allow us to explore redshift effects or quantify the absolute accuracy of the fit parameters. This can only be accomplished with a set of simulations for which we know the intrinsic properties of each simulated galaxy. To that end, we simulated three kinds of galaxies with a range of parameters: pure de Vaucouleurs bulge (B) galaxies with fixed S´ersic n = 4, pure exponential disk (D) galaxies with n = 1, and composite galaxies with both bulge and disk (B+D) components in varying proportions. The galaxy parameters were chosen to be typical of bright, local galaxies (Binney & Merrifield 1998) and placed at z = 0.125. Redshifted samples were developed from these, as described below. For the B and D type galaxies, the component parameters occupy a grid of four values for each parameter across the ranges 16.9 ≤ IAB ≤ 20.4 and 1.5 ≤ re ≤ 6.0 kpc (for bulges); 4.0 ≤ re ≤ 10.0 kpc (for disks), with axis ratios 0.25 ≤ b/a ≤ 1.0. For the single-component fits, the position angle was fixed at 45.0◦. This results in 64 B and 64 D galaxies, each of which is "inactive", i.e., without a central point source. We create simulated AGN host galaxies by adding point sources to these. We used five values between 16.5 ≤ IAB ≤ 24 for the central point source, resulting in 320 B-type and 320 D-type AGN host galaxies, with contrast ratios ranging from −1.3 ≤ log (Lhost/LPS) ≤ 2.8. Lastly, in order to ensure enough galaxies for statistical analysis after binning the fitted single-component sample, we made four copies of each B and D inactive galaxy and two copies of each B and D AGN host galaxy. This resulted in a total of 256 B and 256 D inactive single-component galaxies, and 640 B and 640 D AGN host galaxies. For the B+D type galaxies, the parameter space includes more magnitudes so as to pro- vide 20 input bulge-to-total ratios with 0.028 ≤ (B/Tot) ≤ 0.97 [plus the single-component B and D fits, which have (B/Tot) = 1 and 0, respectively]. In addition, the position angles of the individual components were allowed to vary such that some of the simulated B+D galaxies have components that are slightly (≤ 15◦) off-axis with respect to each other. The axis ratios between bulges and disks were also allowed to vary with respect to each other. These changes significantly increase the total number of double-component galaxies created; thus the number of radius, b/a, and point-source magnitude parameters used was decreased in order to keep the total number of galaxies from being prohibitively large from a computa- tional perspective. This resulted in the creation of 2700 inactive double-component (B+D) galaxies, and 8100 B+D AGN host galaxies. Table 1 gives the parameter values used to create the entire local suite of 12,592 simulated single- and double-component galaxies. -- 10 -- Since we are also interested in distinguishing the effects of redshifting galaxies from the effects of evolving galaxies, the initial sample was defined to lie at z = 0.125, and three additional samples were placed at z = 0.413, 0.738, and 1.075, corresponding to the redshifts at which the centers of each of our GOODS filters (F435W, F606W, F775W, and F850LP) are in the rest-frame B (F435W) band. To redshift each galaxy, we used a concordance cosmology with Ωtot = 1, ΩΛ = 0.73, and H0 = 71 km s−1 Mpc−1 (Spergel et al. 2003) to calculate the cosmological dimming and loss of resolution corresponding to each redshift. Because the size of the ACS PSF does not change with redshift, we assume that the size of AGN central point sources will also stay fixed with redshift, unlike the host galaxy. We therefore created the redshifted AGN by redshifting the model host galaxy (in flux and size) separately from the central point source (in flux only) and adding them together before convolving with the ACS PSF and adding noise to the image. The redshifting of each of the 12,592 galaxies produced another 12,592 galaxies at each final redshift, for a total of 50,368 fully simulated galaxies located at four different redshifts from 0.125 < z < 1.075. We fit each using our batch-fitting algorithm. Each galaxy was fit twice: once with a generalized S´ersic profile plus a point-source component, and once with a combination of de Vaucouleurs bulge, exponential disk, and point source. Inactive galaxies were also fit without point-source components. Examples of fit results for typical galaxies in our sample are shown in Figure 3. 4. Results And Discussion 4.1. Existing Galaxies with Added Point Sources We are interested mainly in three properties of these models: how well they recover the input point-source magnitude, how well they recover the baseline host galaxy parameters, and how the accuracy of the host galaxy parameters relates to the point-source magnitude and/or the contrast ratio between central point source and host galaxy. 4.1.1. AGN (Point Source) Recovery Figure 4 shows the degree of recovery of the central point-source magnitude for input point sources. As expected, brighter point sources are more accurately recovered. At all -- 11 -- Fig. 3. -- Example fit results for four fully simulated galaxies at four different redshifts. Each row shows one galaxy. The first two columns show the image of the galaxy at z = 0.125 and residual. Each successive pair of columns from left to right shows galaxy and residual for z = 0.4125, 0.7375, and 1.075, respectively. Note: These cutouts are zoomed-in to show the galaxies; the actual fitting regions are significantly larger. -- 12 -- values of input point-source magnitude, the fitted value of the point-source magnitude tends to either converge to the true input magnitude or remain at the input guess magnitude. In all cases, we can easily separate these two groups with a simple reduced χ2 ν cut (χ2 ν < 2.0). After the cut, the number of remaining poor fits is very low, as is the number of good fits removed by the cut: both are less than 5%, even in the limiting cases of very high and very low input point-source magnitude. This level of contamination/excess removal of good fits is not strongly dependent on the value at which we choose to cut χ2, nor is it strongly dependent on the input point-source magnitude. The rms uncertainties of the recovered point-source magnitudes for each input point- source value are shown in Table 2. These values reflect the systematic uncertainties in the fitted parameters, and should be added to the statistical error in fitted point-source magnitude returned by GALFIT. 4.1.2. Host Galaxy Parameter Recovery Figure 5 shows the correlation between the fitted model's host galaxy magnitude and the actual host magnitude (as determined before the addition of the central point source). The percent completeness of the low-error sample compared to the full sample varies greatly across a range of contrast ratios between host galaxy and central point source. For very faint central point sources, the host galaxy magnitude is very well recovered, whereas for very bright point sources the fraction of host galaxies with well-recovered parameters is very low (∼ 15%). Our simulations show that the reduced χ2 ν parameter is generally not sufficient to dis- tinguish good from poor host galaxy fits. If we also limit the relative error of the effective radius parameter, however, the fraction of selected galaxies that recover the true magnitude of the host galaxy improves significantly. With the combination of χ2 ν and σre/re parameters, we remove more than 85% of the fits with magnitude differences of 0.5 dex or more from the actual galaxy magnitudes. The number of well-fitted galaxies that are also removed from the sample ranges from 5% to 20%, depending primarily on how conservatively we choose our relative σre cut. Figure 5 reflects a cutoff value of (σre/re) = 0.8, which removes 88% of poor fits and 12% of the good fits from the IPS = 24 sample. Figure 6 shows how well we recover the S´ersic index, n. The same error cuts (χ2 ν and σre/re) efficiently reject the inaccurate fits. The fraction of models whose fitted n values deviate by more than ∆n = 2 compared to the n values of the unaltered galaxies increases significantly with increasing point-source flux. That is, a very bright central point source -- 13 -- Fig. 4. -- Recovery of magnitude of a point source added to real galaxy images, for: input point-source magnitudes of 22, 24, 26, and 28 mag. The fitting recovers the point-source magnitude nicely after making a simple χ2 ν < 2.0 cut (black crosses), with a larger spread for fainter values of the input point source, as expected. Fits with a large value of χ2 ν are shown as green open circles; Nconv indicates the number of fits (out of 450) that converged automat- ically for this input point-source magnitude. The dashed horizontal line shows mout = min, so fits falling on this line perfectly recover the actual point-source magnitude. The dotted line shows mout = mguess: fits on this line converged to their initial guess point-source magnitude, and the final fitted magnitude is wrong; nearly all of these fits are removed by the χ2 ν cut. -- 14 -- Fig. 5. -- Fitted host galaxy magnitude versus actual galaxy magnitude for fits to actual galaxies with added point sources. Excluding high χ2 ν values and effective radii with large fractional errors ensures a high fraction of good fits, with moderate uncertainties. Green open circles indicate fits with high χ2 ν values. Cyan open squares indicate fits with excessive errors on the effective radius fit parameter (σre ≥ 0.8 × re). Purple open triangles represent fits with both high χ2 ν and high σre. The dotted line represents equal input and recovered magnitudes (i.e., perfect recovery). The error cut removes some of the fits that do lie on the 1:1 line, regardless of the precise value for the σre cut. -- 15 -- can easily be confused with a concentrated (high-n) S´ersic value. For all values of the added point-source magnitude, however, the same combination of χ2 ν and relative re error is sufficient to remove 88% of the inaccurate fits from the data, while removing less than 15% of the well-fitted models. 4.1.3. Dependence of Parameters on Host:AGN Contrast Ratio Figure 7 shows the relation of fitted host galaxy parameters to the contrast ratio between the host galaxy and the added central point source, for the 1491 (of 4050) sources that are not eliminated by the combination of χ2 ν and σre/re cuts. These include galaxies with all possible values of input point-source magnitude. The left panel of Figure 7 shows the difference between the fitted host galaxy magnitude and original galaxy magnitude for the PS-added models. The recovered magnitudes are no different on average than the correct parameters. For galaxies with contrast ratios of at least log (Lhost/LPS) < −0.6, the host galaxy magnitude is well recovered, such that a weighted least-squares fit to those data with contrast ratios (Lhost/LPS) greater than 1:4 has a slope within 1% of zero, indicating no correlation between contrast ratio and recovered host galaxy magnitude. The scatter in the plot indicates that, while the sample as a whole is well-recovered, individual galaxies may have large differences in recovered parameters due to the presence of a point source. We see this result again in the right panel, which shows the difference between fitted and original S´ersic index as a function of host-to-PS contrast ratio. The inaccuracy in S´ersic index is larger when the added central point source is more than 4 times as bright as the host galaxy: the slope of the weighted least-squares fit is 6% for ∆IAB and 11% for ∆n. With proper error cuts, and for galaxies brighter than 1/4 of the point-source magnitude, our host galaxy fits are not significantly contaminated by the central point source. However, the presence of a point source increases the uncertainty in fit parameters. Figure 8 shows the median change in fitted parameters with input point-source magni- tude. For very bright added point sources (IAB = 21, hHost:PSi ≈ 1 : 11), all fit parameters deviate from their initially-determined values. The median S´ersic index is low by approx- imately ∆n = 1.75, indicating that on average, galaxies with such bright point sources could be classified as disk-dominated even if they have more bulge-dominated intrinsic mor- phologies. The distribution is also very wide in the brightest-point-source bin: for such low host-to-PS contrast ratios, the S´ersic index is essentially completely uncertain. For fainter point-source magnitudes, recovery of the S´ersic index is far more reliable, although the un- certainty (indicated by the width of the distribution in each bin) is considerably larger than -- 16 -- Fig. 6. -- Change in S´ersic indices of host galaxies after adding point sources with original host magnitude. Excluding high χ2 ν and σre/re yields the correct S´ersic index (a proxy for morphological type). The color-coding of the points is the same as in Figure 5: green open circles for high χ2 ν, cyan open squares for high relative σre, and purple open triangles for both high χ2 ν and relative σre. The dotted lines at ∆n = 0 indicate perfect recovery of the S´ersic index. As expected, the bright-point-source fits return mostly unreliable host galaxy parameters, in contrast to the faint-point-source fits, which return mostly reliable host galaxy parameters. -- 17 -- Fig. 7. -- Accuracy of the host galaxy magnitude (left) and S´ersic index (right) as a func- tion of the ratio of intrinsic galaxy luminosity to the added point-source luminosity. In both figures, contour plots are overlaid on the data points. Weighted least-squares fits to all data points (dashed lines) indicates that the recovery of galaxy magnitude and S´ersic index is very slightly correlated with host-to-PS contrast ratio. Excluding points with log (Lhost/LPS) < −0.6, the weighted least-squares fit (solid lines) is consistent with zero slope (i.e., no correlation between Host-to-PS contrast ratio) for both plots. -- 18 -- that reported by the fitting routine. The ratio of host and point-source luminosities is also incorrect at the brightest point- source magnitudes: in these situations the typical fitted point-source magnitude is very close to the input value, and the fitted host magnitude is too low (bright) by approximately 1.45 dex. Interestingly, the deviations in magnitude do not add to zero: because bright point- sources are also slightly too bright, the total magnitude is also underestimated (too bright) in the presence of a very bright central point source. For point sources brighter than IAB = 24, the distribution of recovered host galaxy magnitude is wide, such that the 68% confidence intervals span more than 1 dex in all bins. For those point sources with magnitudes fainter than 24 (which is approximately equal to the average magnitude of our initial galaxy sample), on the other hand, the model's host galaxy magnitude is well recovered. Since the faintest point-source magnitudes (IAB = 28, 29) are below the 10σ point- source detection limit of our band, the fitted point-source magnitude for these cases is often at least 1 mag brighter than the input point-source magnitude, with 68% confidence intervals larger than 1 mag. The S´ersic indices for the host galaxies are within ∆n = 0.25 on average for most values of the added point-source magnitude. Since disks are generally more extended than bulges, the S´ersic parameter is lower (on average) when the central point source is so bright that the central portion of the host galaxy is dwarfed by the light in the wings of the central point source. In all but the brightest cases, however, the median difference between the model S´ersic index and the no-point-source index is smaller than the size of the morphological bins discussed for a typical AGN sample (e.g., S´anchez et al. (2004); Ballo et al. (2007), S08), so it is unlikely that a significant fraction has been mis-classified as disk-like or bulge-like due to the presence of a central point source. The rms values of the recovered S´ersic indices and recovered host magnitudes for each input point-source magnitude are given in Table 2. 4.2. Fully Simulated Galaxies 4.2.1. Input Parameter Recovery For our 12,592 fully simulated GOODS galaxies, we apply the error cut determined above, using a combination of χ2 ν and σre/re to distinguish good fits from poor fits. Within the single component galaxies, this error cut results in 706 of 896 B type galaxies with "good" -- 19 -- Fig. 8. -- Median change in S´ersic index n (upper), host magnitude (middle), and point- source magnitude (lower) versus input point-source magnitude. Each point represents the median value of all the galaxies with the given input point-source magnitude whose fits pass the χ2 ν and σre/re cut described in the text. The host S´ersic indices (top panel) are generally well-recovered for all but the brightest input point sources. The host magnitudes (middle panel) are also well-recovered in all but the brightest input point-source bin, with the error bars (representing the width of the distribution encompassing 68% of sources) increasing with increasing point-source strength. Point-source magnitudes (bottom panel) are recovered well until the input magnitude is fainter than the published 10σ point-source detection limit of the data. -- 20 -- fits and 708 of 896 D type galaxies with "good" fits. The recovery of the S´ersic index for both of these types of galaxy models is shown for the z = 0.125 sample in Figure 9. If we classify galaxies as bulge-dominated when their fitted n > 2.0 and disk-dominated when their fitted n < 2.0, the level of misclassification based on these simulations is less than 15% for AGN host disk galaxies and less than 10% for AGN host bulge galaxies. Fewer than 10% of inactive disks are misclassified as bulges using the above cut; no inactive bulges are misclassified as disks. For any local sample of single-component galaxies (the simplest possible case), then, the S´ersic parameter is a very reliable indicator of morphological type. The recovery of the z = 0.125 sample of two-component (B+D) galaxies is shown in Figure 10. The fitted bulge-to-total light ratio, B/Tot = LB/ (LB + LD), is within ∆ (B/Tot) ≤ 0.15 for 50% of our AGN host sample and 66% of our inactive sample. A Gaussian fit to the data in Figure 10 gives an estimate of uncertainty in the bulge-to-total parameter of σB/Tot = 0.04. However, the contamination from very deviant fits (i.e., the wings of the histogram in Figure 10) is large. Approximately 14% and 13% of host and inactive galaxies remaining in the sample after our error cut have ∆ (B/Tot) < −0.15, and 36% of host galaxies and 21% of inactive galaxies have ∆ (B/Tot) > 0.15. Figure 11 shows the ratio of fitted-to-input bulge-to-total flux ratio as a function of the intrinsic contrast ratio between the host galaxy and the central point source. The median and 1σ error bars (defined here as the values encompassing 68% of the points in each bin) for the distribution of simulated galaxies without central point sources are plotted for comparison at a value of Lhost/LP S = 3. In general, the presence of a central point source skews the fits to slightly high values of the bulge-to-total ratio. For galaxies where the central point source is very bright, the spread is higher, such that a galaxy may fit to a bulge-to-total ratio twice the intrinsic value and still be within 1σ of the median value. Based on this, one could attempt to correct the bulge-to-total ratio for the presence of a central point source, but a bulge-to-total ratio fitted to a host galaxy could be overestimated by as much as 100%. This large tail of high fitted bulge-to-total values has several causes. The radius of the bulge may converge to an unphysically high value; this causes the luminosity bulge-to-total ratio to skew high. This appears to be slightly more common when the intrinsic luminosity of the galaxy is disk-dominated, i.e., with (B/Tot) ≤ 0.3. In addition, if the total source magnitude is not well-conserved during the fit, the bulge-to-total ratio typically deviates from the input value by at least 10% (high or low). This is more likely to occur when the bulge and disk components have approximately equal luminosity, or the bulge is slightly brighter [0.45 ≤ (B/Tot) ≤ 0.7]. -- 21 -- Fig. 9. -- Recovery of S´ersic index n for the single-S´ersic fits to simulated single-component B and D AGN host galaxies (blue, top) and inactive galaxies (black, bottom). For the bulge- dominated B galaxies (solid histograms), the average fitted S´ersic index is slightly higher than the input n = 4 in both AGN and inactive galaxies; however, these are unambiguously classified as bulge-dominated galaxies, with fewer than 10% of bulge host galaxies mis- classified as disk-dominated galaxies and no bulge inactive galaxies mis-classified as disks. For the disk-dominated D galaxies (dot-dash histogram), the recovery of the input n = 1 is also excellent, with 13% of D host galaxies and 9% of D inactive galaxies having n > 2.0. The separation between the typical fitted n for B and D galaxies is robust and indicates that we can use the S´ersic index as a strong indicator of galaxy morphology for single-component local galaxies with and without central point sources. -- 22 -- Fig. 10. -- Recovery of input bulge-to-total ratio for the fully-simulated two-component galaxies. We define ∆ (B/Tot) > 0 to mean the fit was more bulge-like than the input galaxy. A Gaussian fit to the Host galaxy distribution gives σBT = 0.04, but there is a large asymmetric tail on each distribution indicating that the samples are skewed toward more bulge-like fits than the intrinsic distribution. 36% of host galaxies and 21% of inactive galaxies have ∆ (B/Tot) > 0.15. -- 23 -- Fig. 11. -- Ratio of fitted-to-input bulge-to-total vs. input host-to-PS contrast ratio. Points plotted are the median values in each bin, with error bars marking the range of values containing the central 68% of the points in each bin. The open circle shows simulated galaxies with no added point source; this bin recovers input bulge-to-total ratio closely, with a tail of high bulge-to-total values. For simulated host galaxies, the recovered bulge-to-total ratio is generally high by 10-20%, and the distribution of points in each bin is wider for galaxies where the central point source is at least as bright as the host galaxy. -- 24 -- About 18% of the two-component AGN host fits result in either a bulge or disk half- light radius that is unphysically low given the input sizes of both in our sample. These fits span the entire range of ∆ (B/Tot) values, and are far more common when the central point source is intrinsically bright. For the locally-defined sample, 79% of B+D bulge components and 63% of B+D disk components with fitted re < 5.0 pixels are those with the brightest central point source; this comprises 26% of all two-component sources with the brightest central point source. It is worth noting that these fits typically show residuals that clearly indicate a poor fit, and they can therefore be removed from the well-fitted sample on visual inspection, or followed up with individual fitting to improve the fit. The recovery of bulge and disk sizes in our sample is shown in Figure 12. The distribution of ∆re for single-component galaxies (pure bulges and disks) is sharply peaked at a value of ∆r ≈ 0 (in pixels). A total of 66% of pure bulges and 65% of pure disks recover their input radii within −1 ≤ ∆re ≤ 12 pixels. The discrete peaks at negative ∆re represent those fits that converged to unphysically low values (re < 5.0 pixels for the local sample). Less than 5% of single-component fits converged to extremely large radii (∆re > 150 pixels). The radius parameter converges to extremely high values far more frequently for the two-component fits, and with higher frequency for the bulge components of two-component galaxies (20%, B+D bulges) than for the disk components of two-component galaxies (13%, B+D disks). The high-∆re tail for B+D bulges is a result of confusion with the central point source of the simulated AGN host. Not only are bulges more compact than disks within our B+D galaxies (and typically within real galaxies as well), but the slope of their light profile more closely resembles the light profile of the ACS PSF than does an exponential disk profile. This can lead to uncertain bulge radii and magnitudes, especially when the intrinsic luminosity of the bulge is similar to that of the central point source. Disk components of B+D galaxies are not as readily confused with the central point source: their fit typically fails when the surface brightness of the disk component is too low compared to the bulge and point source components. Figure 13 demonstrates a possible method of separating good from poor two-component fits. After removing those fits where either the bulge or disk converged to re < 5.0 pixels, additionally requiring that the total source magnitude be conserved to within ∆mtot < 0.5 (determined by comparing the fitted total source magnitude with a measured, aperture- corrected magnitude), and/or that the flux-weighted effective radius of the fit be no more than 90 pixels larger than the measured half-light source radius1 removes 81% of sources with ∆ (B/Tot) > 0.15 and retains 83% of sources within 0.15 of the input bulge-to-total 1This value was chosen because it is twice the observed radius of the largest source in our sample. -- 25 -- Fig. 12. -- Histograms of the change in half-light radius for pure bulge (top left), pure disk (top right), B+D bulge (bottom left) and B+D disk (bottom right) fits to simulated AGN hosts. The majority of fits converge to a value that is equal to or greater than the input value of the half-light radius for that source or source component. The small discrete peaks at negative ∆re values are comprised mostly of fits for which re converged to an unphysical value for our sample, re < 5 pixels. Single-S´ersic fits typically recover the sizes of pure bulges and pure disks with about the same level of success. However, the local sample of two-component galaxies indicates that bulge component radii are more likely to converge to a value that is at least 50 pixels too high, or at least twice the intrinsic radius. Local disk components are more likely to be underestimated, with a lower σ for the high-∆re distribution than for B+D bulges. -- 26 -- Fig. 13. -- Deviation from input host-to-PS contrast ratio versus deviation in bulge-to- total ratio. Fits located at (0, 0) perfectly recovered the input bulge-to-total ratio and the Galaxy-AGN luminosity ratio (Lhost/LPS). Requiring that the difference between the half- light radii of the host be no more than 90 pixels larger than that observed for the entire source (∆re ≤ 90), that the total source magnitude luminosity be conserved to better than ∆mtot ≤ 0.5 mag, and that the radii of neither the bulge nor the disk component be less than 5 pixels (typically indicative of a poor fit) cuts 81% of sources with ∆ (B/Tot) > 0.15, and retains 83% of sources with ∆ (B/Tot) ≤ 0.15. -- 27 -- ratio. This also constrains the fitted host-to-PS contrast ratio to within 8% of the input value for 95% of the remaining "good" sample. However, this selection biases the remaining sample toward sources with faint central nuclei and preferentially retains bulge-dominated sources with intrinsic (B/Tot) > 0.7. In general, the result of two-component fits should be used with caution, even for a local, well-understood sample. 4.2.2. Detection of Central Point Sources We established in §4.1.1 that central point sources can be recovered on average for a sample of simulated host galaxies created from the addition of a central point source to a real GOODS inactive galaxy, although the fraction of recovered point sources varies depend- ing on the luminosity of the added point source. Using the χ2 ν cut previously determined from this sample, we now examine the recovery of point sources with redshift and intrinsic morphological type. Figure 14 shows the fraction of fits in which the central point source was detected for our entire sample of fully simulated galaxies. In this case, our criteria for a "detected" point source is simple: the point source is considered undetected only if the magnitude parameter converged to a value considerably fainter than the detection limit of our simulated data (mP S = 27.1 for our sample; Giavalisco et al. (2004)). Thus Figure 14 shows the fraction of fits for which any point-source magnitude was recovered, rather than those fits for which the magnitude is close to the intrinsic magnitude. Figure 14 indicates that the point source is detected within the fit at least 90% of the time for all values of input log (Lhost/LPS). The fraction is lowest when the point source and host galaxy have about equal luminosity. The detection fraction increases to more than 95% at the highest values of input log (Lhost/LPS), even for cases in which the point source is 100 times fainter than its host galaxy. This is due to the fact that the convergence fraction is lower for these bins; fits where the central point source may not have been detected are less likely to converge, driving the detection fraction higher for those fits that did converge. While we are primarily interested in the recovery of host galaxy parameters and optical detection of central AGNs, we also want to characterize the fraction of spurious optical point-source identifications. Thus, Figure 15 shows the results of a S´ersic + PS fit to inactive galaxies with no intrinsic central point source. We find a significant difference in the number of false point-source detections between pure disk and pure bulge galaxies. Only ∼ 1% of pure disk inactive galaxies detect2 a central point source when none are present. However, we 2We define "detected" as having been fit with a magnitude brighter than the published 10σ point-source detection limit of the GOODS survey (from which we take our noise properties). -- 28 -- Fig. 14. -- Fraction of point sources recovered for single-component fits to all simulated AGN host galaxies. The fraction of recovered point sources is slightly lower when the with input log (Lhost/LPS) is near zero (LP S ≈ Lhost or slightly fainter). However, even in these cases, the fits detect at least 90% of point sources. -- 29 -- find a point source nearly one-quarter of the time when the simulated galaxy is a pure bulge. Because our total sample is composed of equal numbers of pure bulge and pure disk galaxies, this gives an overall false detection of a central point source in 12% of inactive galaxies. However, this effect is strongly dependent on morphology: fits to elliptical galaxies are far more likely to detect a faint point source when none are present than fits to disk galaxies. Host galaxy studies estimating the uncertainty in the fraction of hosts with detected optical point sources should therefore consider the morphological composition of their data sample: if the sample contains mostly elliptical galaxies, the number of false detections could be as high as 25%. 4.2.3. Redshift Effects on AGN Host Galaxies At higher redshifts, surface-brightness dimming causes loss of extended components, and an unresolved central AGN corresponds to a larger fraction of a galaxy's central region at high redshift for fixed spatial resolution. Because our GOODS data sample (S08), as well as those of others (S´anchez et al. 2004; Grogin et al. 2005; Ballo et al. 2007; Alonso-Herrero et al. 2008), extends to z ≈ 1, we consider the quantitative effects of redshift on morphological classification. After applying data cuts requiring ∆re ≤ 90 pixels and ∆mtot ≤ 0.5 to our results at all redshifts, we assess the reliability of using the S´ersic index to classify single-component galaxies at different redshifts in Figure 16. We have "classified" a galaxy as bulge-dominated if it has n > 2.0 and disk-dominated if it has n < 2.0. Figure 16 then shows the fraction of correctly classified B- and D-type galaxies at each redshift. In general, the classification of pure bulges and pure disks is highly reliable for AGN host galaxies and point sources. At all redshifts, disks and bulges are correctly classified at least 90% of the time using these data cuts. Without these cuts, the misidentification rate increases to as much as 15%. The average recovery of input half-light radius for all our simulated galaxies (pure bulges and disks as well as B+D bulges and disks) is shown in Figure 17. Each point represents the peak of the distribution of ∆re values for a single galaxy type/component in each redshift bin. The error bars give the limits of the central 68% of data points; the peak of each distribution is generally not coincident with the numerical "center" of each bin. Single-component S´ersic fits to pure bulges and disks recover the radius reliably at all redshifts. Fitted radii to bulge and disk components in B+D hosts recover the intrinsic radius of the component to within 1 kpc on average; however, the errors are larger and these fits are more uncertain. For example, the width (containing 68% of sources) of the ∆re distribution of z ≈ 1 pure disks is less than 1 kpc, whereas the width of the z ≈ 1 distribution of ∆re for disk components in B+D hosts -- 30 -- Fig. 15. -- Distribution of fitted point-source magnitude for simulated inactive galaxies. Only 1% of pure disk inactive galaxies (blue dashed histogram) recovers point-source mag- nitudes brighter than the 10σ detection limit (dotted line) for the GOODS survey (from which our noise properties are taken). When the galaxy is a pure bulge (black histogram), however, a point source is found above the 10σ threshold in approximately 25% of fits. The overall percentage of fits where a point-source is found when none exists is 12% (for both pure bulges and pure disks combined), but this number is strongly dependent on intrinsic morphology. -- 31 -- Fig. 16. -- Dependence of galaxy classification on redshift for simulated bulge galaxies (nin = 4; top, solid lines) and disk galaxies (nin = 1; bottom, dot-dashed lines) with (blue) and without (black) central point sources. The y-axis is the fraction of galaxies correctly identified as bulges or disks given a threshold of n > 2.0 for bulge-dominated classification and n < 2.0 for a galaxy with a significant disk. Although we have here adopted a cutoff of n = 2.0, the results change by less than 5% when we change the cutoff to values between 1.5 < n < 3. -- 32 -- Fig. 17. -- Recovery of input radii for pure and B+D bulges (top panel) and disks (bottom panel). Each point represents the peak of the distribution of single-component (black, solid lines) and two-component (green, dashed lines) ∆re at each redshift. The numbers next to each point indicate the percentage of completed fits with re > 5 and ∆re ≤ 90 pixels, so that the fit converged to a physical value and would not be rejected by eye based on the fit residuals. While median fitted bulge and disk sizes are generally accurate to within 1 kpc at all redshifts, the presence of a central point source increases the spread in the distribution by up to a factor of 3. In addition, the majority of bulge fits within B+D galaxies at z ≈ 1 are not able to converge to a sensible value for the bulge radius. -- 33 -- is approximately 3 kpc. Although the average recovery of bulge and disk component radii in B+D galaxies is accurate to within 1 kpc, individual radii are highly uncertain. The percentage of sources remaining after removing fits with incorrect total magnitudes and unphysically low or high values of re, as previously discussed, is also shown for each bin in Figure 17. The number of fits for which the fitted re < 5 pixels remains relatively constant with redshift, but the number of fits for which the fitted re value diverges increases with redshift, more dramatically so for bulges than for disks. In the z ≈ 1 sample, fewer than 40% of bulges within two-component fits have fitted radii within a reasonable physical range. This is likely due not just to the presence of the additional disk component, but also to the fact that the bulge is much smaller at higher redshift compared to the size of the central point source. In these cases, the bulge and the point source can be confused, often resulting in a divergent bulge fit. This happens less frequently for disks for two reasons: first, disks are generally more extended than bulges, and second, the radial profile of the ACS PSF more closely resembles that of a de Vaucouleurs bulge than an exponential disk. Because different fractions of galaxies are removed from the sample at different redshifts (and for different bulge sizes), the morphological composition of a galaxy sample may be al- tered from its intrinsic composition simply by applying a reasonable cut on the morphological fit parameters, even if this cut is applied evenly (as it is in Figure 17). Figure 18 thus shows the fitted morphological fraction of bulges with redshift of our simulated sample of pure bulges and disks for both AGN hosts and inactive galaxies. After applying the radius and magnitude cuts above (i.e., removing only unphysical fit results from the sample), we calculate the fraction of galaxies in each redshift bin that would be classified as bulge-dominated according to an n > 2 cut. For a perfectly-recovered sample, this fraction would be N(n > 2.0)/Ntot = 0.5 at all redshifts because we simulated equal numbers of bulges and disks. However, the convergence rate for pure bulge and disk fits is slightly different at each redshift, and even our set of reasonable and conservative error cuts removes different fractions of bulge and disk galaxies at each redshift. While this does affect the morphological composition of the sample at each redshift, its effect is small: the morphological fraction of n > 2 galaxies for AGN hosts is typically within 2% of the intrinsic fraction at all redshifts. The inactive galaxy morphological fraction is within 5% of the intrinsic fraction at all redshifts; both samples are consistent within their uncertainties with the intrinsic fraction. For the two-component B+D galaxies, the dependence of fitted S´ersic index on input bulge-to-total ratio is shown in Figure 19. We see a general dependence of fitted n on the intrinsic bulge-to-total ratio. The relation changes somewhat as the sample is redshifted, but at all redshifts the S´ersic index indicates that galaxies may be fitted with a S´ersic index -- 34 -- Fig. 18. -- Fraction of galaxies classified as bulges with redshift. Blue circles represent simulated AGN hosts, and black triangles represent simulated inactive galaxies (error bars determined using Poisson statistics). Because the initial simulation contains equal numbers of pure bulges and disks, the actual bulge fraction is 0.5 for both samples. At the lowest redshift bin, these fractions are recovered to within 1% for both samples. The sample of simulated AGN host galaxies is mis-classified by up to 2% (0.02), and the inactive sample is mis-classified by up to 5%. Uncertainties are higher for inactive galaxies due to lower numbers of galaxies in the inactive sample, but in both cases the recovered fractions are consistent with the correct fraction within the uncertainties. -- 35 -- Fig. 19. -- Fitted S´ersic index vs. input bulge-to-total ratio, for simulated normal galaxies (black triangles) and galaxies with added point sources (blue circles). At all redshifts (low to high, top to bottom) the fit correctly finds disks (n = 1) at (LB/Ltot) = 0 and bulges (n = 4) at (LB/Ltot) = 1. However, some galaxies fit to S´ersic values of n ≥ 4 despite having up to 30% of their intrinsic light from a disk. Galaxies with intermediate values of n (2 < n < 4) likely have bulges that contribute between 20% and 70% of the total galaxy light. Points plotted are the median of each bin's distribution; error bars mark the widths of each distribution enclosing the central 68% of sources in the bin. -- 36 -- indicative of a pure bulge (n = 4) even in cases where a disk is present and contributing up to 45% of the total galaxy light. However, a fitted S´ersic index consistent with an exponential disk (n = 1) typically indicates very little bulge contribution (< 10%); galaxies fitted with n < 1.5 have (LB/Ltot) < 0.2. Intermediate values of the S´ersic parameter, 1.5 < n < 3, generally indicate a galaxy with both bulge and disk, but the bulge-to-total ratio may vary between 0.2 ≤ (LB/Ltot) ≤ 0.65, indicating that it is impossible to determine with a single- component fit whether a host galaxy with an intermediate S´ersic index is intrinsically bulge- or disk-dominated. Given that two-component B+D fits are also uncertain in the presence of a central point source, we conclude that determination of the bulge-to-total ratio is uncertain by at least 20% in AGN host galaxies, with a higher uncertainty for hosts with intermediate S´ersic indices and/or fitted (LB/Ltot). Figure 20 shows the recovery of detected input point sources, hosts, and host-to-PS contrast ratios for our simulated galaxies (B, D, and B+D) at all redshifts. Again we see that for the local sample, the recovery is excellent: not only are point sources detected at least 95% of the time, but the recovered magnitude is typically within at least 0.1 ± 0.12 dex of the intrinsic magnitude. In fact, point-source recovery is very good at all redshifts: the recovered magnitude is within 0.5 dex of the input magnitude for all redshifts and input point-source magnitudes. The dispersion in recovered values for the host galaxy magnitude is higher than that of point-source magnitude values for all redshifts. While in the local redshift bin the median recovered host magnitude is within 0.1 dex of the input magnitude, the central 68% of values has a larger spread. At higher redshifts, the host galaxy is more likely to converge to magnitudes that are too bright compared to the input magnitude. This effect is more pronounced for sources with a brighter input point source, as expected. For sources with a bright point source compared to the host, the difference is ∆mhost = 0.6 for z > 0.7. For z ≈ 1, the distribution of ∆mhost has an extended tail to negative (too-bright) values: a significant fraction of high-redshift sources overestimates the host brightness by up to 1.8 dex regardless of the host-to-PS contrast ratio. For z < 1, this tail also exists for all host- to-PS values. However, when the input point source is faint compared to the host, the host magnitude is fitted to within −0.8 < ∆mhost < 0.3 (1σ) for these redshifts. 5. Summary and Conclusions We simulated 54,418 GOODS ACS galaxy images of inactive and AGN host galaxies at redshifts 0.125 < z < 1.25, the largest sample of simulated two-dimensional galaxy morphologies to date. -- 37 -- Fig. 20. -- Effects of redshift on recovery of input point sources (top), input host magnitudes (middle), and host-to-PS contrast ratio, shown for all our simulated host galaxies. Each panel shows the fitted quantity versus the input point-source magnitude. Each line in a panel represents a different redshift bin: z = 0.125 (blue, circles), z = 0.4125 (green, squares), z = 0.7375 (orange, triangles), and z = 1.075 (red, stars). -- 38 -- Using a robust initial guess routine followed by batch-fitting, we performed single- component and bulge-disk host galaxy fits on these simulated galaxies, while also extracting the central point-source components of the AGN. For batch-fit galaxies with central point-source components, typically 60%-70% of the fits converge successfully and pass our error cuts. This percentage can be significantly increased by follow-up case-by-case fitting, starting from the results of the batch fitting routines. ν cut on the recovered fits. Using a combination χ2 We reliably extract central point-source magnitudes for the simulated AGNs using a simple χ2 ν < 2 and relative effective radius error cut of (σre/re) = 0.8 removes 88% of the poor host galaxy fits while excluding 12% of good host galaxy fits. The accurate extraction of AGN and host galaxy parameters does not depend on the host:AGN contrast ratio (Lhost/LPS), as long as it is greater than 1:4. Galaxy size is recovered to within ∆re < 1 kpc for single-component galaxies. For indi- vidual bulges and disks within two-component bulge + disk host galaxy fits, the distribution of recovered radii peaks within ∆re < 1 kpc at all redshifts, but the large width of the dis- tributions indicates that individual fits to bulges and disks within two-component galaxies are highly uncertain. The average recovered morphological fraction of our sample is within 5% of the true fraction, even at z ≈ 1. Of course, this number should be regarded as a lower limit since actual data samples of two-dimensional AGN host galaxy morphologies are likely to be significantly smaller than our simulations, and thus suffer from more uncertain statistics. We recover the correct morphology 99% of the time for inactive galaxies out to z = 1.25. The average bulge-to-total ratios are within 10% of the true ratio at all redshifts, but with large scatter (∼ 50%). This means that values for individual hosts are unreliable at the 50% level, but statistical population averages are more accurate. Thus, uncertainties in individual black hole masses determined using Lbulge − MBH relations (McLure & Dunlop 2002; Marconi & Hunt 2003; Ferrarese & Ford 2005) are higher when using fitted bulge lu- minosities of AGN host galaxies than those of inactive galaxies, and may be systematically overestimated due to the presence of central point sources. Galaxies with S´ersic n < 1.5 are generally disk-dominated, with at least 80% of their total light coming from a disky component. Galaxies with intermediate 1.5 < n < 3 have larger bulge components (occupying from 20% to 65% of the total galaxy light). Galaxies that appear to be bulge-dominated, i.e., with n ≥ 3, may derive as little as 45% (depending on redshift) of their total light from a bulge component; for n = 4, typically ∼ 70% of the galaxy light comes from the bulge. -- 39 -- In fully simulated AGN host galaxies, the central point source is correctly detected over 90% of the time, with only weak dependence on the intrinsic contrast ratio between the host galaxy and point source or the host morphology. This indicates that a sample of AGN host galaxy fits to ACS data with z < 1.25 is at least 90% complete in its detection of central point sources. However, we also detect spurious point sources in simulated galaxies where no central point source is present, for as little as 1% or as high as 25% of the galaxies, depending on the morphological composition of the sample. Fits to bulge-dominated galaxies are far more likely to detect a point source when none are present than fits to disk-dominated galaxies. We have used these simulations to evaluate the robustness of our results for AGN host galaxies observed with ACS. This is important for assessing the accuracy of recovered host morphologies, as well as the evolution of AGN host morphology with redshift. All of our results that can be quantitatively applied to a sample of real AGN host galaxies are also presented in tabular format (Tables 2, 3, and 4). We have performed these simulations using the image depths and noise properties of GOODS, but they may also be used for other ACS surveys such as GEMS (Rix et al. 2004), COSMOS (Scoville et al. 2007), and AEGIS (Davis et al. 2007). The extent to which these simulations are directly applicable or provide limits on the accuracy of AGN host galaxy morphological fits for other ACS data depends on the relative depth of observations in each survey. It is important to note that these results apply to redshifts where the ACS bands are observing rest-frame optical data, e.g. z < 1.25. As we move into the era of newer telescopes such as the James Webb Space Telescope and observing rest-frame optical wavelengths at very high resolution becomes possible for higher redshifts, an understanding of the effects of a central point source on high-redshift morphologies will become crucial. Simulations to probe these effects have begun on small scales (Dasyra et al. 2008), but more such work is needed. The authors wish to thank C. Conselice, L. Moustakas, P. Natarajan, S. Ravindranath, and the entire GOODS team for helpful discussions that improved this work. Thanks espe- cially to C. Peng for making GALFIT publicly available and for many enlightening discus- sions. We acknowledge support from NASA through grants HST-AR-10689.01-A, HST-GO- 09425.13-A and HST-GO-09822.09-A from the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy under NASA contract NAS 5-26555. -- 40 -- REFERENCES Alonso-Herrero, A., P´erez-Gonz´alez, P. G., Rieke, G. H., Alexander, D. M., Rigby, J. R., Papovich, C., Donley, J. L., & Rigopoulou, D. 2008, ApJ, 677, 127 Ballo, L., Cristiani, S., Fasano, G., Fontanot, F., Monaco, P., Nonino, M., Pignatelli, E., Tozzi, P., Vanzella, E., Fontana, A., Giallongo, E., Grazian, A., & Danese, L. 2007, ApJ, 667, 97 Bertin, E., & Arnouts, S. 1996, A&AS, 117, 393 Binney, J., & Merrifield, M. 1998, Galactic astronomy (Galactic astronomy / James Binney and Michael Merrifield. Princeton, NJ : Princeton University Press, 1998. (Princeton series in astrophysics) QB857 .B522 1998 ($35.00)) Conselice, C. J. 2003, ApJS, 147, 1 Dasyra, K. M., Yan, L., Helou, G., Surace, J., Sajina, A., & Colbert, J. 2008, ArXiv e-prints, 802 Davis, M., Guhathakurta, P., Konidaris, N. P., Newman, J. A., Ashby, M. L. N., Biggs, A. D., Barmby, P., Bundy, K., Chapman, S. C., Coil, A. L., Conselice, C. J., Cooper, M. C., Croton, D. J., Eisenhardt, P. R. M., Ellis, R. S., Faber, S. M., Fang, T., Fazio, G. G., Georgakakis, A., Gerke, B. F., Goss, W. M., Gwyn, S., Harker, J., Hopkins, A. M., Huang, J.-S., Ivison, R. J., Kassin, S. A., Kirby, E. N., Koekemoer, A. M., Koo, D. C., Laird, E. S., Le Floc'h, E., Lin, L., Lotz, J. M., Marshall, P. J., Martin, D. C., Metevier, A. J., Moustakas, L. A., Nandra, K., Noeske, K. G., Papovich, C., Phillips, A. C., Rich, R. M., Rieke, G. H., Rigopoulou, D., Salim, S., Schiminovich, D., Simard, L., Smail, I., Small, T. A., Weiner, B. J., Willmer, C. N. A., Willner, S. P., Wilson, G., Wright, E. L., & Yan, R. 2007, ApJ, 660, L1 Ferrarese, L., & Ford, H. 2005, Space Science Reviews, 116, 523 Ferrarese, L., & Merritt, D. 2000, ApJ, 539, L9 Gebhardt, K., Bender, R., Bower, G., Dressler, A., Faber, S. M., Filippenko, A. V., Green, R., Grillmair, C., Ho, L. C., Kormendy, J., Lauer, T. R., Magorrian, J., Pinkney, J., Richstone, D., & Tremaine, S. 2000, ApJ, 539, L13 Giavalisco, M., Ferguson, H. C., Koekemoer, A. M., Dickinson, M., Alexander, D. M., Bauer, F. E., Bergeron, J., Biagetti, C., Brandt, W. N., Casertano, S., Cesarsky, C., Chatzichristou, E., Conselice, C., Cristiani, S., Da Costa, L., Dahlen, T., de Mello, -- 41 -- D., Eisenhardt, P., Erben, T., Fall, S. M., Fassnacht, C., Fosbury, R., Fruchter, A., Gardner, J. P., Grogin, N., Hook, R. N., Hornschemeier, A. E., Idzi, R., Jogee, S., Kretchmer, C., Laidler, V., Lee, K. S., Livio, M., Lucas, R., Madau, P., Mobasher, B., Moustakas, L. A., Nonino, M., Padovani, P., Papovich, C., Park, Y., Ravindranath, S., Renzini, A., Richardson, M., Riess, A., Rosati, P., Schirmer, M., Schreier, E., Somerville, R. S., Spinrad, H., Stern, D., Stiavelli, M., Strolger, L., Urry, C. M., Vandame, B., Williams, R., & Wolf, C. 2004, ApJ, 600, L93 Graham, A. W. 2001, AJ, 121, 820 Grogin, N. A., Conselice, C. J., Chatzichristou, E., Alexander, D. M., Bauer, F. E., Horn- schemeier, A. E., Jogee, S., Koekemoer, A. M., Laidler, V. G., Livio, M., Lucas, R. A., Paolillo, M., Ravindranath, S., Schreier, E. J., Simmons, B. D., & Urry, C. M. 2005, ApJ, 627, L97 Haussler, B., McIntosh, D. H., Barden, M., Bell, E. F., Rix, H.-W., Borch, A., Beckwith, S. V. W., Caldwell, J. A. R., Heymans, C., Jahnke, K., Jogee, S., Koposov, S. E., Meisenheimer, K., S´anchez, S. F., Somerville, R. S., Wisotzki, L., & Wolf, C. 2007, ApJS, 172, 615 Jogee, S., Barazza, F. D., Rix, H.-W., Shlosman, I., Barden, M., Wolf, C., Davies, J., Heyer, I., Beckwith, S. V. W., Bell, E. F., Borch, A., Caldwell, J. A. R., Conselice, C. J., Dahlen, T., Haussler, B., Heymans, C., Jahnke, K., Knapen, J. H., Laine, S., Lubell, G. M., Mobasher, B., McIntosh, D. H., Meisenheimer, K., Peng, C. Y., Ravindranath, S., Sanchez, S. F., Somerville, R. S., & Wisotzki, L. 2004, ApJ, 615, L105 Koekemoer, A. M., Fruchter, A. S., Hook, R. N., & Hack, W. 2002, in The 2002 HST Calibration Workshop : Hubble after the Installation of the ACS and the NICMOS Cooling System, Proceedings of a Workshop held at the Space Telescope Science Institute, Baltimore, Maryland, October 17 and 18, 2002. Edited by Santiago Arribas, Anton Koekemoer, and Brad Whitmore. Baltimore, MD: Space Telescope Science Institute, 2002., p.339, 339 -- + Kormendy, J., & Richstone, D. 1995, ARA&A, 33, 581 Magorrian, J., Tremaine, S., Richstone, D., Bender, R., Bower, G., Dressler, A., Faber, S. M., Gebhardt, K., Green, R., Grillmair, C., Kormendy, J., & Lauer, T. 1998, AJ, 115, 2285 Marconi, A., & Hunt, L. K. 2003, ApJ, 589, L21 Marleau, F. R., & Simard, L. 1998, ApJ, 507, 585 -- 42 -- McLure, R. J., & Dunlop, J. S. 2002, MNRAS, 331, 795 Peng, C. Y., Ho, L. C., Impey, C. D., & Rix, H.-W. 2002, AJ, 124, 266 Ravindranath, S., Ferguson, H. C., Conselice, C., Giavalisco, M., Dickinson, M., Chatzichris- tou, E., de Mello, D., Fall, S. M., Gardner, J. P., Grogin, N. A., Hornschemeier, A., Jogee, S., Koekemoer, A., Kretchmer, C., Livio, M., Mobasher, B., & Somerville, R. 2004, ApJ, 604, L9 Rix, H.-W., Barden, M., Beckwith, S. V. W., Bell, E. F., Borch, A., Caldwell, J. A. R., Haussler, B., Jahnke, K., Jogee, S., McIntosh, D. H., Meisenheimer, K., Peng, C. Y., Sanchez, S. F., Somerville, R. S., Wisotzki, L., & Wolf, C. 2004, ApJS, 152, 163 S´anchez, S. F., Jahnke, K., Wisotzki, L., McIntosh, D. H., Bell, E. F., Barden, M., Beckwith, S. V. W., Borch, A., Caldwell, J. A. R., Haussler, B., Jogee, S., Meisenheimer, K., Peng, C. Y., Rix, H.-W., Somerville, R. S., & Wolf, C. 2004, ApJ, 614, 586 Scoville, N., Aussel, H., Brusa, M., Capak, P., Carollo, C. M., Elvis, M., Giavalisco, M., Guzzo, L., Hasinger, G., Impey, C., Kneib, J.-P., LeFevre, O., Lilly, S. J., Mobasher, B., Renzini, A., Rich, R. M., Sanders, D. B., Schinnerer, E., Schminovich, D., Shop- bell, P., Taniguchi, Y., & Tyson, N. D. 2007, ApJS, 172, 1 S´ersic, J. L. 1968, Atlas de galaxias australes (Cordoba, Argentina: Observatorio Astro- nomico, 1968) Sheth, K., Elmegreen, D. M., Elmegreen, B. G., Capak, P., Abraham, R. G., Athanassoula, E., Ellis, R. S., Mobasher, B., Salvato, M., Schinnerer, E., Scoville, N. Z., Spalsbury, L., Strubbe, L., Carollo, M., Rich, M., & West, A. A. 2008, ApJ, 675, 1141 Simard, L., Koo, D. C., Faber, S. M., Sarajedini, V. L., Vogt, N. P., Phillips, A. C., Gebhardt, K., Illingworth, G. D., & Wu, K. L. 1999, ApJ, 519, 563 Spergel, D. N., Verde, L., Peiris, H. V., Komatsu, E., Nolta, M. R., Bennett, C. L., Halpern, M., Hinshaw, G., Jarosik, N., Kogut, A., Limon, M., Meyer, S. S., Page, L., Tucker, G. S., Weiland, J. L., Wollack, E., & Wright, E. L. 2003, ApJS, 148, 175 Trujillo, I., & Aguerri, J. A. L. 2004, MNRAS, 355, 82 This preprint was prepared with the AAS LATEX macros v5.2. -- 43 -- Pure Bulges Pure Disks Bulge + Disk mgal (IAB, L∗) re (pixels, kpc) 16.898, 5.0 18.645, 1.0 19.398, 0.5 20.393, 0.2 16.898, 5.0 18.645, 1.0 19.398, 0.5 20.393, 0.2 13.54, 1.5 27.09, 3.0 40.63, 4.5 54.18, 6.0 36.12, 4.0 54.18, 6.0 72.23, 8.0 90.09, 10.0 Bulge Disk 16.898, 5.0 18.645, 1.0 19.398, 0.5 20.393, 0.2 20.750, 0.14 27.09, 3.0 40.63, 4.5 54.18, 6.0 16.898, 5.0 18.645, 1.0 19.398, 0.5 20.393, 0.2 20.750, 0.14 54.18, 6.0 72.23, 8.0 90.29, 10.0 b/a PA (degrees) mP S (IAB) 0.25 0.50 0.75 1.0 45.0 none 16.500 18.375 20.250 22.125 24.000 0.25 0.50 0.75 1.0 45.0 none 16.500 18.375 20.250 22.125 24.000 0.65 1.0 45.0 60.0 0.30 0.65 1.0 45.0 none 16.500 20.250 22.125 Table 1: Grid values for simulated galaxies at z = 0.125. The effective radius re values differ between the bulge and disk galaxies due to the different physical sizes of these two classes of galaxies. The parameter space varies slightly between single-component galaxies and double- component galaxies in order to increase the number of mesh points in the bulge-to-total ratio without increasing the simulated double-component galaxies to a computationally prohibitive number. Conversion from magnitude to luminosity and between pixels and physical size uses a concordance cosmology (Ωtot = 1, ΩΛ = 0.73, and H0 = 71 km/s/Mpc) and a redshift z = 0.125. -- 44 -- Source mP S Nconv σmP S N1 σmhost σn σre N2 (AB) 21 22 23 24 25 26 27 28 29 374 389 397 394 382 372 355 315 286 0.33 0.48 0.27 0.15 0.17 0.27 0.47 0.87 1.20 210 273 359 328 328 295 271 228 193 1.09 0.12 0.15 0.20 0.14 0.17 0.12 0.10 0.09 (pixels) 15.6 21.8 11.0 26.7 15.9 8.2 7.9 6.3 1.6 45 85 138 184 223 215 215 192 161 2.10 0.97 0.67 0.54 0.40 0.44 0.27 0.22 0.27 Table 2: RMS values for fit parameters vary with central point-source magnitude. Nconv is the number of fits which converged (out of 450) for each input point-source magnitude value. N1 is the number remaining after applying a χ2 ν < 2.0 cut; N2 is the number remaining after further requiring σre/re < 0.8. The RMS value for the point-source magnitude is calculated from N1 galaxies, and the values for σmhost, σn, and σr are calculated from N2 objects. Note: Results for these magnitudes are based on a survey with the GOODS depths (Giavalisco et al. 2004). Source log (Lhost/LPS) σmP S σmhost σn σre -1.5 -1.0 -0.5 0.0 0.5 1.2 1.9 0.26 0.31 0.31 0.36 1.69 1.92 0.69 0.78 0.29 0.31 0.35 0.32 0.23 0.07 (pixels) 7.2 6.2 7.1 5.4 4.4 4.6 3.4 0.70 0.38 0.31 0.27 0.24 0.33 0.20 Table 3: Determined RMS values for fit parameters to sources with different log (Lhost/LPS). -- 45 -- Fitted S´ersic index Intrinsic bulge-to-total ratio n ≤ 1.5 1.5 < n < 3 n ≥ 3 n ≥ 4 (B/Tot) < 0.2 0.2 < (B/Tot) < 0.65 (B/Tot) > 0.45 (B/Tot) > 0.55 Table 4: Relationship of a host galaxy's fitted S´ersic index to its intrinsic bulge-to-total light ratio. Hosts with intermediate S´ersic index may have a wide range of bulge-to-total ratios; hosts with fitted n = 4, typically classified as pure deVaucouleur bulges, may in fact have bulge-to-total ratios as low as 55%.
astro-ph/0009058
1
0009
2000-09-05T12:19:11
Search for High Proper Motion White Dwarfs
[ "astro-ph" ]
Recent results of microlensing surveys, show that 10-20% of the dark halo mass of the Milky Way consists of compact objects. The masses of these compact objects range from 0.1 to 1 solar mass. New theoretical cooling models for white dwarfs, and the detection of faint blue high proper motion objects, imply that ancient white dwarfs might make up part of this population. In this pilot project, using data from the C and D patches of the ESO Imaging Survey and data obtained at CTIO, we attempted to find such halo white dwarfs in the solar neighbourhood. With a time baseline of approximately one year between data sets and a limiting I-band magnitude of 23, we can find these high proper motion objects up to distances of 85 parsecs. In the 2.5 square degrees studied so far we have found three high proper motion candidates.
astro-ph
astro-ph
SEARCH FOR HIGH PROPER MOTION WHITE DWARFS J.T.A. DE JONG, K.H. KUIJKEN & M.J. NEESER Kapteyn Astronomical Institute, Postbus 800, 9700 AV Groningen, The Netherlands Recent results of microlensing surveys, show that 10-20 % of the dark halo mass of the Milky Way consists of compact objects. The masses of these compact objects range from 0.1 to 1 solar mass. New theoretical cooling models for white dwarfs, and the detection of faint blue high proper motion objects, imply that ancient white dwarfs might make up part of this population. In this pilot project, using data from the C and D patches of the ESO Imaging Survey and data obtained at CTIO, we attempted to find such halo white dwarfs in the solar neighbourhood. With a time baseline of approximately one year between data sets and a limiting I-band magnitude of 23, we can find these high proper motion objects up to distances of 85 parsecs. In the 2.5 deg2 studied so far we have found three high proper motion candidates. 1 Introduction: dark matter and white dwarfs It is generally accepted that the Milky Way and most external galaxies contain more mass than observed directly. What makes up this so-called 'dark mass' is, however, not so clear. This mass is often assumed to be distributed in an isothermal halo, which means that dark matter would also be present in the solar neighbourhood. Recent results from galactic microlensing experiments like MACHO 1 indicate that 10 to 20 % of the dark halo mass of the Milky Way may be in compact objects. Recent number counts of faint blue objects2 in the Hubble Deep Fields North and South and the detection of high proper motion objects3,4, indicate the presence of a population of faint blueish compact objects in the galaxy. Furthermore, new cooling models for white dwarfs5,6 show that they are much bluer than was previously believed. The above suggests that the galactic dark halo might, at least partially, consist of very old white dwarfs. In the project presented here we try to find faint high proper motion objects in the solar neighbourhood, which are likely to be dark halo white dwarfs (see the table on the right). For this purpose we use the C and D patches of the ESO Imaging Survey (EIS) Wide7 and a smaller survey done with the CTIO 4m telescope. These data serve as a pilot project for a larger survey for which we plan to observe the complete C and D patches of the EIS Wide survey with a larger temporal baseline. 2 Data and Method The data that are being used for the current project consist of two overlapping sets. The first data set is taken from the C and D patches of the EIS Wide survey7, which are 6 deg2 each. These data were obtained between July 1997 and March 1998 in the I-band with individual exposure times of 300 seconds. The second data set was obtained with the 4m telescope at CTIO. This survey covers almost 2 deg2 of both the EIS C and D patches, making up a total area of nearly 4 deg2. The CTIO data were taken in the R-band with single pointing exposure times of 2000 seconds. As this second survey was performed in December 1998, we have a time baseline of approximately one year for the high proper motion search. To find high proper motion objects from the data we use a straightforward method. First, the CTIO data are transformed to the EIS coordinate system. The EIS pixels are smaller than the CTIO pixels, so in this way we do not lose information. Following this, objects are matched within a 2 arcsecond radius. Objects that seem to be offset are selected with a magnitude dependent cutoff. Finally the resulting outliers are visually inspected to filter out close objects, cosmic rays, extended sources, etc. With the time baseline of ∼ 1 year we are sensitive to proper motions of 0.5 to 2.0 arcsecond per year. 3 What can we find? A high proper motion search can be very efficient in finding halo objects because of their large velocities with respect to the solar standard of rest. While the sun follows the galactic rotation, the halo does not. Therefore, halo objects have typical velocities of 200 km/s. In the event that we find an object with a high proper motion, how do we know whether it's part of the dark halo? In table 1 we compare the chances of finding objects belonging to the galactic components present in the solar neighbourhood. Our sensitivity to proper motions of 0.5 to 2.0 arcseconds, together with the typical velocities of the objects w.r.t. the sun, gives the distance range where we can find these objects. With these densities we can calculate the average number of objects per square degree. This shows that when we find a high proper motion object, it is most likely part of the dark halo. According to cooling models by Saumon & Jacobsen6, white dwarfs as cool as 3000 K have MI ∼ 16; easily detectable in our survey to distances of 85 pc. There are, however, some issues that should be mentioned. The two data sets are in different bands (R and I), and have different noise properties because of the different exposure times. Other possible explanations for apparent object motion include close (variable star) binaries, distant supernovae and even Kuiper belt and Oort cloud objects. The motions on the sky of Table 1: Calculated estimates of the average number of high proper motion objects one will find for different galactic populations. For these calculations, objects of 0.3 M⊙ are assumed. Population v distances ρ pop.II halo thin disk thick disk dark halo (km/s) 200 30 60 200 (pc) 20-85 3-13 6-25 20-85 (M⊙pc−3) 5 × 10−5 0.05 0.003 0.01 objects (deg−2) 0.001 0.03 0.01 2 the Kuiper and Oort objects are dominated by parallax, but because our temporal baseline is ∼ 1 year this possibility can not be excluded with our current data. All of these problems can be solved by doing a third observation run. 4 Results We have now analyzed 5/8 of the nearly 4 deg2 of data, and have found three promising candi- dates for dark halo objects. These are shown in the figures 1 to 3. The first two candidates are located in the C patch of the EIS Wide survey while the third one is in the D patch. The objects are very faint, which is a further indication that they are not normal stars. Both candidate 1 and 2 have an apparent I-band magnitude of 22. The third candidate is even fainter, around mI ∼ 23. It is also striking that the two C patch objects move approximately in the same direction. This direction is consistent with it being due to the galactic rotation of the sun. The same is true for the motion of the D patch candidate. 5 Future The recently discovered population of faint, high proper motion objects which may be part of the galactic dark halo needs to be studied in detail. For this purpose, a large number of objects is necessary. Therefore, we see the current project as a pilot project for a larger survey. The most efficient way of increasing the chances of finding these objects is to increase the field of view. For the candidates we have so far, further observations need to be done to confirm the proper motions and to get more color information and/or spectra. References 1. Alcock, C. et al, 2000, astro-ph/0001272 2. Mendez, R.A. & Minniti, D., 2000, ApJ, 529, 911, astro-ph/9908330 3. Ibata, R.A. et al, 1999, ApJ, 524, L95, astro-ph/9908270 4. Ibata, R.A. et al, 2000, ApJ, 532, L41, astro-ph/0002138 5. Hansen, B.M.S., 1999, ApJ, 520, 680, astro-ph/9903025 6. Saumon, D. & Jacobsen, S.B., 1999, ApJ, 511, L107, astro-ph/9812107 7. Benoist, C. et al, 1999, A&A, 346, 58, astro-ph/9807334 Figure 1: Candidate 1. The two images show the same 16x16 arcsec patch of the sky (located in the EIS Wide C field). On the left is the EIS I-band image, on the right the CTIO 4m R-band image, which was transformed to the EIS coordinate system. The apparent shift of the image is ∼ 1 arcsec. Figure 2: Candidate 2. The shift of this candidate is also ∼ 1 arcsec. This candidate is also located in the EIS Wide C patch and seems to be moving approximately in the same direction as candidate 1. This direction is consistent with it being due to the galactic rotation of the sun. Figure 3: Candidate 3. This candidate has an apparent shift of ∼ 0.5 arcsec and is near our brightness and proper motion detection limits. The direction of motion of this object is also consistent with it being due to the galactic rotation of the sun. This candidate is located in the EIS Wide D patch.
astro-ph/0210450
2
0210
2003-07-11T07:44:20
Cosmic shear statistics in the Suprime-Cam 2.1 sq deg field: Constraints on Omega_m and sigma_8
[ "astro-ph" ]
We present measurements of the cosmic shear correlation in the shapes of galaxies in the Suprime-Cam 2.1 deg^2 R_c-band imaging data. As an estimator of the shear correlation originated from the gravitational lensing, we adopt the aperture mass variance. We detect a non-zero E mode variance on scales between 2 and 40arcmin. We also detect a small but non-zero B mode variance on scales larger than 5arcmin. We compare the measured E mode variance to the model predictions in CDM cosmologies using maximum likelihood analysis. A four-dimensional space is explored, which examines sigma_8, Omega_m, Gamma and zs (a mean redshift of galaxies). We include three possible sources of error: statistical noise, the cosmic variance estimated using numerical experiments, and a residual systematic effect estimated from the B mode variance. We derive joint constraints on two parameters by marginalizing over the two remaining parameters. We obtain an upper limit of Gamma<0.5 for zs>0.9 (68% confidence). For a prior Gamma\in[0.1,0.4] and zs\in[0.6,1.4], we find sigma_8=(0.50_{-0.16}^{+0.35})Omega_m^{-0.37} for flat cosmologies and sigma_8=(0.51_{-0.16}^{+0.29})Omega_m^{-0.34}$ for open cosmologies (95% confidence). If we take the currently popular LCDM model, we obtain a one-dimensional confidence interval on sigma_8 for the 95.4% level, 0.62<\sigma_8<1.32 for zs\in[0.6,1.4]. Information on the redshift distribution of galaxies is key to obtaining a correct cosmological constraint. An independent constraint on Gamma from other observations is useful to tighten the constraint.
astro-ph
astro-ph
SUBMITTED TO APJ; DRAFT VERSION NOVEMBER 10, 2018 Preprint typeset using LATEX style emulateapj v. 11/12/01 COSMIC SHEAR STATISTICS IN THE SUPRIME-CAM 2.1 SQ DEG FIELD: CONSTRAINTS ON ΩM AND σ8 1 TAKASHI HAMANA2,3, SATOSHI MIYAZAKI2, KAZUHIRO SHIMASAKU4, HISANORI FURUSAWA2, MAMORU DOI5, MASARU HAMABE6, KATSUMI IMI6, MASAHIKO KIMURA8, YUTAKA KOMIYAMA2, FUMIAKI NAKATA2, NORIO OKADA2, SADANORI OKAMURA4, MASAMI OUCHI4, MAKI SEKIGUCHI8, MASAFUMI YAGI2 AND NAOKI YASUDA2 [email protected] 3 0 0 2 l u J 1 1 2 v 0 5 4 0 1 2 0 / h p - o r t s a : v i X r a Submitted to ApJ; Draft version November 10, 2018 ABSTRACT We present measurements of the cosmic shear correlation in the shapes of galaxies in the Suprime-Cam 2.1 deg2 Rc-band imaging data. As an estimator of the shear correlation originated from the gravitational lensing, we adopt the aperture mass variance, which most naturally decomposes the correlation signal into E and B (non- gravitational lensing) modes. We detect a non-zero E mode variance on scales between θap = 2′ and 40′. We also detect a small but non-zero B mode variance on scales larger than θap > 5′. We compare the measured E mode variance to the model predictions in CDM cosmologies using maximum likelihood analysis. A four-dimensional space is explored, which examines σ8, Ωm, Γ (the shape parameter of the CDM power spectrum) and ¯zs (a mean redshift of galaxies). We include three possible sources of error: statistical noise, the cosmic variance estimated using numerical experiments, and a residual systematic effect estimated from the B mode variance. We derive joint constraints on two parameters by marginalizing over the two remaining parameters. We obtain an upper limit of Γ < 0.5 for ¯zs > 0.9 (68% confidence). For a prior Γ ∈ [0.1,0.4] and ¯zs ∈ [0.6,1.4], we find σ8 = (0.50+0.35 - 0.37 m for Ωm + ΩΛ = 1 and σ8 = (0.51+0.29 for ΩΛ = 0 (95% confidence). If we take the currently popular ΛCDM model (Ωm = 0.3, Ωλ = 0.7, Γ = 0.21), we obtain a one-dimensional confidence interval on σ8 for the 95.4% level, 0.62 < σ8 < 1.32 for ¯zs ∈ [0.6,1.4]. Information on the redshift distribution of galaxies is key to obtaining a correct cosmological constraint. An independent constraint on Γ from other observations is useful to tighten the constraint. Subject headings: cosmology: observations -- cosmological parameters -- dark matter -- gravitational lensing - 0.16)Ω - 0.16)Ω - 0.34 m 1. INTRODUCTION Cosmic shear, that is coherent distortions in distant galaxy images resulting from weak gravitational lensing by large-scale structures, has been recognized as a powerful tool for cosmol- ogy because it directly probes matter distribution regardless of any relation between mass and light (for reviews see, Mellier 1999, Bartelmann & Schneider 2001). Since the first reports of the detection of cosmic shear correlations (Van Waerbeke et al., 2000; Witteman et al., 2000; Bacon, Refregier & Ellis 2000; Kaiser, Wilson & Luppino 2000; Maoli et al. 2001; ), cosmic shear statistics have become a promising probe of cosmologi- cal parameters. Indeed, recent measurements of cosmic shear correlation have put useful constraints on the matter density pa- rameter Ωm and the matter power spectrum normalization σ8 (Maoli et al. 2001; Bacon et al. 2002; Van Waerbeke et al. 2001; 2002; Hoekstra et al. 2002a; Hoekstra, Yee & Gladders 2002b; Brown et al. 2003; Jarvis et al. 2003). As the cosmic shear cor- relation is primarily sensitive to density fluctuation at interme- diate redshifts (0.2 < z < 0.7) and on scales from quasi-linear to nonlinear (1 < θ < 50 arcmin corresponding to comoving scales of 0.38 < r < 19.2h- 1Mpc at z = 0.5 for a cosmological model with Ωm = 0.3 and ΩΛ = 0.7), it provides cosmological information that is independent of other observations, such as galaxy clustering, cluster number counts, and the cosmic mi- crowave background anisotropies, and thus is complementary to these techniques. Cosmic shear measurement is not an easy task; it requires a rigorous observation strategy, specifically, a deep and wide- field survey with very high image quality. A large number den- sity of distant galaxies, ideally ng > 30arcmin- 2, is needed to suppress random noise due to the intrinsic galaxy ellipticity. A wide survey area is necessary for suppression of the cos- mic variance. In addition, good seeing conditions are essential for precise measurements of a galaxy's shape. Suprime-Cam, a wide-field camera mounted on the prime focus of the 8.2-m Subaru telescope, is an almost ideal camera for a weak lensing survey. It has a field of view of 34′ × 27′ with 0′′.202 pixel- 1. The median seeing in the Ic band, monitored over a period of one-and-a-half years, is reported to be ∼ 0.6 arcsec (Miyazaki et al. 2002b). Subaru's light-gathering power enables a com- plete magnitude of RC = 25.2 (the turn-around of galaxy counts) to be obtained by a 30-min exposure, which provides a galaxy number density of ng ∼ 33arcmin- 2, after object selection (see §3). These advantages allow a weak lensing survey to be car- 1 Based on data collected at Subaru Telescope, which is operated by the National Astronomical Observatory of Japan. 2 National Astronomical Observatory of Japan, Mitaka, Tokyo 181-8588, Japan 3 Institut d'Astrophysique de Paris, 98bis Boulevard Arago, F 75014 Paris, France 4 Department of Astronomy, University of Tokyo, Bunkyo, Tokyo 113-0033, Japan 5 Institute of Astronomy, University of Tokyo, Mitaka, Tokyo 181-0015, Japan 6 Department of Mathematical and Physical Sciences, Japan Women's University, Bunkyo, Tokyo 112-8681, Japan 7 Communication Network Center (Tsu-den) , Mitsubishi Electric, Amagasaki, Hyogo 661-8661, Japan 8 Institute for Cosmic Ray Research, University of Tokyo, Kashiwa,Chiba 277-8582, Japan 1 1992), the top-hat shear variance (Bernardeau, Van Waerbeke & Mellier 1997) and the aperture mass variance (Schneider et al. 1998). For our maximum likelihood analysis, we adopt the aperture mass variance, which has the very useful property that it naturally carries out E/B mode decomposition (Schneider et al. 1998; Crittenden et al. 2002; Pen, Van Waerbeke & Mellier 2002). As gravitational lensing produces only E mode shear, E/B mode decomposition allows contamination from B mode shear (which is not caused by gravitational lensing) to be sup- pressed, and the amplitude of the B mode variance can be used to estimate the amplitude of the systematic error. The aperture mass is defined by Map =Z d2θ U(θ)κ(θ), where κ(θ) is the lensing convergence field, and U(θ) is the compensated filter, for which we adopt the following function proposed by Schneider et al. (1998) U(θ) = 9 πθ2 ap(cid:18)1 - θ2 θ2 ap(cid:19)(cid:18) 1 3 θ2 θ2 ap(cid:19) , for θ < θap, and 0 otherwise. It should be emphasized that this filter probes an effective scale of θap/5 not θap. The aperture mass can be calculated directly from the tangential shear γt (the tangential component of shear defined in the local frame con- necting the aperture center to a galaxy), without the need for a mass reconstruction, by where Q(θ) is given from U(θ): Map =Z d2θ Q(θ)γt(θ), θ2Z dθ′ U(θ′) - U(θ). 2 Q(θ) = (3) (4) (5) (6) 2 T. Hamana et al. ried out efficiently. In this paper we present the results of our analysis of Rc- band imaging data from a Suprime-Cam 2.1 deg2 field. This field is composed of nine pointings in a 3 × 3 mosaic configu- ration which covers a contiguous 1.64◦ × 1.28◦ area. A great advantage of these data is their very good and homogeneous image quality over the whole field, which allows us to obtain secure weak lensing measurements. These data were also used for the first halo number counts using the weak lensing tech- nique (Miyazaki et al., 2002a), which demonstrated that a weak lensing halo survey is indeed a promising way to study the dis- tribution and evolution of large-scale structures. In this paper, we concentrate on the measurement of cosmic shear correla- tions caused by large-scale structures. We carry out a full max- imum likelihood analysis of the cosmic shear correlation over four parameters, Ωm, σ8, Γ (a shape parameter of the Cold Dark Matter power spectrum) and ¯zs (the mean redshift of the source galaxy distribution). We derive joint constraints on two param- eters by marginalizing over the two remaining parameters, and obtain a constraint on Ωm and σ8. We also obtain an upper limit on Γ. Confidence intervals on σ8 for the currently popu- lar ΛCDM model (Ωm = 0.3, ΩΛ = 0.7 and Γ = 0.21) are also derived. The outline of this paper is as follows. In §2 we briefly dis- cuss the theory of cosmic shear statistics and summarize the analytical formulas that are used to compute the theoretical pre- dictions. Details of the observations and data are described in §3. A galaxy shape analysis is presented in §4. Measurement of the cosmic shear correlation is presented in §5. The measured cosmic shear correlation signal is compared with the theoretical prediction using maximum likelihood analysis in §6. Finally, §7 presents a summary and discussion. In the Appendix, we discuss tests of the anisotropic point spread function correction procedure in some detail. 2. BASICS OF COSMIC SHEAR CORRELATION In this section, we provide a basic description of the theory of the cosmic shear correlation (see the reviews by Mellier 1999; Bartelmann & Schneider 2001 for details). The observable shear two-point statistics can be related to the convergence power spectrum defined by Pκ(l) = 9Ω2 m c (cid:19)4 4 (cid:18) H0 Z χH 0 a(χ) (cid:21)2 dχ (cid:20)W (χ) Pδ(cid:20) l fK(χ) ; χ(cid:21) , (1) where H0 is the Hubble parameter, χ is the radial comoving distance, χH corresponds to the horizon, a(χ) is the scale fac- tor and fK(χ) is the comoving angular diameter distance. Pδ(k) is the matter power spectrum, for which we adopt the fitting function of the CDM power spectrum given by Bardeen et al. (1986), but we treat the shape parameter, Γ, as a free pa- rameter. To take into account the effect of nonlinear growth of the density field, which has a significant impact on the shear correlation function on scales below one degree (Jain & Seljak 1997), we use the fitting function of Peacock & Dodds (1996). W (χ) is the source weighted distance ratio given by, W (χ) =Z χH χ dχ′ ns(χ′) , (2) fK(χ′ - χ) fK(χ) here ns(χ) is the normalized redshift distribution of source galaxies, which we discuss later (§6.1). Several real space estimators of the shear correlation have been proposed, including: the shear two-point correlation func- tion (Blandford et al. 1991; Miralda-Escude 1991; Kaiser The aperture mass variance is related to the convergence power spectrum eq. (1) by hM2 api(θap) = 2πZ dl l Pκ(l)(cid:20) 12 π(lθap)2 J4(lθap)(cid:21)2 , (7) where J4 is a fourth-order Bessel function of the first kind. We compute the aperture mass variances from the shear cor- relation functions using relations eqs. (11) and (12). This ap- proach has some advantages over direct measurement in that (i) it is the least affected by defects in the data, such as masking by bright stars, (ii) and it does not depend on the geometry of the data; (iii) thus it uses all information in the data (Hoekstra et al. 2002b). The two shear two-point correlation functions that are measured are: i, j wiw jγt,iγt, j i, j wiw j i, j wiw jγr,iγr, j i, j wiw j ξtt(θ) = PNs PNs ξrr(θ) = PNs PNs , , (8) (9) where θ = xi - xj, Ns is the number of source galaxies, γt and γr are the tangential and 45◦ rotated shear in the frame con- necting the pair of galaxies, and w is a weight that expresses the reliability of the shape measurement for each galaxy (dis- cussed in §4). For the following discussion, it is useful to de- fine, ξ+(θ) and ξ- (θ), which are the sum and difference of the and - Cosmic shear statistics in the Suprime-Cam 2.1 sq deg field 3 shear two-point correlation functions defined by eqs. (8) and (9), respectively ξ+(θ) = ξtt(θ) + ξrr(θ), and ξ- (θ) = ξtt(θ) - ξrr(θ). (10) The E and B modes' (which we denote by M⊥) aperture mass variances are derived by integration of these correlation func- tions with an appropriate window and hM2 hM2 0 api(θap) = πZ 2θap ⊥i(θap) = πZ 2θap 0 dθ θhW(θ)ξ+(θ) + W(θ)ξ- (θ)i , dθ θhW(θ)ξ+(θ) - W(θ)ξ- (θ)i , (11) (12) where W and W are given in Crittenden et al. (2002); useful analytical expressions were derived by Schneider et al. (2002). FIG. 1. -- Top panel: the size distribution of objects used for the cosmic shear measurements. Bottom panel: Average tr(Pγ ) as a function of rg. FIG. 2. -- Top panel:the solid histogram shows the number counts of objects used for the cosmic shear measurements, while the dashed histogram is for the product of the number counts and the average weight. Bottom panel: Average weight as a function of RC magnitude. 3. DATA Observations were made with the Suprime-Cam on the Sub- aru 8.2-m telescope during its commissioning phase. We used a field size of 2.1 deg2, which was the largest size possible during that period. The field that we chose was centered at R.A. = 16h04m43s, decl. = +43◦12′19′′(J2000.0). We obtained Rc-band images on the nights of 2001 April 23-25. Suprime-Cam has a field of view of 34′× 27′with a scale 0′′.202 pixel- 1 (see Miyazaki et al. 2002 for instrumental details of Suprime-Cam). Nine con- tiguous fields were observed in a 3×3 mosaic configuration. Each exposure on a given field was offset by 1 ∼ 2′from the other exposures to allow the removal of cosmic rays and de- fects on the CCDs. The total exposure time of each field was 1800 sec (360 sec × 5). We apply the weak lensing mass reconstruction technique to these data (Miyazaki et al. 2002a). The number counts of high peaks (above 5-σ), which represent massive dark halos, in the reconstructed convergence field is 4.9 ± 2.3, where the Gaus- sian smoothing radius of the convergence map is 1′. This result is consistent with predictions thatassume the Press-Schechter mass function (Press & Schechter 1974; we used the version modified by Sheth & Tormen 1999) and the universal NFW halo profile (Navarro, Frenk & White 1996) under the cluster normalized CDM cosmology. Thus, it is unlikely that this field is significantly far from the cosmic mean. The individual images were de-biased and then flattened us- ing a median of all the object frames taken during the observing run. Stacking the dithered images is not a trivial procedure be- cause of the large distortion in the optics, combined with the alignment error of each CCD with displacement and rotation from its nominal position. We present here an outline of how to obtain parameters that transform a CCD coordinate to a stan- dard celestial coordinate. First, we employ a geometrical model of the field distortion using a fourth order polynomial function R - r r = ar + br2 + cr3 + dr4, (13) where R and r are the distance from the optical axis in units of pixels on the face of the CCD and in celestial coordinates, re- spectively. We typically changed the telescope pointing by 1∼2 arcmin between successive exposures to fill the gaps between the CCDs. The offset and rotation of the telescope pointing between the exposures are also set as free parameters. All of these parameters can be determined by minimizing the distance of the same stars identified on different exposures. To do this, we adopt a modified version of mosaic f it, which is one of the functions of the imcat suite. Once we have obtained these parameters, we use them to warp each image before stacking. The residual of the distances between corresponding star images is a measure of the error of this mosaic-stacking procedure. The RMS value of the resid- uals is typically about 0.5 pixels. As shown in Miyazaki et al. (2002b), distortion parameters obtained in this way match quite well with the residuals predicted by optical ray-tracing programs, which implies that our solution is satisfactory. The RMS residual of 0.5 pixels is due to several effects that are not considered in the simple model, including atmospheric dispersion and asymmetric aberration of the optics. We note, however, that the residual vector changes smoothly with posi- tion and can be well modeled as a third order bi-linear poly- nomial function of position. This model then gives a fine cor- rection to our geometrical solution described above. The mea- 4 T. Hamana et al. surements of the displacement and the warping of the images are carried out using fitgeometry2 and mosaicmap of imcat, re- spectively, as described by Kaiser et al. (1999). As a result of these procedures, the final residual typically decreases down to 0.07 pixel RMS (14 milliarcsec). To derive a better astromet- ric solution, an external reference star catalog would have to be used. However, we simply employ each first exposure as a reference and accept the resulting moderate astrometric accu- racy, as it does not significantly affect the weak lensing analy- sis. The individual images are then warped using the solution, and stacked. The seeing in the resulting image is 0′′.68 FWHM and the scatter among the fields is quite small at 0′′.04 rms. To carry out object detection, photometry and shape mea- surements of objects, we use hfindpeaks, apphot and getshapes of the imcat suite, which are an implementation of Kaiser et al. (1995). Catalogs created for the nine fields are registered using stars in the overlapping regions to generate a final cata- log whose total field of view is 1.64◦× 1.28◦. Differences in the photometric zero point among the fields due to variation in the sky conditions (in turn, due to thin cirrus and differing air mass) are compensated for at this stage, but the adjustment is not significant (∼ 0.05 mag). We adopt slightly different object selection criteria to those used by Miyazaki et al. (2002a) to optimize our cosmic shear correlation measurement. Our criteria are (i) 22.5 < Rc < 26. (ii) The signal-to-noise (S/N) ratio, nu, calculated in imcat, ex- ceeds 7 (Erben et al. 2001). (iii) The image size is larger than the PSF size, rg > 1.45 ∼ 1.65, where rg is a measure of the size of objects (in pixel units) yielded by imcat. The PSF size is identified from the stellar branch in the size-magnitude (rg-Rc) plane. The PSF size varies slightly from pointing to pointing because of changes in the seeing conditions. (iv) Objects with rh > 10 pixels (where rh is the half light radius) are considered too large and removed. (v) Highly elliptical objects, where the observed ellipticity, eobs > 0.5, are removed. (vi) Objects that have a close companion, with a separation of less than 10 pixels (≃ 2 arcsec), are removed to avoid the problem of overlapping isophotes reported by Van Waerbeke et al. (2000). The num- ber of objects that pass these various selection criteria (i)-(vi) is 249,071 (32.9 arcmin- 2). Van Waerbeke et al. (2000) reported that regions where data from different CCDs are stacked together as a result of offsets between exposures (specifically, the edges of CCDs) can poten- tially produce discontinuities in the properties of the field and thus make the PSF correction difficult. This effect could cause a systematic error in the cosmic shear correlation measurement. Therefore, we decided to mask such regions. That is, we use only objects that were observed by the same CCD. About 35% of the objects are removed by this masking, and the number of objects in the final catalog is 161,740. The image size distribu- tion and magnitude distribution of the catalog are shown in the upper panel of Figure 1 and Figure 2, respectively. 4. GALAXY SHAPE ANALYSIS The observed ellipticity of galaxies is measured from the weighted quadrupole moments Ii j of the surface brightness f (θ): (14) where WG(θ) is the Gaussian window function. Estimating the shear, γ, from the observed ellipticities, eobs, involves two eobs =(cid:18) I11 - I22 I11 + I22 , I12 I11 + I22(cid:19) , Ii j =Z d2θ WG(θ)θiθ j f (θ), steps: First, the point spread function (PSF) anisotropy is cor- rected using the star images as references, e = eobs - Psm P∗ sm e∗ obs, (15) sm)- 1e∗ where Psm is the smear polarisability tensor, which is mostly obs was calculated for diagonal (Kaiser et al. 1995). (P∗ stars scattered over the field of view. We use unsaturated stars selected by the following criteria: (i) 20.6 < Rc < 23.0 (us- ing fainter stars than this does not change the results, see the Appendix). Note that the saturation level identified from the size-magnitude (rg-Rc) plane is Rc = 19.5 ∼ 20.5 which corre- lates with the seeing (with a fainter saturation level for a better seeing). (ii) nu > 15. (iii) The image size is within a narrow range of the seeing size, rg∗ - 0.05 < rg < rg∗ + 0.05, where rg∗ denotes the central rg value (in pixel units) of the stellar branch in the rg-Rc plane. rg∗ varies from pointing to pointing and is rg∗ = 1.27 ∼ 1.52. (iv) Highly elliptical objects, with an observed ellipticity, e > 0.3, are removed. As a result, the av- erage number density of the selected stars is about 1/arcmin2, and there are on average about 60 stars in each chip. How- ever, the chip-by-chip variation in the number of stars is quite large; some chips have only ∼20 stars because of the presence of a large saturated star in the field. We make the first order obs as a function of posi- bi-polynomial fit to values of (P∗ tion (a second order fit does not change the results, see the Ap- pendix). To make this fit, we use the efit command in imcat, and flux weighting is not applied. This function is used in eq. (15) to correct the ellipticities of faint galaxies. This correction is made with the ecorrect command in imcat. Note that each CCD chip is treated independently in this procedure. We found that a first order fit corrects the PSF anisotropy well, and fur- thermore does not introduce a systematic artificial residual due to the wings of a higher order fit (Van Waerbeke et al. 2002), provided that overlapped CCD regions are masked. The RMS value of the ellipticities of the reference stars, he∗2i1/2, is re- duced from 2.8% to 1.0% as a result of the correction. Note that the RMS before the correction is already small, because of the superb image quality of the Subaru telescope. Figure 3 shows the star ellipticities before (left panel) and after (right panel) the anisotropic PSF correction from one pointing. The observed ellipticities not only have a large scatter but also show a systematic effect. After the PSF correction this tendency is removed and the corrected star ellipticities are distributed sym- metrically around zero with a small scatter. sm)- 1e∗ FIG. 3. -- Ellipticities distribution of stars, before (left) and after (right) the correction for PSF anisotropies is made. The mean and dispersion of the ellipticites (e) of all the stars are shown inside the frame. Cosmic shear statistics in the Suprime-Cam 2.1 sq deg field 5 of a certain observable, hAi, is calculated asPN instead of PN where α2 is the variance of all the objects in the catalog and α ≃ 0.4 here. Under the weighting scheme, an averaged value i=1 wi i=1 Ai/N. The lower panel of Figure 2 shows the averaged weight as a function of Rc magnitude. As expected, less weight is given to fainter objects because the shape mea- surements of these objects are noisier. The dashed histogram in the upper panel of Figure 2 plots the product of the num- ber counts and the average weight and shows that the effective counts peak at Rc ∼ 24.5. i=1 wiAi/PN FIG. 4. -- Components of galaxy ellipticities as a function of the star elliptic- ity component used for the anisotropic PSF correction. Top and middle panels show the averaged galaxy ellipticities before and after correction, respectively. Bottom panels show the ellipticity distribution of galaxies (number per bin) used for the cosmic shear correlation measurement. Figure 4 shows the average ellipticity of galaxies as a func- tion of the value of the star ellipticity used for the anisotropic PSF correction. The top panels show the observed ellipticities, and the expected strong correlation is present. Note that for the majority of galaxies, the PSF anisotropy is very small, as shown in the bottom panels. The corrected ellipticities plotted in the middle panels comprise only small values. The averaged values of the corrected ellipticities are he1i = - 5.0 × 10- 4 and he2i = 1.8 × 10- 3. Thus, no significant offset is found. The second step is the isotropic correction, caused by the window, WG, and the seeing. Luppino & Kaiser (1997) reported a method to correct the ellipticities for these effects. The pre- seeing shear γ is described as FIG. 5. -- Shear correlation functions. Top panel is for ξtt, middle panel for ξrr, and bottom for the cross correlation ξtr which should vanish if the data are not contaminated by systematics. The error bars present the statistical error computed from 100 randomized realizations. 5. COSMIC SHEAR CORRELATIONS In this section, we present the cosmic shear correlations mea- sured from the 2.1 deg2 Suprime-Cam data, and discuss their statistical and possible systematic errors. The top and middle panels of Figure 5 show the shear cor- relation functions, ξtt and ξrr, respectively, computed using the estimator eqs. (8) and (9). The bottom panel presents the cross- correlation ξtr. The error bars indicate the RMS among 100 ran- domized realizations, in which the orientations of galaxies are randomized, and presumably represent the statistical error. In the following, statistical errors are computed in this manner. As the two top panels clearly show, we detect non-zero shear corre- lation signals on scales below 30 arcmin, except for the smallest two bins. On small scales (θ < 1 arcmin) the number of pairs corresponding to the separations decreases, and consequently the signals become noisy and the statistical errors become large. As the cross-correlation should be zero for a signal due to grav- itational lensing, it provides a check on systematic effects in the data. The bottom panel shows that the cross-correlation is indeed consistent with zero at all scales. This result indicates γ = P- 1 γ e, Pγ = Psh - P∗ sh P∗ sm Psm, (16) where Psh is the shear polarisability tensor. The Pγ of individual galaxies is, however, known to be a noisy estimate, and thus we adopt the smoothing and weighting method developed by Van Waerbeke et al.(2000; see also Erben et al. 2001 for a detailed study of the smoothing scheme). For each object, 20 neigh- bors are first identified in the rg-Rc plane. A median value of Pγ among these neighbors is adopted as the smoothed Pγ of the ob- ject. The averaged tr(Pγ) for all objects is calculated as 0.40 but Pγ depends on the object size. Figure 1 plots the averaged tr(Pγ) as a function of rg. This graph shows that the average tr(Pγ) is almost constant, ∼ 0.65 for rg > 3.5, but becomes smaller for smaller rg. As the estimated γ is still noisy, especially for small and faint objects, it is important to weight the galaxies according to the uncertainty in the shape measurements. We weight the objects using the procedure developed by Van Waerbeke et al. (2000; 2002; see also Erben et al. 2001). The variance of raw γ before the smoothing among the neighbors, σ2 γ, is used to estimate the weight of each object, w, as w = σ2 γ 1 + α2 (17) 6 T. Hamana et al. that our PSF corrections perform well and do not introduce a systematic effect into the data. galaxies is not sufficient to obtain meaningful statistics. This possibility should be tested in future studies. Currently, it is not clear how to correct the E mode vari- ance, given the observed B mode. We follow Van Waerbeke et al. (2002b) and add the B mode in quadrature to the uncer- tainty in the E mode for the maximum likelihood analysis (see §6). FIG. 6. -- Upper panel shows the E-mode aperture mass variance hM2 api, while the lower panel for B-mode aperture mass variance hM2 ⊥i. The error bars present the statistical error computed from 100 randomized realizations. Let us now turn to the aperture mass variance. We compute the E and B mode aperture mass variances from the two-point shear correlation functions via eqs. (11) and (12). We use the analytic expressions for the window function W and W given in Schneider et al. (2002). The shear correlation functions defined by eq. (10) are computed over the range 0.04′ < θ < 90′ on 168 bins equally spaced with a log-interval of ∆log θ = 0.02. The E and B mode aperture mass variances are plotted in Figure 6. For the E mode, we detect positive, non-zero signals on scales larger than θap > 2′. Since, as shown in Figure 5, the two-point correlation function becomes very noisy on scales smaller than 1 arcmin, and also considering that the aperture mass effec- tively probes a real scale of ∼ θap/5, we use only the signals on scales larger than θap = 2 arcmin for our maximum likelihood analysis in the next section. The amplitude and main features of the E mode variance are broadly consistent with theoretical predictions under the cluster normalized CDM cosmology (see Figure 8). The B mode aperture mass variance is shown in the lower panel of Figure 6. Small but non-zero signals are found on scales larger than 5 arcmin. Currently, the origin of this B mode variance is not clear. One possibility is incorrect anisotropic PSF correction. We tested this possibility by repeating the PSF correction using different procedures, and found no significant problems (see the Appendix). It is interesting to note that Van (2002) reported the detection in their data Waerbeke et al. ⊥i ∼ 3 × 10- 6) on scales of a non-zero B mode variance (hM2 10′ < θap < 40′, the amplitude of which is very similar to ours. Note that their survey depth (a limiting magnitude of IAB = 24.5) was slightly shallower than ours. On the other hand, Hoekstra et al. (2002b) reported a vanishing B mode variance on scales larger than 10 arcmin in their shallow data (Rc < 24). These results suggest that the current procedures for galaxy shape cor- rections become problematic for fainter objects. Unfortunately, we cannot test this possibility because the number of brighter FIG. 7. -- Cross-correlation coefficient r(θ, θ′) for the statistical noise as a function of scale θ for three scales, θ′ = 2.5 (filled circles with solid line), 6.3 (filled squares with dashed line)and 20 arcmin (filled triangles with long- dashed line). 6. LIKELIHOOD ANALYSIS In this section, we compare the measured aperture mass vari- ance to the model predictions in CDM cosmologies using max- imum likelihood analysis. We present constraints on cosmo- logical parameters obtained from the cosmic shear variance ob- served in the Suprime-Cam 2.1 degree2 field data. 6.1. Source redshift distribution To compute the theoretical prediction of the aperture mass variance using eq. (7), we need to fix the redshift distribution of the source galaxies. However, no redshift information about the galaxies in our catalog is available. Therefore, we decided to adopt a parameterized model that provides a good fit to the red- shift distribution of deep surveys (e.g., Hoekstra et al. 2002b; Van Waerbeke et al. 2002), ns(z) = β z∗(cid:19)α z∗Γ(cid:2)(1 + α)/β(cid:3)(cid:18) z exp"- (cid:18) z z∗(cid:19)β# , (18) with α = 2 and β = 1.5. The mean redshift relates to z∗ as ¯zs = z∗Γ(cid:2)(2 + α)/β(cid:3)/Γ(cid:2)(1 + α)/β(cid:3) and for these values of α and β, it gives ¯zs ≃ 1.5z∗. The median redshift is approximately given by zmed ≃ 1.4z∗. We treat the mean redshift of the distri- bution as a model parameter in the maximum likelihood analy- sis. 6.2. Maximum likelihood analysis In performing the maximum likelihood analysis, we basi- cally follow the procedure described in Van Waerbeke et al. (2002; see also Hoekstra et al. 2002b). The theoretical pre- dictions are computed in a four-dimensional space, but we re- strict the parameter space to realistic but conservative ranges: Ωm ∈ [0.1,1] (either ΩΛ = 0 or Ωm + ΩΛ = 1), σ8 ∈ [0.2,2], Γ ∈ [0.05,0.75] and ¯zs ∈ [0.3,2.5] with a sampling of 19 × 19 × 11 × 23. In what follows, we refer to this parameter range (Ωm, σ8, Γ, ¯zs) as the default prior space. The model predictions are then interpolated with an oversampling fivefold higher in each dimension. Cosmic shear statistics in the Suprime-Cam 2.1 sq deg field 7 distribution from the observer (z = 0) to high z (z ∼ 3) is gen- erated by stacking 10 snapshot outputs from the N-body sim- ulations. N-body data from the Very Large Simulation (VLS) which followed 5123 particles in a cubic box of 479h1Mpc on a side, carried out by the Virgo Consortium (Jenkins et al. 2001; Yoshida, Sheth & Diaferio 2001) are used. A ΛCDM model (Ωm = 0.3, Ωλ = 0.7 and h = 0.7) is assumed with the CDM ini- tial power spectrum computed using CMBFAST (Seljak & Zal- darriaga 1996). The multiple-lens plane ray-tracing algorithm is used to follow the light rays (Hamana & Mellier 2001, see also Jain, Seljak & White 2000 for the basic theory). The lens- ing convergence and shear is computed for 10242 pixels with a pixel size of 0.25 arcmin for a single source plane of zs = 1. We compute the covariance matrix due to the cosmic variance from 36 random mock observations (but without the intrinsic ellipticity of galaxy images) generated by the numerical exper- iment. As the ratio of the non-Gaussian to Gaussian contribu- tions to cosmic variance does not vary significantly with the underlying cosmology (Van Waerbeke et al. 2002), the cosmic variance also does not play an important role at all scales; given the large residual error from the B mode variance at large scales, we use the covariance matrix obtained from the ΛCDM numer- ical experiment regardless of the cosmological model consid- ered. Figure 8 shows the aperture mass variance with error bars from the statistical error only (the left error bar on each point) and a sum of the statistical, systematic and cosmic variance in quadrature (the right error bar). On small scales, statistical error dominates [the statistical error (left) is comparable to the total error (right)], while at large scales the systematic error domi- nates given the large B mode variance. FIG. 9. -- The gray-scale shows the ∆χ2 map on Γ and ¯zs obtained after a marginalization over Ωm ∈ [0.1, 1] and σ8 ∈ [0.2, 2]. A darker gray-scale indicates a lower ∆χ2 value (thus more likely). Contours indicate 68.3 and 95.4% confidence levels. 6.3. Results and discussion Here, we present two-parameter space constraints obtained from marginalizations over the two remaining parameters. Fig- ure 9 shows the likelihood map on the Γ-¯zs plane where the de- fault prior is applied for Ωm and σ8. This Figure clearly presents the correlation between Γ and ¯zs; that is, a flatter (steeper) power spectrum (a larger (smaller) Γ) prefers a lower (higher) ¯zs, (this was reported previously by Van Waerbeke et al. 2002). The message here is that, given a relatively narrow range of sig- nal detection (2′ < θap < 40′) with large error bars (as shown in Figure 8), the slope of the power spectrum and its normalization are degenerate. Similar considerations apply to a constraint on FIG. 8. -- The aperture mass variance. For each measurement point, the left error bars show statistical error while right error bars present a sum of the statistical error, a residual systematic error estimated from B mode vari- ance and the cosmic variance in quadrature. On smaller scales, statistical error dominates, while on larger scales residual error dominates. Comparing with them, the cosmic variance does not have a serious impact on the final error on all scales. The solid curve shows, for an illustrative example, the theoretical prediction of CDM model with Ωm = 0.3, ΩΛ = 0.7, Γ = 0.21, σ8 = 0.85 and ¯zs = 1. For a given theoretical model, we compute the χ2 (log- likelihood) χ2 = (di - mi)C- 1(di - mi)T , (19) where di is the measurement at scale θi, and mi is the corre- sponding theoretical prediction. Confidence values are com- puted from this χ2 in the standard manner. The covariance ma- trix consists of three contributions C(θi, θ j) = Cs(θi, θ j) +Cb(θi, θ j) +Ccv(θi, θ j), (20) where Cs, Cb and Ccv are the statistical error, the residual sys- tematics, and the cosmic variance, respectively, computed by the procedures described below. The statistical error was com- puted from the 100 randomized realizations catalog. Figure 7 shows the cross-correlation coefficient for the statistical noise for three scales, 2.5, 6.3, and 20 arcmin, with other scales. The non-zero B mode variance could indicate the existence of a residual systematic, although its origin is not yet clear. It would be natural to consider that a residual systematic of similar size also exists in the E mode. However, there is not yet a clear scheme to deal with this residual systematic. Therefore, we adopt the simple and conservative procedure described in Van Waerbeke (2002). We quadratically add the B mode variance to the error of the signal. As the E and B mode covariance matri- ces for the statistical noise are identical, the diagonal part of the matrix Cb is given by the B mode signal and off-diagonal terms follow the same correlation properties as the E mode. Estimation of the cosmic variance is more complicated, be- cause the observed scales are in the quasi-linear to nonlinear regime and thus random Gaussian theory cannot be applied. The cosmic variance is estimated using weak lensing numer- ical experiments, which are performed using a ray-tracing tech- nique combined with large N-body simulations; the details are described in Hamana et al. (in preparation; see also Menard et al. 2003; Takada & Hamana 2003). Briefly, the dark matter 8 T. Hamana et al. σ8-Γ, as shown in the lower left panel of Figure 10. Although the constraint on Γ and ¯zs is not as tight, we can see from Figure 9 that for ¯zs > 0.9 (> 0.7) a flatter power spectrum with Γ > 0.5, would be ruled out at more than the 68% confidence level for the flat (open) model. FIG. 10. -- The gray-scale shows the ∆χ2 on two-parameter space obtained after a marginalization over two remaining parameters for the default prior, a darker gray-scale for a lower ∆χ2 value (thus more likely). Plots are for σ8-¯zs (upper left), Ωm-¯zs (upper right), σ8-Γ (lower left) and Ωm-Γ (lower right). Contours indicate 68.3, 95.4 and 99.73% confidence levels. FIG. 11. -- Constrains on Ωm and σ8 obtained after a marginalization over Γ ∈ [0.1, 0.4] and ¯zs ∈ [0.6, 1.4]. Contours indicate 68.3, 95.4 and 99.73% confidence levels. Figure 10 shows the degeneracy among the four parameters. The left panels clearly show the strong degeneracy between σ8 and ¯zs (or Γ). From these plots the following two points be- come clear: (i) redshift information about the source galaxies is crucial to obtaining a tight constraint on σ8 (Bernardeau et al. 1997; Jain & Seljak 1997). (ii) A constraint on Γ from other in- dependent observations (such as galaxy clustering and/or CMB anisotropies) is very useful to break the degeneracy among the parameters. The right panels of Figure 10 show the same likeli- hood maps but for Ωm. The constraints on Ωm are weaker than those on σ8, because the cosmic shear correlation is less sen- sitive to Ωm than σ8 (Bernardeau et al. 1997; Jain & Seljak 1997). In addition to the default prior, we adopt other priors from information obtained from other observations. We estimate the mean redshift of our galaxy catalog using redshift distributions from other deep surveys. The redshift distributions of faint galaxies in the Hubble Deep Fields North and South have been estimated using spectroscopic and photometric redshift tech- niques by several groups (Fernández-Soto et al. 1999; Cohen et al. 2000; Fontana et al. 1999; 2000; Yahata et al. 2000). Note that there is a small discrepancy between the two datasets, which is probably due to field-to-field variation. If we assume that our galaxy catalog has a similar redshift distribution to the Hubble Deep Field data, the mean redshift of our catalog (22.5 < Rc < 26) is estimated to be larger than z = 1. On the other hand, the redshift distribution of galaxies in the Subaru deep survey field has been estimated using the photometric red- shift technique with B V Rc i′ z′ data (Furusawa 2002; Furu- sawa et al. in preparation, see also Kashikawa et al. 2003). They found that the mean redshift of faint galaxies (R > 24) is systematically lower than the HDF data. Note that the discrep- ancy is not very significant, given the large error bars due to small-number statistics. If we take the redshift distribution of the Subaru deep survey data, the median redshift of our cata- log can be as low as z = 0.7. The difference between these two estimates suggests that the field-to-field variation can be quite large. It is also important to take into account that the weight- ing of galaxies might change the redshift distribution in a non- trivial manner. Considering these points, we adopt a conserva- tive constraint of ¯zs ∈ [0.6,1.4], and two tighter constraints of ¯zs ∈ [0.6,1.2] and ¯zs ∈ [0.8,1.4] to determine the impact of the source redshift information. We constrain Γ using the studies of galaxy clustering by the SDSS (Dodelson et al. 2002; Szalay et al. 2001) and the APM survey (Eisenstein & Zaldarria 2001). As there is a wide dispersion in Γ values among these stud- ies, whereas the statistical error bars in each measurement are small, we adopt a conservative constraint, Γ ∈ [0.1,0.4], which covers the 68% confidence intervals of these three studies. We also take an extreme constraint of Γ = 0.21, where the 68% con- fidence intervals of the three studies overlap, to determine how well independent information on Γ improves the constraint on Ωm and σ8. Figure 11 shows the constraint on Ωm and σ8 obtained from marginalization over Γ ∈ [0.1,0.4] and ¯zs ∈ [0.6,1.4]. As ex- pected, a strong degeneracy between Ωm and σ8 is found. The confidence intervals of Ωm and σ8 from various priors are sum- marized in Table 1. It is clear from this Table that informa- tion about the redshift distribution of galaxies and Γ can indeed give a tighter constraint on Ωm and σ8. Although the current constraint is not very tight, we may rule out the following two models.The COBE normalized high density CDM model (Bunn & White 1997, i.e., Ωm = 1, Ωλ = 0, σ8 = 1.2) is ruled out at more than the 99.9% confidence level. This model predicts too high an amplitude at all scales. For the open model, low density models with Ωm < 0.2 are ruled out at more than the 68% con- fidence level. As pointed out by Van Waerbeke et al. (2002; see also Schneider et al. 1998), the incompatibility is because such low density models predict very large power at small scales. We obtain the following fitting function for the 95% confidence contours plotted in Figure 11, σ8 = (0.50+0.35 - 0.16)Ω - 0.37 m , for Ωm + ΩΛ = 1, (21) Cosmic shear statistics in the Suprime-Cam 2.1 sq deg field 9 95.4% ONE-(TWO-)PARAMETER CONFIDENCE INTERVALS OBTAINED FROM THE APERTURE MASS VARIANCE, ∆χ TABLE 1 2 = 4.00 (6.17), OF Ωm AND σ8 FOR DIFFERENT PRIORS. (...) INDICATES THAT NO CONSTRAINT IS PLACED WITHIN THE INTERVALS 0.1 < Ωm < 1.0 AND 0.2 < σ8 < 2.0. Priors Γ 0.21 hzsi ∈ [0.6,1.4] ∈ [0.05,0.75] ∈ [0.6,1.4] ∈ [0.1,0.4] ∈ [0.6,1.4] ∈ [0.6,1.2] ∈ [0.05,0.75] ∈ [0.6,1.2] ∈ [0.1,0.4] ∈ [0.6,1.2] ∈ [0.8,1.4] ∈ [0.05,0.75] ∈ [0.8,1.4] ∈ [0.1,0.4] ∈ [0.8,1.4] 0.21 0.21 Ωm ... (...) ... (...) ... (...) ... (...) ... (...) ... (...) ... (...) ... (...) ... (...) Ωm + ΩΛ = 1 σ8 ΩΛ = 0 Ωm σ8 > 0.22 (...) > 0.29 (> 0.26) 0.33 -- 1.92 (> 0.31) > 0.24 (> 0.21) > 0.31 (> 0.29) > 0.37 (> 0.23) > 0.41 (> 0.26) > 0.43 (> 0.26) > 0.36 (> 0.24) > 0.40 (> 0.25) > 0.43 (> 0.28) > 0.36 (> 0.21) 0.28 -- 1.44 (0.25 -- 1.56) > 0.39 (> 0.22) 0.32 -- 1.29 (0.31 -- 1.39) > 0.41 (> 0.23) 0.35 -- 1.92 (> 0.33) 0.21 -- 1.65 (< 1.75) 0.22-0.93 (< 1.03) 0.29 -- 0.79 (0.26 -- 0.86) 0.33 -- 0.70 (0.31 -- 0.76) 0.23-0.93 (0.21 -- 1.03) 0.31 -- 0.79 (0.29 -- 0.87) 0.35 -- 0.70 (0.33 -- 0.76) 0.21 -- 0.76 (< 0.81) 0.27 -- 0.63 (0.25 -- 0.68) 0.32 -- 0.55 (0.31 -- 0.61) and σ8 = (0.51+0.20 - 0.16)Ω - 0.34 m , for ΩΛ = 0. (22) FIG. 12. -- The one-dimensional likelihood function (∆χ2) for σ8 in the flat ΛCDM cosmological model with Ωm = 0.3 (ΩΛ = 0.7) and Γ = 0.21 for priors ¯zs ∈ [0.6, 1.2] (the solid line) and ¯zs ∈ [0.8, 1.4] (dashed line). The horizontal dotted lines indicate, form lower to upper, 68.3, 95.4 and 99.73% confidence levels. Figure 12 shows the one-dimensional likelihood function for σ8 in the currently popular flat ΛCDM cosmological model with Ωm = 0.3 (ΩΛ = 0.7) and Γ = 0.21. The solid curve is for the prior ¯zs ∈ [0.6,1.2], while the dashed curve is for ¯zs ∈ [0.8,1.4]. Specifically, the one-dimensional confidence intervals of σ8 for the 95.4% level are: 0.68 < σ8 < 1.33 for ¯zs ∈ [0.6,1.2], 0.62 < σ8 < 1.02 for ¯zs ∈ [0.8,1.4], and 0.62 < σ8 < 1.32 for ¯zs ∈ [0.6,1.4]. A strong degeneracy be- tween σ8 and ¯z is evident. Especially, the upper bound of the σ8 confidence interval is very sensitive to the choice of the lower limit of the mean redshift. This is also seen in the upper left panel of Figure 10. From these results we may say that, for the flat ΛCDM cosmological model, we obtain a relatively reliable lower bound of σ8 > 0.62 (95.4% confidence) for a reasonable choice of ¯zs. However, the upper limit is uncertain; it depends strongly on the choice of the mean redshift. A conservative conclusion is σ8 < 1.33 (95.4% confidence), but it is tightened by σ8 < 1.02 if the mean redshift is as large as ¯zs = 0.8. Our results are broadly consistent with constraints obtained from other cosmic shear surveys (Maoli et al. 2001; Bacon et al. 2002; Van Waerbeke et al. 2002; Hoekstra et al. 2002b; Re- fregier, Rhodes & Groth 2002; Brown et al. 2003; Jarvis et al. 2003). This consistency is remarkable, given that the data have been compiled from different instruments, filters and sur- vey depths. However, there is a spread of confidence intervals among the surveys. In fact, Bacon et al.(2002) and Van Waer- beke et al. (2002) obtained a slightly higher normalization, σ8 ∼ 0.95 for the ΛCDM model, which is incompatible with our 68.3% confidence interval for ¯zs ∈ [0.8,1.4]. Note that Bacon et al.(2002) did not decompose the shear correlation function into E and B modes. Thus, their correlation function could be biased on the high side. The source of the small discrepancy among the cosmic shear surveys is unclear: it could be field-to-field variance, it could arise from the different analysis schemes, or it could be due to a mis-choice of the redshift distribution of galaxies. Hirata & Seljak (2003) investigated biases induced by the conversion between the observed image shape to shear distortion in current weak lensing analyses. They found that the non-Gaussianity of the point spread function has a signifi- cant effect and can lead to up to a 15% error in σ8 depending on the method of analysis. A wider field, multi-color survey, and an analysis scheme calibrated using mock observations (as re- ported by Erben et al. 2000) are needed to improve the accuracy of the cosmic shear analysis. It is interesting to compare our results with the σ8 values obtained from the number density of rich clusters of galaxies published by many other groups. It should, however, be em- phasized that there is a large dispersion in σ8 values among these studies, whereas the errors in each measurement are quite small. The σ8 values reported range from σ8 . 0.7 (e.g., Bor- gani et al. 1999; 2001; Seljak 2002; Vianna, Nichol & Liddle 2002) to σ8 & 0.9 (e.g., Eke, Cole & Frenk, 1996; Kitayama & Suto 1996; 1997; Bahcall & Fan 1998; Pen 1998; Pierpaoli et al. 2001) for the standard ΛCDM model. This spread may reflect mainly the uncertainty in the rela- tion between the mass and X-ray temperature of clusters. If we take the prior of ¯zs ∈ [0.6,1.4], both values are well within our 95.4% confidence interval. However, larger σ8 values are not preferred by the constraint obtained from the high-¯zs prior, 10 T. Hamana et al. but are well within the 95.4% confidence interval of the low-¯zs prior. Thus, if the mean redshift of our galaxy catalog is as high as that estimated from the HDF data (¯zs & 1), our result is more in accord with the lower σ8 value. Recently, Spergel et al. (2003) combined CMB measure- ments from WMAP (Bennet et al. 2003 and references therein), CBI (Person et al. 2002) and ACBAR (Kuo et al 2002), the galaxy power spectrum from the 2dF galaxy redshift survey (Percival et al. 2001; Verde et al. 2002), and the measurements of the Lyman α power spectrum (Croft et al. 2002; Gnedin & Hamilton 2002) to find the best fit cosmological model and ob- tained σ8 = 0.84 ±0.04 (68% confidence). This result is in good agreement with our cosmic shear constraints. 7. SUMMARY We analyzed the Suprime-Cam 2.1 deg2 Rc-band data and measured the cosmic shear correlation. Suprime-Cam has a wide field of view of 34′ × 27′, and its superb imaging quality provides a very small RMS value of star ellipticities of 2.8%, which after PSF corrections is reduced to 1.0%. These advan- tages combined with the large light gathering power of the 8.2- m Subaru telescope make Suprime-Cam an almost ideal camera for a weak lensing survey. For the cosmic shear correlation measurement we used galaxies with 22.5 < Rc < 26 and an image size larger than the seeing size. We detected a non-zero cosmic shear two-point correlation function of up to 40′. However, this result may be contaminated by shear that is not caused by gravitational lens- ing. We thus adopted the aperture mass variance, which nat- urally decomposes the correlation signal into E and B modes (the latter arises from shear whose origin is not in gravitational lensing). We detected a non-zero E mode variance on scales from θap = 2′ to 40′. As the aperture mass probes a scale of θap/5, the signals come from effective scales of 0.5′ < θ < 10′, cor- responding to the quasi-linear to nonlinear regimes. We also detected a small but non-zero B mode variance on scales larger than θap > 5′. Currently, the origin of this B mode variance is not clear. One possibility is an incorrect anisotropic PSF correction. To test this possibility, we repeated the anisotropic PSF correction using different procedures (see the Appendix for details), namely: (i) a second order bi-polynomial fit to PSF, (ii) a pointing-by-pointing correction without masking over- lapped regions and (iii) using fainter stars for modeling the PSF anisotropy. We did not find significant problems in our PSF correction procedure. Interestingly, the amplitude of the B mode variance on larger scales is similar to that found by Van Waerbeke et al. (2002), though their survey depth was slightly shallower than ours. On the other hand, Hoekstra et al. (2002b) found a vanishing B mode variance on scales larger than 10 arcmin in their shal- low data (Rc < 24). These results may suggest that the current procedures for galaxy shape correction become problematic for fainter objects. Future detailed studies of the origin of the B mode shear are required to understand and suppress this possi- ble source of residual systematic error. Also, a calibration of the analysis scheme using mock data is needed to improve the accuracy of the cosmic shear analysis. - 0.34 m - 0.16)Ω - 0.16)Ω We performed a maximum likelihood analysis in a four- dimensional space of σ8, Ωm, Γ and ¯zs. We included three possible sources of error: the statistical noise, the cosmic vari- ance, and the residual systematic estimated from the B mode variance. We derived joint constraints on two parameters by marginalizing over the two remaining parameters. We ob- tained a weak upper limit of Γ < 0.5 for ¯zs > 0.9 (68% con- fidence). We also showed that independent information on Γ can reduce the degeneracy among the parameters. For the prior Γ ∈ [0.1,0.4] and ¯zs ∈ [0.6,1.4], we found σ8 = (0.50+0.35 - 0.37 m for Ωm + ΩΛ = 1 and σ8 = (0.51+0.29 for ΩΛ = 0 (95.4% confidence). Although the current constraint is not very tight, we can rule out the following two models: the COBE normal- ized high density CDM model (Ωm = 1, Ωλ = 0, σ8 = 1.2) by more than a 99.9% confidence level, and low density open mod- els (Ωm < 0.2) by more than 68% confidence. If we take the currently popular ΛCDM model (Ωm = 0.3, Ωλ = 0.7, Γ = 0.21), we obtain a one-dimensional confidence interval on σ8 for the 95.4% level, 0.62 < σ8 < 1.32 for ¯zs ∈ [0.6,1.4]. This result is broadly consistent with constraints from other cosmic shear sur- veys and from the cluster abundance. However, we found that the confidence interval is sensitive to the choice of the mean redshift: 0.68 < σ8 < 1.33 for the prior of ¯zs ∈ [0.6,1.2], while 0.62 < σ8 < 1.02 for ¯zs ∈ [0.8,1.4]. The latter is incompati- ble with the higher σ8 values obtained from some cluster abun- dance studies. This result clearly demonstrates that informa- tion on the redshift distribution of the source galaxies is cru- cial and can significantly tighten the confidence interval of σ8 and Ωm. The improvement of the constraint on σ8 from the redshift information can be estimated as follows: the cosmic shear correlation roughly scales with σ8 and the mean redshift as ξ ∝ σ2.5 , thus the uncertainly in the median redshift con- tributes to the error in σ8 as δσ8/σ8 = 0.6δzs/zs. Therefore, the error in σ8 due to uncertainly in the median redshift can be reduced to 10% by the current photometric redshift technique (e.g., Bolzonella, Miralles & Pelló 2000). 8 z1.5 s We would like to thank Y. Mellier for useful discussions and comments on the manuscript and L. Van Waerbeke for help- ful discussions about galaxy shape analysis. We also thank the anonymous referee for detailed and constructive comments on an earlier manuscript, which improved the paper. T.H. thanks K. Umetsu and T. Futamase for useful discussions. T.H. F.N. and M.O. acknowledge support from Research Fellowships of the Japan Society for the Promotion of Science. REFERENCES Bacon, D. J., Refregier, A. R., & Ellis, R. S. 2000, MNRAS, 318, 625 Bacon, D. J., Massey, R. J., Refregier, A. R., & Ellis, R.S. 2002, submitted to MNRAS (astro-ph/0203134) Bahcall, N. A., & Fan, X., 1998, ApJ, 504, 1 Bardeen, J., Bond, J., Kaiser, N., & Szalay, A. S., 1986, ApJ, 304, 15 Bartelmann, M., & Schneider, P., 2001, Phys. Rep., 340, 291 Bennett, C. L., et al., 2003, ApJ, submitted (astro-ph/0302207) Blandford, R. D., Saust, A. B., Brainerd, T. G., & Villumsen, J. V., MNRAS, Borgani, S., Rosati, P., Tozzi, P., & Norman, C., 1999, A&A, 363, 476 Borgani, S., et al. 2001, ApJ, 561, 13 Brown, M. L., Taylor, A. N., Bacon, D. J., Gray, M. E., Dey, S., Meisenheimer, K., Wolf, C., 2003, MNRAS, 341, 100 Bunn, E. F., & White, M., 1997, ApJ, 480, 6 Cohen, J. G., Hogg, D. W., Blandford, R., Cowie, L. L., Hu, E., Songaila, A., Shopbell, P., Richberg, K., 2000, ApJ, 538, 29 Crittenden, R. G., Natarajan, P., Pen, U-L., & Theuns, T., 2002, ApJ, 568, 20 Croft, R. A. C., Weinberg, D. H., Bolte, M., Burles, S., Hernquist, L., Katz, N., 251, 600 Bolzonella, M., Miralles, J.-M., & Pelló, R. 2000, A& 363, 476 Kirkman, D., & Tytler, D. 2002, ApJ, 581, 20 Cosmic shear statistics in the Suprime-Cam 2.1 sq deg field 11 Dodelson, S., et al., (the SDSS collaboration), 2002, ApJ, 572, 140 Eisenstein, D. J., & Zaldarria, M., 2001, ApJ, 546, 2 Eke, V. R., Cole, S., & Frenk, C. S., 1996, MNRAS, 282, 263 Erben, T., van Waerbeke, L., Bertin, E., Mellier, Y. & Schneider, P., 2001, A&A, 366, 717 Fernández-Soto, A., Lanzetta, K. M., & Yahil, A., 1999, ApJ, 513, 34 Fontana, A., D'Odorico, S., Fosbury, R., Giallongo, E., Hook, R., Poli, F., Renzini, A., Rosati, P., & Viezzer, R., 1999, A&A, 343, L19 Fontana, A., D'Odorico, S., Poli, F., Giallongo, E., Arnouts, S., Cristiani, S., Moorwood, A., & Saracco, P., 2000, AJ, 120, 2206 Furusawa, H., 2002, Ph.D. thesis, Tokyo University Gnedin, N. Y. & Hamilton, A. J. S. 2002, MNRAS, 334, 107 Hamana, T. & Mellier, Y. 2001, MNRAS, 169, 176 Hirata, C., & Seljak, U., 2003, MNRAS in press (astro-ph/0209489) Hoekstra, H., Yee, H. K. C., Gladders, M. D., Barrientons, L. F., Hall, P. B., & Infante, L., 2002a, ApJ, 572, 55 Hoekstra, H., Yee, H. K. C., & Gladders, M. D., 2002b, ApJ, 577, 595 Hu, W., 1999, ApJ, 522, L21 Jain, B., & Seljak, U., 1997, ApJ, 484, 560 Jain, B. Seljak, U. & White, S. D. M. 2000, ApJ, 530, 547 Jarvis, M., Bernstein, G. M., Fischer, P., Smith, D., Jain, B., Tyson, J. A., Wittman, D. 2003, AJ, 125, 1014 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, 324, 450 Kaiser, N., 1992, ApJ, 388, 272 Kaiser, N., Squires, G. & Broadhurst, T. 1995, ApJ, 449, 460 Kaiser, N., Wilson, G., Luppino, G., Dahle, H. 1999, PASP, submitted (astro- ph/9907229) Kashikawa, N., et al., 2003, AJ, 125, 53 Kitayama, T., & Suto, Y., 1996, ApJ, 469, 480 Kitayama, T., & Suto, Y., 1997, ApJ, 490, 557 Kuo, C. L. et al. 2002, ApJ, submitted (astro-ph/0212289) Luppino, G.A. & Kaiser, N. 1997, ApJ, 475, L20 Maoli, R., et al., 2001, A&A, 367, 766 Mellier, Y. 1999, ARA&A, 37, 127 Ménard, B., Hamana, T., Bartelmann, M., & Yoshida, N., 2003, A&A, 403, 817 Miralda-Escude, J., 1991, ApJ, 380, 1 Miyazaki, S., Hamana, T., Shimasaku, K., Furusawa, H., Doi, M., Hamabe, M., Imi, K., Kimura, M., Komiyama, Y., Nakata, F., Okada, N., Okamura, S., Ouchi, M., Sekiguchi, M., Yagi, M., & Yasuda, N., 2002, ApJ, 580, L97 Miyazaki, S., Komiyama, Y., Okada, N., Imi, K., Yagi, M., Yasuda, N., Sekiguchi, M., Kimura, M, Doi, M., Hamabe, M., Nakata, F., Shimasaku, K., Furusawa, H., Ouchi, M. & Okamura, S., 2002, PASJ, 54, 833 Navarro, J., Frenk, C. & White, S. 1996, ApJ, 462, 563 Peacock, J. A., & Dodds, S. J., 1996, MNRAS, 280, L19 Pearson, T. J. et al., ApJ, submitted (astro-ph/0205388) Pen, U.-L., 1998, ApJ, 504, 601 Pen, U-L., Van Waerbeke, L., Meller, Y., 2002, ApJ, 567, 31 Percival, W. J., et al., 2001, MNRAS, 327, 1297 Pierpaoli, E., Scott, D., White, M., 2001, MNRAS, 327, 1297 Press, W. & Schechter, P. 1974, ApJ, 185, 397 Refregier, A., Rhodes, J., & Groth, E. D., 2002, ApJ, 572, L131 Schneider, P., van Waerbeke, L., Jain, B. & Kruse, G. 1998, MNRAS, 296, 873 Schneider, P., van Waerbeke, L.& Mellier, 2002, A&A, 389, 729 Seljak, U., 2002, MNRAS, 337, 769 Seljak, U. & Zaldarriaga, M. 1996, ApJ, 469, 437 Sheth R. K., Tormen G., 1999, MNRAS, 308, 119 Spergel, D. N., et al. 2003, ApJ, submitted (astro-ph/0302209) Szalay, A. S., et al.., (the SDSS collaboration), submitteed to ApJ, (astro- ph/0107419) Takada, M., & Hamana, T., 2003, submitte to ApJ, (astro-ph/0305381) Van Waerbeke, L. et al., 2000, A&A, 358, 30 Van Waerbeke, L. et al., 2001, A&A, 374, 769 Van Waerbeke, L. et al., 2002, A&A, 393, 369 Verde, L., et al.. 2002, MNRAS, 335, 432 Vianna, P. T. P., Nichol, R. C., & Liddle, A. R., 2002, ApJ, 569, L75 Wittman, D. M., Tyson, J. A., Kirkman, D., Dell'Antonio, I., & Bernstein, G., 2000, Nat, 405, 143 Yahata, No., Lanzetta, K. M., Chen, H.-W., Fernández-Soto, A., Pascarelle, S. M., Yahil, A., & Puetter, R. C., 2000, ApJ, 538, 493 Yoshida, N., Sheth, R. K. & Diaferio, A. 2001, MNRAS, 328, 669 APPENDIX ANISOTROPIC PSF CORRECTION In §5, we found a small but non-zero B mode aperture mass variance on scales larger than 5 arcmin. Currently, the origin of this B mode variance is not clear. One possibility is an incorrect anisotropic PSF correction. To test this possibility, we repeated the anisotropic PSF correction, but adopting different procedures: 1. use a second order bi-polynomial fit to (P∗ sm)- 1e∗ obs. 2. use the pointing-by-pointing correction without masking the overlapping regions. 3. use fainter stars for modeling the PSF anisotropy. First, we repeated the anisotropic PSF correction adopting a second order bi-polynomial fit (the primary analysis uses a first order fit). In this case, we found that both the E and B mode aperture mass variances are almost identical to the results from our primary data, the differences are 1 × 10- 6 at largest. Further, higher order fits are not feasible, because of the small number of stars in some chips. Note that Van Waerbeke et al. (2002) reported that a higher order polynomial fit to the PSF (third order in their case) caused a wing at the edge of fields and produced an artificial B mode signal. Second, we applied the anisotropic PSF correction not to each chip separately but to each pointing, which is composed of ten chips (see Miyazaki et al. 2002 for instrumental details of Suprime-Cam). In this case, we did not mask the overlapping regions between different CCD chips that result from stacking dithering exposures. The second and fifth order bi-polynomial fits were adopted. No significant differences were found in either the E or B mode variances between the second and fifth order corrections. The E mode variance is consistent with our primary data (plotted in Figure 6). The amplitude of B mode variance is also similar to the primary data, but there is a bump at 6′ < θap < 15′. This scale is translated into a real scale of 1′ < θ < 2′, which corresponds to the dithering angles between exposures. Thus, it is very likely that the bump arises from an inaccurate PSF correction at the overlapping regions where the PSF anisotropy pattern becomes very irregular9. Because of this result, we decided to mask the overlapping regions. Also, we decided to adopt the chip-by-chip correction to avoid poor modeling of the anisotropic PSF due to discontinuities between the chips. Finally, we repeated the PSF correction but adopted slightly fainter stars to test the possibility of different responses to the PSF between bright and faint stars. Stars in the magnitude range 21.5 < Rc < 23.5 were used for the PSF correction with the first order fit (20.6 < Rc < 23.0 for the primary procedure). The number of stars selected is almost the same as the primary selection (∼ 1/arcmin2). Both the E and B mode variance from these data are consistent with the primary data (specifically the results are within the error bars of the primary data). A much fainter criterion for star selection gives a poor PSF model because of contamination by small galaxies. Thus, it gives a very poor PSF correction. 9 A similar B mode excess is found in CFHT data (L. Van Waerbeke & Y. Mellier, private communication). 12 T. Hamana et al. In conclusion, as far as can be determined from the tests, we did not find a significant problem with our PSF correction procedure. A future detailed examination of the PSF correction method should be carried out using realistic simulation data, similar to the exercise performed by Erben et al. (2001). These tests, however, are beyond the scope of this paper and will be reported elsewhere.
astro-ph/0312524
1
0312
2003-12-19T12:48:04
Radiative transfer through the Intergalactic Medium
[ "astro-ph" ]
We use a probabilistic method to compute the propagation of an ionization front corresponding to the re-ionization of the intergalactic medium in a LCDM cosmology, including both hydrogen and helium. The effects of radiative transfer substantially boost the temperature of the ionized gas over the case of uniform re-ionization. The resulting temperature-density relation of the ionized gas is both non-monotonic and multiple-valued, reflecting the non-local character of radiative transfer and suggesting that a single polytropic relation between local gas density and temperatue is a poor description of the thermodynamic state of baryons in the post-reionization universe.
astro-ph
astro-ph
Preprint-03 Radiative transfer through the Intergalactic Medium James Bolton1, Avery Meiksin2, Martin White3 1Institute of Astronomy, University of Cambridge, The Observatories, Madingley Road, Cambridge CB3 0HA, UK 2Institute for Astronomy, University of Edinburgh, Blackford Hill, Edinburgh EH9 3HJ, UK 3Departments of Astronomy and Physics, University of California, Berkeley, CA 94720, USA 31 October 2018 ABSTRACT We use a probabilistic method to compute the propagation of an ionization front corresponding to the re-ionization of the intergalactic medium in a ΛCDM cosmology, including both hydrogen and helium. The effects of radiative transfer substantially boost the temperature of the ionized gas over the case of uniform re-ionization. The resulting temperature-density relation of the ionized gas is both non-monotonic and multiple-valued, reflecting the non-local character of radiative transfer and suggesting that a single polytropic relation between local gas density and temperatue is a poor description of the thermodynamic state of baryons in the post-reionization universe. Key words: methods: numerical -- intergalactic medium -- quasars: absorption lines 1 INTRODUCTION Hydrodynamical simulations of structure formation in the universe have led to fundamental insights into the struc- ture and evolution of the Intergalactic Medium (IGM) (Cen et al. 1994; Zhang, Anninos & Norman 1995; Hernquist et al. 1996; Zhang et al. 1997; Bond & Wadsley 1997; Theuns, Leonard & Efstathiou 1998). Comparisons with the Lyα forest, as measured in high redshift Quasi-Stellar Ob- ject (QSO) spectra, show that the simulations broadly repro- duce the statistical properties of the IGM. Precision compar- isons with the highest spectral resolution measurements re- cover the cumulative flux distributions and H I column den- sity distributions to an accuracy of a few percent (Meiksin, Bryan & Machacek 2001). More problematic are the pre- dicted widths of the absorption features, which appear to require additional sources of broadening, perhaps late He II re-ionization, to match the measured widths (Theuns et al. 1999; Bryan & Machacek 2000; Meiksin et al. 2001). Many of the properties of the IGM may be attributed to the gravitational instability of the dark matter alone, as the baryon density fluctuations closely follow those of the dark matter (Zhang et al. 1998). This has led to the development of pseudo-hydrodynamic simulations based on pure gravity in which the baryon fluctuations are assumed to exactly fol- low the dark matter and the temperature to be derived from an effective equation of state (Petitjean, Mucket & Kates 1995; Croft et al. 1998; Gnedin & Hui 1998; Meiksin & White 2001). In recent years, such simulations have been in- creasingly relied on for predicting the flux power spectrum of the Lyα forest (Meiksin & White 2001; Zaldarriaga, Hui & Tegmark 2001; Croft et al. 2002; Meiksin & White 2003b; Seljak, McDonald & Makarov 2003). c(cid:13) 0000 RAS The simulations have generally been done assum- ing sudden early homogeneous H I and He II re-ionization. While tentative steps have been taken to introduce ra- diative transfer into the simulations using approximate schemes (Abel, Norman & Madau 1999; Gnedin & Abel 2001; Nakamoto, Umemura & Susa 2001; Ciardi, Stoehr & White 2003), these simulations have emphasized the prop- agation of H I ionization fronts and the predicted mean op- tical depths. Except for the simulations of Gnedin & Abel, they have not solved fully self-consistently for the combined gas dynamics and radiative transfer. In this Letter, we extend the photon-conserving scheme of Abel et al. to include the treatment of helium and ex- plore the thermal effects of radiative transfer in the inhomo- geneous medium predicted in a ΛCDM cosmology. 2 RADIATIVE TRANSFER 2.1 Simulation data We use the data from a previously run ΛCDM simula- tion used to investigate effects of radiative transfer on the Lyα forest (Meiksin & White 2003b). The simulation was run using a pure Particle Mesh (PM) dark matter code, and it was assumed the gas and dark matter have the same spa- tial distribution. A description of the parallel PM code is given in Meiksin & White (2003a). The parameters used for the simulation are ΩM = 0.30, ΩΛ = 0.70, Ωb = 0.045, h = H0/100 km s−1 Mpc−1 = 0.70, and slope of the pri- mordial density perturbation power spectrum n = 1.05. The model is consistent with existing large-scale structure, Lyα forest flux distribution, cluster abundance and WMAP constraints (Meiksin & White 2003b). The simulation was 2 J. Bolton, A. Meiksin and M. White Figure 1. The baryon density fluctuations at z = 6 against co- moving distance from the ionizing source for the line of sight considered in our radiative transfer simulations. run using 5123 particles and a 10243 force mesh, in a cubic box with (comoving) side length 25 h−1Mpc, adequate for obtaining converged estimates of the Lyα pixel flux distri- bution and flux power spectrum (Meiksin & White 2003a,b). A typical line-of-sight density run at z = 6, used for the ra- diative transfer calculations here, is shown in Fig. 1. 2.2 Equations of radiative transfer To consider the effect of radiative transfer when modelling the propagation of ionization fronts (I-fronts) into the IGM, we extend the photon-conserving algorithm of Abel et al. to include helium. The advantage of this scheme is that energy is conserved independently of numerical resolution, ensuring that I-fronts propagate at the correct speeds in our simula- tions. In practice, this means that larger step sizes may be taken on the simulation space grid without the associated loss of accuracy which would occur when solving the radia- tive transfer equation through direct numerical integration. The probabilities for the absorption of an ionizing pho- ton by H I , He I and He II , respectively, are abs = pHIqHeIqHeII (cid:2)1 − exp(cid:0)−τ tot P HI abs = qHIpHeIqHeII (cid:2)1 − exp(cid:0)−τ tot P HeI = qHIqHeIpHeII (cid:2)1 − exp(cid:0)−τ tot P HeII ν (cid:1)(cid:3) /D, ν (cid:1)(cid:3) /D, ν (cid:1)(cid:3) /D. abs (1) (2) (3) ν = τ HI ν + τ HeI ν), qi = exp(−τ i Here we have defined the auxiliary absorption and transmis- sion probabilities pi = 1 −exp(−τ i ν), i denot- ing the species being referred to, τ tot ν + τ HeII , where τ i ν is the optical depth for a given species, and D = pHIqHeIqHeII + qHIpHeIqHeII + qHIqHeIpHeII. These prob- abilities can be used to calculate the ionization rate for a given species per unit volume, niΓi, as follows. If in a time δt, δt N l−1 photons enter grid zone l from zone l −1, then the number of photons that will be absorbed in zone l by species i is δt N l−1 abs is the absorption probability for species i within zone l. The ionization rate of species i in zone l of volume V l is then abs, where P i ν P i ν ν nl iΓl i = 1 V l X g Nνg P i abs(νg), (4) L and 10ν HI where we have divided the frequencies into 100 discrete groups g evenly spaced between ν HI L . (While we found this spacing to be adequate, we have made no attempt to optimize it.) These are used in the following set of cou- pled equations to solve for the positions of the three I-fronts over time: dnHII = nHIΓHI − nenHIIαHI(T ) − 3 nHII, a a (5) dt dnHeII dt dnHeIII dt = nHeIΓHeI + nenHeIIIαHeII(T ) − nHeIIΓHeII − nenHeIIαHeI(T ) − 3 a a nHeII, = nHeIIΓHeII − nenHeIIIαHeII(T ) − 3 (6) a a nHeIII, (7) where ni denotes number density, Γi (s−1) is the photoion- ization rate per atom, αi(T ) is the total radiative recombina- tion coefficient, and a is the cosmological expansion factor. The time evolution of the gas temperature T is given by: dT dt = 2(G − L) 3kn + T n dne dt − 3 a a (cid:16) 2 3 T + ne n T(cid:17) . (8) where n = nHI + nHeI + nHII + nHeII + nHeIII + ne, ne = nHII + nHeII + 2nHeIII, k is the Boltzmann constant, and G and L (J m−3 s−1) are the atomic heating and cooling rates, respectively. The last term is the adiabatic cooling term resulting from cosmological expansion, which will dom- inate the thermal effects of gas motions at the densities we consider. While computing the photoionization, we as- sume the gas over-density (not the density) stays frozen, which is a good approximation on the scales relevant to the Lyα forest (Zhang et al. 1998). As in similar previous stud- ies, we do not include the effects of adiabatic compression heating as this will only affect the temperature for the rela- tively rare virialised halos with circular velocities exceeding ∼ 20 km s−1 (Meiksin 1994; Meiksin & White 2003b). The heating rate Gl in cell l is due to the photoioniza- tion of H I , He I and He II according to Gl = Gl HeI + Gl i is evaluated in a similar manner to the ionization rate per unit volume as given in Eq. (4), HeII, where for each species i, Gl HI + Gl nl iGl i = 1 V l X g (hνg − χi) Nνg P i abs(νg), (9) where χi is the ionization potential of species i. The atomic cooling rate L includes radiative recombination cooling to H I , He I and He II , electron excitation of H I , and Comp- ton cooling off the Cosmic Microwave Background photons. The radiative recombination and cooling rates are taken from Meiksin (1994). Eqs. (5) -- (8) are solved using explicit forward Euler in- tegration. Although an implicit numerical scheme such as backward Euler integration results in a more stable solution at low numerical resolution (Anninos et al. , 1997), we find that to obtain the required numerical accuracy as well as to maintain stability, both methods require similar numerical resolution. For the sake of speed and simplicity, the forward method is preferred. The timestep is restricted to being no more than several times greater that the hydrogen ionization timescale, Γ−1 HI , for numerical accuracy. c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Radiative transfer through the Intergalactic Medium 3 We take the frequency specific luminosity of the ionizing source in our simulations to be described by a power law spectrum with index α = 1.5, and a luminosity of LQ ν = 1023 W Hz−1 at the Lyman edge. This spectrum is typical of that of QSOs, which are probable candidate sources for contributing to the re-ionization of the IGM. The assumed mass fraction of helium in the IGM is Y ≃ 0.235. We tested the photon-conserving algorithm on the pho- toionization of gas with uniform density around a point source. The resulting solutions for the ionization and tem- perature profiles were accurate to better than 10 per cent for optical depths at the Lyman edge of up to ∆τν ≃ 20 per cell on the space grid. This results in a significant re- duction in the computational resources compared with a radiative transfer scheme in which the ionization rates are computed by direct numerical integration over the inten- ν e−τν /[4πr]2) and photoionization cross-sections (eg, sity (LQ Madau, Meiksin & Rees 1997). Typically, we find that to obtain an accuracy comparable to the photon-conserving scheme at ∆τ HI ν ≃ 20, a direct integration algorithm re- quires ∆τ HI ν ≃ 1/4 in each cell. For the simulation results presented in this Letter, the original PM grid of 1024 sepa- rate cells was used over a line of sight 25h−1 comoving Mpc in length. This provided accurate spatial resolution, as at most ∆τ HI ν ≃ 6.8 in each cell at the average baryonic den- sity assumed for the neutral IGM at z = 6. 3 RESULTS We set up a problem of a QSO source of luminosity LQ ν at z = 6 placed at a comoving distance of 10h−1 Mpc from the left edge of the density run shown in Fig. 1, and assume the gas surrounding the QSO up to that point has already been ionized. Placing the source at this distance avoids having to impose the restriction that the I-fronts propagate no faster than the speed of light. We indicate the displaced position of the source in the figures by starting the (comoving) spatial axes at R = 10h−1 Mpc. Two cases were compared: computing the ionization and temperature profiles by incorporating radiative trans- fer as discussed in section 2.2 (hereafter the RT simulation), and a second case neglecting radiative transfer, assuming instead an instantaneous uniform ionization rate across the whole line of sight (hereafter the nRT simulation), as is usu- ally done in simulations of the Lyα forest. Our intention is to compare the final gas temperatures to determine how large an effect including radiative transfer may have. Fig. 2 shows a comparison between the IGM tempera- tures at z = 5 computed with and without radiative trans- fer. The inclusion of radiative transfer results in a signifi- cant boost to the temperature of the ionized IGM, a point illustrated by Abel & Haehnelt (1999) in the context of a smooth IGM. The positions of the H II and He II I-fronts at z = 5 in the RT simulation are 24h−1 comoving Mpc from the source, while the He III front lags behind at about 19h−1 comoving Mpc; their speeds of propagation are limited by the reduction in the ionization rate due to the increasing optical depth through the box. Consequently, the heating of the IGM, which we find to be dominated by the photoion- ization of neutral hydrogen at the H II I-front, is also con- strained to lie behind the H II I-front. All gas beyond 24h−1 c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Figure 2. Comparison of the IGM temperature at z = 5 against comoving distance from the ionizing source. The solid line is com- puted using radiative transfer, while the dashed line result as- sumes a uniform ionization field through the entire line of sight. Figure 3. Comparison of the IGM temperature at z = 3 against comoving distance from the ionizing source. The line types are as in Fig. 2. Mpc is still much cooler. In contrast, the temperature com- puted by the nRT simulation is much lower and the heating extends across the whole line of sight. The temperature is also much less sensitive to variations in the gas density for the nRT simulation. Fig. 3 compares the results obtained at z = 3. By this time, both the H II and He II I-fronts have reached 35h−1 Mpc from the source. The material which has been ionized early on in the RT simulation is now beginning to cool to the same temperature as that computed by the nRT simulation, although the temperature differences re- main large. In particular, the gas that was photoionized at z = 5 (R < 24 h−1 Mpc ), remains nearly twice as hot as the uniformly ionized gas by z = 3. By contrast, the regions which were ionized more recently are much hotter than for the nRT simulation. We find that only by z = 1, once the gas has been in an ionized state for a sufficiently long period of time, do the temperatures predicted in the two scenarios converge. The reason for the temperature boost in the RT sim- ulation is made clear by considering the energy per ioniza- tion (E = G/Γ) at the I-front. In Fig. 4, we show that 4 J. Bolton, A. Meiksin and M. White Figure 4. Plot of the energy per ionization for HI (solid line) HeI (dashed line) and HeII (dot-dashed line) at z = 5. The bold lines correspond to the results computed using radiative transfer, while the lighter (flat) lines correspond to the uniform ionization field. the average energy of a photon absorbed by either H I , He I or He II increases at the relevant I-front, as proportion- ally more of the lower energy photons have been absorbed already -- a consequence of including radiative transfer. This increase to the heating energy per ionization is responsible for the temperature boost. The assumption of an instanta- neous ionization field across the line of sight results in an underestimate of the IGM temperature at high redshift. A comparison of the IGM temperature-density rela- tion computed by the two different simulations is made in Figs. 5 -- 7. Only data points where He II has been 90 per cent ionized are plotted, so that gas which is neutral or in the process of being ionized is not considered. In Fig. 5, the underdense regions in the line of sight (Fig. 1), are ini- tially heated to a higher temperature in the RT simulation. A non-monotonic temperature-density relation results. The relation actually splits into two trends. To understand the origin of the split, we divide the data points into two dis- tinct sets corresponding to two regions in which the average density fluctuations are slightly higher or lower than the overall average in the line of sight (see caption to Fig. 5). The gas in the denser region, which is closer to the source and thus is ionized sooner, shows the same overall trend as the gas in the less dense region but is slightly cooler at a fixed over-density. In contrast, in the nRT simulation the temperature-density relation is both monotonic and essen- tially single-valued. By z = 3 (Fig. 6), the ionized gas has cooled adiabatically, with the overdense regions cooling at a slower rate than the underdense regions, as the denser re- gions are better able to maintain thermal balance because of the shorter cooling (and photoionization heating) times. The IGM temperatures computed by the two simulations are now beginning to converge at high density, but at low den- sity there is still a significant difference between the results. A third set of data points is also evident, corresponding to the latest locally underdense region to be ionized. Finally, by z = 1, the gas temperatures computed by the RT and nRT simulations have nearly converged. The gra- dient of the line in the temperature-density plane indicates that adiabatic cooling due to expansion is largely responsi- Figure 5. Comparison of the IGM temperature-density relation at z = 5 for regions with fully ionized helium (nHeIII /nHe > 0.9). The triangles correspond to the uniform ionization case. The data points for the results incorporating radiative transfer are split into two sets: + corresponds to gas in the range 10h−1 < R < 14h−1, for which the mean over-density hρ/¯ρi = 1.452, (where ¯ρ is the global mean gas density); and ∗ designates gas in the range 14h−1 < R < 23.6h−1, with mean over-density 0.519. Figure 6. Comparison of the IGM temperature-density relation at z = 3. The symbols are as in Fig. 5. We also add a third point × designating gas in the range 23.6h−1 < R < 25.8h−1, with mean over-density 0.671. ble for cooling the gas. We find the entropy is now nearly constant throughout the line of sight. The striations in the RT data points have also disappeared. By z = 1, the results from the two simulations are in good agreement. 4 CONCLUSIONS We extend the probabilistic method of radiative transfer de- scribed for hydrogen by Abel et al. (1999) to include helium and tested the method on the propagation of an ionization front through a uniform medium. We compare the method against direct numerical integration of the exact time-steady radiative transfer equation. We confirm their conclusion that the probabilistic method accurately reproduces the position of the ionization fronts even for large optical depths at the photoelectric edges. Specifically, we recover the numerically c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Radiative transfer through the Intergalactic Medium 5 is insufficient in itself to account for the broader mea- sured Lyα absorption systems than predicted by numeri- cal simulations assuming sudden uniform photoionization. Meiksin, Bryan & Machacek (2001) find that a boost of ∆T ≈ 1.7×104 K is required at z < ∼ 3.5, larger than the tem- peratures we find here. Although it may be possible to ob- tain a larger boost for an alternative re-ionization scenario, such a large boost may still require invoking late He II re- ionization. Our principal conclusion here is that for numeri- cal simulations to accurately predict the temperature of the IGM, they must include the effects of radiative transfer. We are currently extending our methods to incorporate radiative transfer with ray tracing (eg, Abel & Wandelt 2002), into fully self-consistent simulations to explore this and related issues, including alternative sources of re-ionization like low luminosity AGNs and stars. A.M. thanks the University of Edinburgh Development Trust for its financial support. REFERENCES Abel T., Haehnelt M. G., 1999, ApJ, 520, L13 Abel T., Norman M. L., Madau P., 1999, ApJ, 523, 66 Abel T., Wandelt B. D., 2002, MNRAS, 330, L53 Anninos P., Zhang Y., Abel T., Norman M., 1997, NewA, 2, 209 Bond J. R., Wadsley J. W., 1997, in Petitjean P., Charlot S., eds, Structure and Evolution of the Intergalactic Medium from QSO Absorption Line Systems. Editions Fronti`eres, Paris, p. 143 Bryan G. L., Machacek M. E., 2000, ApJ, 534, 57 Cen R., Miralda-Escud´e J., Ostriker J. P., Rauch M., 1994, ApJ, 437, L9 Ciardi B., Stoehr F., White S. D. M., 2003, MNRAS, 343, 1101 Croft R.A.C., Weinberg D. H., Katz N., Hernquist L., 1998, ApJ, 495, 44 Croft R. A. C., et al. , 2002, ApJ, 581, 20 Gnedin N. Y., Abel T., 2001, NewA, 6, 437 Gnedin N. Y., Hui L., 1998, MNRAS, 296, 44 Hernquist L., Katz N., Weinberg D., Miralda-Escud´e J., 1996, ApJ, 457, L51 Madau P., Meiksin A., Rees M. J., 1997, ApJ, 475, 429 Meiksin A., 1994, ApJ, 431, 109 Meiksin A., Bryan G. L., Machacek M. E., 2001, MNRAS, 327, 296 Meiksin A., White M., 2001, MNRAS, 324, 141 Meiksin A., White M., 2003a, MNRAS, 342, 1205 Meiksin A., White M., 2003b, MNRAS (astro-ph/0307289) Nakamoto T., Umemura M., Susa H., 2001, MNRAS, 321, 593 Peebles, P.J.E, Principles of Physical Cosmology, 1993, Princeton University Press Petitjean P., Mucket J. P., Kates R. E., 1995, A&A, 295, L9 Seljak U., McDonald P., & Makarov A., 2003, MNRAS, 342, L79 Theuns T., Leonard A., Efstathiou G., 1998, MNRAS, 297, L49 Theuns T., Leonard A., Schaye J., Efstathiou G., 1999, MNRAS, 303, L58 Zaldarriaga M., Hui L., Tegmark M., 2001, ApJ, 557, 519 Zhang Y., Anninos P., Norman M. L., 1995, ApJ, 453, L57 Zhang Y., Anninos P., Norman M. L., Meiksin A., 1997, ApJ, 485, 496 Zhang Y., Meiksin A., Anninos P., Norman M. L., 1998, ApJ, 495, 63 Figure 7. Comparison of the IGM temperature-density relation at z = 1. The symbols are as in Figs. 5 and 6. integrated converged solutions for the ionization fractions and temperature to an accuracy of 10 per cent for an incre- mental optical depth per grid zone at the H I photoelectric edge as high as 20, while direct numerical integration of the radiative transfer equations requires an incremental optical depth less than 1/4 to obtain comparable accuracy. We use the probabilistic method to compute the prop- agation of an I-front generated by a QSO source through the IGM, using the density distribution drawn from a nu- merical simulation in a ΛCDM universe. Including the ef- fects of radiative transfer results in a substantial boost in the post-ionized gas temperature compared with the case of sudden uniform ionization with no radiative transfer. The boost is due chiefly to the increased energy per photoioniza- tion within the ionization fronts, a consequence of the higher probability for high energy photons to pass through the gas compared with lower energy photons. A consequence of the radiative transfer is to alter the de- pendence of temperature on density from the approximately polytropic relation found for moderate to low densities as- suming uniform photoionization (Meiksin 1994; Gnedin & Hui 1998) to a generally non-monotonic trend. While the dif- ferences between the temperatures for over-densities above ∼ 5 are small (less than 25 per cent), at lower densities the two temperatures converge only slowly with time, nearly re- covering the polytropic relation only by z ≈ 1. At z > 1, the relation between temperature and den- sity is not single-valued: multiple temperatures may result for different gas parcels of the same over-density. Although this is in part a consequence of the delay in time for gas at different distances from the source to be photoionized as the ionization front propagates, an additional contributing factor is the environment of the gas parcel: if the ionization front has passed through a region of dense gas, sufficient to drive the optical depth at the H I photoelectric edge to near unity, before reaching the parcel, the radiation field will de- liver a higher energy per photoionization to the parcel. Thus the determination of the gas temperature is strongly non- local due to the dependence of the heating rate on the optical depth to photoionizing photons between the gas parcel and the source of ionizing radiation. Although the inclusion of radiative transfer results in a substantial and persistent boost in gas temperature, it c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
0812.1984
1
0812
2008-12-10T17:51:17
Potential-density pairs for a family of finite disks
[ "astro-ph" ]
Exact analytical solutions are given for the three finite disks with surface density $\Sigma_n=\sigma_0 (1-R^2/\alpha^2)^{n-1/2} \textrm{with} n=0, 1, 2$. Closed-form solutions in cylindrical co-ordinates are given using only elementary functions for the potential and for the gravitational field of each of the disks. The n=0 disk is the flattened homeoid for which $\Sigma_{hom} = \sigma_0/\sqrt{1-R^2/\alpha^2}$. Improved results are presented for this disk. The n=1 disk is the Maclaurin disk for which $\Sigma_{Mac} = \sigma_0 \sqrt{1-R^2/\alpha^2}$. The Maclaurin disk is a limiting case of the Maclaurin spheroid. The potential of the Maclaurin disk is found here by integrating the potential of the n=0 disk over $\alpha$, exploiting the linearity of Poisson's equation. The n=2 disk has the surface density $\Sigma_{D2}=\sigma_0 (1-R^2/\alpha^2)^{3/2}$. The potential is found by integrating the potential of the n=1 disk.
astro-ph
astro-ph
Potential-density pairs for a family of finite disks Earl Schulz 60 Mountain Road, North Granby, CT 06060 [email protected] ABSTRACT Exact analytical solutions are given for the three finite disks with surface density Σn = σ0(1 − R2/α2)n−1/2 with n = 0, 1, 2. Closed-form solutions in cylindrical co- ordinates are given using only elementary functions for the potential and for the gravi- tational field of each of the disks. results are presented for this disk. The n=1 disk is the Maclaurin disk for which The n=0 disk is the flattened homeoid for which Σhom = σ0/p1 − R2/α2. Improved ΣM ac = σ0p1 − R2/α2. The Maclaurin disk is a limiting case of the Maclaurin spheroid. surface density ΣD2 = σ0(cid:0)1 − R2/α2(cid:1)3/2 The potential of the Maclaurin disk is found here by integrating the potential of the n=0 disk over α, exploiting the linearity of Poisson's equation. The n=2 disk has the . The potential is found by integrating the potential of the n=1 disk. Subject headings: gravitation- methods: analytical-celestial mechanics- galaxies: kine- matics and dynamics - galaxies: individual(NGC 891) 1. Introduction Observations of edge-on galaxies now provide detailed structural and kinematic 3-D informa- tion. However attempts to use this data to generate mass distribution have not been completely successful. One problem is that it is difficult to calculate the potential and force vectors for finite disks. There is a need for fully solved finite disks which can be used for theoretical studies, for benchmarking computer programs, and as basis functions to directly model 3-D observations. Mass modeling commonly assumes that the disk mass in contained in an infinite disk. These include the exponential disk (Freeman 1970), the Mestel disk (Mestel 1963), the Kuzmin-Toomre disk (Toomre 1963; Binney & Tremaine 2008; Evans & de Zeeuw 1992; Conway 2000), and the Rybicki disk (Evans & Collett 1993). Only a few finite disks have been solved analytically for all (R, z). Lass & Blitzer (1983) and Vokrouhlicky & Karas (1998) give a solution for a for all (R, z) for a finite thin disk with constant density. The gravitational attraction approaches infinity at the edge of this disk. The gravitational -- 2 -- attraction of the finite Mestel disk (Mestel 1963; Lynden-Bell & Pineault 1978; Hunter et al. 1984) and the truncated exponential disk (Casertano 1983) are finite at the edge of the disk but these disks have not been solved in closed form for points off the disk. Hur´e et al. (2008); Hur´e (2005) describe a method to approximate the potential of a power law disk for points both on and off the disk. The family of finite disks with surface density Σn(R; α) =(cid:0)1 − R2/α2(cid:1)n−1/2 are related to the Maclaurin spheroid which has been studied since the time of Newton. This family has recently been studied by Gonz´alez & Reina (2006) and Pedraza et al. (2008). Gonz´alez & Reina (2006) use the method of Hunter (1963) to obtain the general solution as a sum of Legendre polynomials in elliptical coordinates and give evaluated expressions for the potentials for the disks 1, 2, and 3. Here we derive complete closed form solutions in cylindrical co-ordinates for the potential and the gravitational fields of the n=0, 1, and 2 disks. We simplify the integration by moving to the imaginary domain in way which is similar to the complex-shift method introduced by Appell. See Ciotti & Marinacci (2008); Ciotti & Giampieri (2007) and references therein The gravitational potential follows Poisson's equation ∇2Φ = 4πGρ. Poissons's equation is linear so that solutions may be added. That is, if ∇2Φ1 = 4πGρ1 and ∇2Φ2 = 4πGρ2 then ∇2(Φ1 + Φ2) = 4πG(ρ1 + ρ2). Similarly, solutions may be differentiated with respect to a parameter to obtain a new solution. Toomre (1963) and Lynden-Bell (1989) exploited this linearity to find a family of velocity-density pairs by differentiating the Kuzmin (1956) model . In the same way, Satoh (1980) derived new models from the Miyamoto & Nagai (1975) set of potential-density pairs by differentiating with respect to a parameter. Here we use a similar approach by integrating a known solution with respect to a parameter. If the integral is tractable the result is a new potential-density pair. Maple 11 was used in this work. Maple was invaluable in simplifying the unwieldy intermediate results but needed a good deal of coaxing to handle the messier equations. The Maple results were cross checked in several ways. A common notation is used throughout. Cylindrical co-ordinates (R, z) are used; Σ is the disk surface density; σ0 is the surface density at R = 0; and α is the disk radius. Φ is the potential where Φ is always negative and Φ(∞) = 0; FR and Fz are the gravitational field vectors with the standard sign convention, ie: F = −∇Φ . r cos(θ) + I r sin(θ); √z = √r cos(θ/2) + I√r sin(θ/2), valid for −π < θ < π. In this way the function √z is unambiguous with continuous derivatives except on the negative real axis where it is discontinuous. The principal values of elementary functions are used so that, for instance, for z = x + I y = -- 3 -- 2. The n=0 Disk: The Flattened Homeoidal Shell A homeoid is a shell of uniform density which is bounded by similar spheroids. Newton was the first to prove that the net force is 0 (i.e., the potential is constant) within these shells. See Chandrasekhar (1987) for historical background. The infinitely thin homeoidal shell is a differential element of a spheroid. The homeoidal disk is the limiting case for which the minor axis approaches 0. The collapsed density of the thin homeoid is: Σhom(R; α) =(σ0/p1 − R2/α2 0 for R < α for R > α (1) Lynden-Bell (1989) gives a formula for the external potential of the thin homeoidal shell. Taking the limit c = (1 − e2)1/2α → 0 gives a solution for the disk which is valid at all R and z. Cuddeford (1993) gives an expression for the potential of this disk which is simpler than previous solutions: Equation 2 can be further simplified. First make the trivial transformation: 2α Φhom(R, z; α) = −2πασ0Gsin−1" pz2 + (R + α)2 +pz2 + (R − α)2# Φhom(R, z; α) = −2πασ0Gsin−1"pz2 + (R + α)2 −pz2 + (R − α)2 # 2R Now use the identity A1 to obtain Φhom(R, z; α) = −πασ0G (cid:20)sin−1(cid:18) α − I z R (cid:19) + sin−1(cid:18) α + I z R (cid:19)(cid:21) (2) (3) (4) Equation 4 can be integrated over α whereas Equation 3 yields an impossible integral. Equation 4 must be real valued on physical grounds. This is easy to prove by noting that sin−1(x) is an odd function of x and so the odd powers of I z in the series expansion of equation 4 cancel, leaving a real valued result. The gravitational field for the collapsed homeoid is obtained from the potential. Equation 4 yields particularly simple expressions for the field vectors FR,hom and Fz,hom: πασ0G R " FR,hom(R, z; α) = − Fz,hom(R, z; α) = −πασ0G" α − I z + I pR2 − (α − I z)2 pR2 − (α − I z)2 − α + I z pR2 + (α − I z)2# pR2 − (α + I z)2# I (5) (6) -- 4 -- These force vectors can be expressed as entirely real functions using identities A8 and A9: FR,hom(R, z; α) = − Fz,hom(R, z; α) = − √2πασ0G √2πασ0G α√f1f2 − f3− z √f1f2 + f3 Rf1f2 sgn(z)√f1f2 + f3 f1f2 Where f1 =pz2 + (R + α)2 f2 =pz2 + (R − α)2 f3 = α2 − R2 − z2 (7) (8) (9) 3. The n=1 Disk: The Maclaurin disk Beginning in the early 18'th century Colin Maclaurin, along with James Ivory and many others, studied the properties of elliptical bodies. Chandrasekhar (1987) includes a very good historical summary. See also Binney & Tremaine (2008); Bertin (2000); Schmidt (1956); Mihalas & Routly (1968); Kalnajs (1971, 1972). The homogeneous oblate spheroid is the simplest case of a spinning body for which the gravi- tational attraction balances the centrifugal force. The Maclaurin disk, also known as the Kalnajs disk (Kalnajs 1972), is a limiting case for which minor axis is 0. The Maclaurin disk is defined by the surface density: ΣM ac(R; α) =(σ0p1 − R2/α2 0 for R < α for R > α (10) There are a few solutions for the potential of the Maclaurin disk in the literature. Mihalas & Routly (1968) gives an expression for the potential of an oblate homogeneous spheroid based on the deriva- tion in Schmidt (1956). The potential of the Maclaurin disk can be found by letting the eccentricity e → 1 while holding the mass constant. Hunter (1963) gives the solution for the Maclaurin disk as a series of Legendre polynomials in elliptical coordinates. Neugebauer & Meinel (1995); Meinel (2001); Gonz´alez & Reina (2006) give a closed form solution for the potential of the Maclaurin disk in elliptic coordinates. The starting point here is the potential-density pair for the n=0 disk, the flattened homeoid for which Σhom(R; α) = σ0/p1 − R2/α2. The surface mass density of the Maclaurin disk is found from the transformation: ΣM ac(R; α) = 1 αZ α 0 Σhom(R; α) dα = 1 αZ α 0 σ0 p1 − R2/α2 dα = σ0p1 − R2/α2 (11) -- 5 -- The corresponding potential is: ΦM ac(R, z; α) = 1 αZ α 0 Φhom(R; α) dα = −πσ0G α Z α 0 α(cid:20)sin−1(cid:18) α − I z R (cid:19) + sin−1(cid:18) α + I z R (cid:19)(cid:21) dα (12) Where the expression for Φhom is given by equation 4 above. Use integral 2.813 and 2.833 from Gradshteyn & Ryzhik (1994) to obtain: ΦM ac(R, z; α) = − πσ0G 4α (cid:20)(2α2 − R2 + 2z2)(cid:18)sin−1(cid:18) α + I z + α(cid:16)pR2 − (α − I z)2 +pR2 − (α + I z)2(cid:17) − 3 z(cid:16)IpR2 − (α − I z)2 − IpR2 − (α + I z)2(cid:17)(cid:21) R (cid:19) + sin−1(cid:18) α − I z R (cid:19)(cid:19) Equation 13 can be converted to an entirely real expression by using the identities A1, A6 and (13) (14) A7: ΦM ac(R, z; α) = − πσ0G 4α (cid:20)2(2α2 − R2 + 2z2)sin−1(cid:18) f1 − f2 2R (cid:19) + √2αpf1f2 − f3 − 3√2 z pf1f2 + f3(cid:21) where f1, f2, f3 are given by equation 9 above. The gravitational field of the Maclaurin disk is found from the potential using Φ as given by equation 13. The resulting expressions were expressed as entirely real functions by using the identities A1, A8, and A9 : FR,M ac(R, z; α) = − Fz,M ac(R, z; α) = − πσ0G 2Rα "2R2sin−1(cid:18) f1 − f2 2R (cid:19) √f1f2 − f3 + √2α(α2 − R2 + z2) √2 z (α2 + R2 + z2) − α " − 2zsin−1(cid:18)f1 − f2 2R (cid:19) πσ0G f1f2 f1f2 √f1f2 + f3 √f1f2 − f3 + 2√2αz + √2 sgn(z) (α2 − R2 − z2) f1f2 √f1f2 + f3 f1f2 # # (15) (16) -- 6 -- where f1, f2, f3 are given by equation 9 above. The potential on the z axis and on the z = 0 plane can be found by by taking limits of equation 14: ΦM ac(0, z; α) = − ΦM ac(R, 0; α) =  πσ0G α (cid:20)(cid:0)α2 + z2(cid:1) sin−1(cid:18) − π2σ0G 4α (cid:0)2α2 − R2(cid:1) 2α h(2α2 − R2)sin−1( α − πσ0G α √α2 + z2(cid:19) − α z (cid:21) R ) + α√R2 − α2i forR ≤ α forR ≥ α (17) (18) The radial force vector in the z = 0 plane is found by taking the limit of equation 15 or by differentiating equation 18 with respect to R. FR,M ac(R, 0; α) =  2α − π2Rσ0 G α hRsin−1(α/R) − αp1 − α2/R2i − πσ0G forR ≤ α forR ≥ α (19) The axial force vector on the z axis is found by taking the limit of equation 15 or differentiating equation 17 with respect to z. Fz,M ac(0, z; α) = − 2πσ0G α (cid:20)zsin−1(cid:18) α √α2 + z2(cid:19) − α sgn(z)(cid:21) (20) 4. The n=2 Disk Gonz´alez & Reina (2006) give a closed form solution for the potential of the n=2 disk in elliptic coordinates. The disk surface density of the n=2 disk is ΣD2(R; α) =(σ0(1 − R2/α2)3/2 0 for R < α for R > α (21) This mass distribution can be obtained from equation 10, the disk surface density of the n=1 disk, with the transformation: ΣD2(R; α) = The corresponding potential is: ΦD2(R, z; α) = 3 α3 Z α 0 3 α3 Z α 0 α2ΣM ac(R; α)dα α2ΦM ac(R, z; α) dα (22) (23) -- 7 -- Where the expression for ΦM ac is given by equation 13 above. Equation 23 contains terms which using Gradshteyn & Ryzhik (1994) 2.262, 2.813, and 2.833. After collecting terms, the result is reasonably compact: after change of variable have with the form R xnsin−1(x)dx and R xn√1 ± x2dx and can be solved R (cid:19) + sin−1(cid:18) α − I z 64α3 "3(8α4 − 8α2R2 + 16α2z 2 + 3R4 − 24R2z 2 + 8z 4)(cid:18)sin−1(cid:18) α + I z ΦD2(R, z; α) = − πσ0G R (cid:19)(cid:19) + α(18α2 − 9R2 + 26z 2)(cid:16)pR2 − (α + I z)2 +pR2 − (α − I z)2(cid:17) − z(58α2 − 55R2 + 50z 2)(cid:16)IpR2 − (α + I z)2 − IpR2 − (α − I z)2(cid:17)(cid:21) (24) Equation 24 can be converted to an entirely real expression by using the identities A1, A6 and A7: ΦD2(R, z; α) = − πσ0G 64α3 "6(8α4 − 8α2R2 + 16α2z2 + 3R4 − 24R2z2 + 8z4)sin−1(cid:18) f1 − f2 2R (cid:19) + √2α(18α2 − 9R2 + 26z2)pf1f2 − f3 √2 z (58α2 − 55R2 + 50z2)pf1f2 + f3# − (25) where f1, f2, f3 are given by equation 9 above. The gravitational field of the n=2 disk is the gradient of the potential using Φ as given by equation 24. The resulting expressions were expressed as entirely real functions by using the identities by using identities A1, A8, and A9 : FR,D2(R, z; α) = − 3πσ0G 16Rα3"2R2(4α2R2 − 3R4 + 12R2z 2)sin−1(cid:18) f1 − f2 2R (cid:19) √f1f2 − f3 + √2α(2α4 − 5α2R2 + 4α2z 2 + 3R4 − 25R2z 2 + 2z 4) + √2 z (2α4 + 9α2R2 + 4α2z 2 − 13R4 − 11R2z 2 + 2z 4) f1f2 f1f2 √f1f2 + f3 Fz,D2(R, z; α) = − πσ0G 4α3 " − 6z(2α2 − 3R2 + 2z 2)sin−1(cid:18) f1 − f2 2R (cid:19) + √2αz(13α2 − 13R2 + 17z 2) + √2 sgn(z)((4α4 − 8α2R2 − 3α2z 2 + 4R4 − 7R2z 2 − 11z 4) f1f2 √f1f2 − f3 where f1, f2, f3 are given by equation 9 above. √f1f2 + f3 f1f2 # # (26) (27) -- 8 -- The potential on the z axis and on the z = 0 plane can be found by by taking the limit of equation 25: ΦD2(0, z; α) = − πσ0G 4α3 (cid:20)3(cid:0)α4 + 2α2z2 + z4(cid:1) sin−1(cid:18) α √α2 + z2(cid:19) − α z (5α2 + 3z2) (cid:21) (28) ΦD2(R, 0; α) =  − 3π2σ0G 64α3 (cid:0)8α4 − 8α2R2 + 3R4(cid:1) 32α h(8α4 − 8α2R2 + 3R4)sin−1( α − 3πσ0G R ) + 3α(2α2 − R2)√R2 − α2i forR ≤ α forR ≥ α (29) The radial force vector in the z = 0 plane is found by taking the limit of equation 26 or differentiating equation 29 with respect to R. FR,D2(R, 0; α) =  16α3 (4α2 − 3R2) − 3π2Rσ0 G 8α3 hR(4α2 − 3R2)sin−1(α/R) − α(2α2 − 3R2)p1 − α2/R2i − 3πσ0G forR ≤ α forR ≥ α (30) The axial force vector on the z axis is found by taking the limit of equation 27 or differentiating equation 28 with respect to z. Fz,D2(0, z; α) = − πσ0G 2α3(α2 + z2)(cid:20) − 6(α2 + z2)2zsin−1(cid:18) α √α2 + z2(cid:19) + αz2(13α2 + 17z2) + α sgn(z)(4α4 − 3α2z2 − 11z4)(cid:21) (31) 4.1. Comparison of the Maclaurin disk and the n=2 disk Table 1 compares important properties of the two disks. Figure 1a compares the surface mass density and potential. Figure 1b compares rotational velocity in the disks. The n=2 disk is more centrally concentrated than the Maclaurin disk. The rotational velocity increases more quickly in the inner disk and begins to fall before reaching the edge of the disk. As is apparent from figure 1b, the derivative of the circular velocity of the Maclaurin disk is discontinuous at the edge of the disk whereas the n=2 disk is better behaved. -- 9 -- 2.0 1.5 1.0 0.5 y t i c o l e v r a l u c r i C 0.0 0.2 0.4 0.6 Radial distance 0.8 1.0 Maclaurin disk n=2 Disk 0.0 0.0 0.2 0.4 0.6 0.8 1.0 1.2 Maclaurin disk Radial distance n=2 Disk Point Mass y t i s n e D s s a M e c a f r u S 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0.0 Fig. 1. -- Compare the surface mass density and circular velocity, Vc = p(−V FR(R, z), for the Maclaurin disk and the n=2 disk. Reduced units are used so that the masses of both disks and the point mass are 1.0 Table 1 comparison of the Maclaurin and the n=2 disks Property Maclaurin disk n=2 Disk Surface density Total mass Σ(R) = σ0p1 − R2/α2 M = 2 Circular velocity V 2 c (R, 0) = = Disk edge velocity V 2 c (α, 0) = = 3 πα2σ0 π2R2σ0G 2α 3πR2M G 4α3 π2ασ0G 2 3πM G 4α = σ0(1 − R2/α2)3/2 = 2 5 πα2σ0 3π2R2σ0G(4α2 − 3R2) 15πR2M G(4α2 − 3R2) 3π2ασ0G 16α3 32α3 = = = = 16 15πM G 32α 5. Example: The force field of a simple galaxy model Three-dimensional problems such as that of the structure and kinematics of the extra-planar gas will benefit from the use of the new density-potential pairs. A simple galaxy model was constructed for illustration. The model consists of a an n=2 disk and a core/bulge region modeled as a point mass. This model is defined by three parameters: the mass of the disk, the mass of -- 10 -- 250 200 150 100 50 1 - s m k y t i c o l e v r a l u c r i C 0 4 6 8 10 12 14 16 18 z=0 Radial distance kpc z=2.6 kpc z=3.9 kpc Fig. 2. -- Circular velocity, Vc = p(−V FR(R, z), of a simple galaxy model at z=0, z=2.6 kpc, and z=3.9 kpc above the disk. The model consists of an n=2 disk and a point mass to represent the core + bulge. The masses of the two components are Mdisk = 3.5 × 1010M⊙ and MP ointM ass = 4 × 1010M⊙. The radius of the disk is α = 19 kpc. the core/bulge region, and the diameter of the disk. As shown in figure 2, the circular velocity, calculated as Vc = p(−V FR(R, z), is nearly constant over much of the disk. Also, the derivative of the circular velocity with z is nearly linear over a wide range of both R and z. Figure 2 agrees surprisingly well with figure 5 of Fraternali et al. (2005) which shows that the measured velocity of HI for NGC891 decreases linearly with the height above the disk. See also Rand (1997); Swaters et al. (1997); Kamphuis et al. (2007); Oosterloo et al. (2007); Fraternali & Binney (2006); Barnabe et al. (2006). Further work is planned on this topic. 6. Summary and conclusion We have presented new solutions for a family of finite disks. Closed form expressions in cylindrical coordinates using elementary functions are given for the potential and gravitational force for the disks with surface density Σn = σ0(1 − R2/α2)n−1/2 with n = 0, 1, 2. Expressions are also given for the limiting cases of R = 0 and z = 0. These solutions fill a need and should make it easier to model 3-D gravitational phenomenon involving disk galaxies. This is particularly important due to the recent availability of detailed kinematic data above the plane of the disk. -- 11 -- I am grateful to the anonymous referee for useful suggestions and comments which improved the presentation. -- 12 -- A. Some simple identities A number of identities are gathered here. In all cases x, y ∈ ℜ The ranges of validity must avoid the discontinuity of the principal value of the square root function on the negative real axis. Equation A1 can be proved by taking the sin of both sides; reducing the terms using the identities sin(a + b) = sin(a) cos(b) + cos(a) sin(b) and sin(2a) = 2 sin(a) cos(b); and substituting cos =p1 − sin2. The other identities can be found by substituting into the relation √x + I y = qpx2 + y2 + x + sgn(y)Iqpx2 + y2 − x √2 and into the expression found by taking the reciprocal of both sides. 2p(x − 1)2 + y2i sin−1(x − I y) + sin−1 (x + I y) = 2sin−1h 1 2p(x + 1)2 + y2 − 1 √x − I y + √x + I y = √2qpx2 + y2 + x I√x − I y − I√x + I y = sgn(y)√2qpx2 + y2 − x = √2qpx2 + y2 + x px2 + y2 = sgn(y)√2qpx2 + y2 − x px2 + y2 √x − I y √x − I y √x + I y − √x + I y + 1 I 1 I q1 − (x + I y)2 +q1 − (x − I y)2 = √2qp1 − 2x2 + 2y2 + x4 + 2x2y2 + y4 + 1 − x2 + y2 Iq1 − (x + I y)2 − Iq1 − (x − I y)2 = sgn(xy)√2qp1 − 2x2 + 2y2 + x4 + 2x2y2 + y4 − 1 + x2 − y2 (A1) (A2) (A3) (A4) (A5) (A6) (A7) (A8) 1 + q1 − (x − I y)2 q1 − (x − I y)2 − I 1 q1 − (x + I y)2 q1 − (x + I y)2 I = √2qp1 − 2x2 + 2y2 + x4 + 2x2y2 + y4 + 1 − x2 + y2 = sgn(xy)√2qp1 − 2x2 + 2y2 + x4 + 2x2y2 + y4 − 1 + x2 − y2 p1 − 2x2 + 2y2 + x4 + 2x2y2 + y4 p1 − 2x2 + 2y2 + x4 + 2x2y2 + y4 (A9) -- 13 -- REFERENCES Barnabe, M., Ciotti, L., Fraternali, F., & Sancisi, R. 2006, A&A, 446, 61 Bertin, G. 2000, Dynamics of galaxies (Cambridge University Press) Binney, J. & Tremaine, S. 2008, Galactic Dynamics: Second Edition (Princeton University Press, Princeton, NJ.) Casertano, S. 1983, MNRAS, 203, 735 Chandrasekhar, S. 1987, Ellipsoidal figures of equilibrium (New York : Dover, 1987) Ciotti, L. & Giampieri, G. 2007, MNRAS, 376, 1162 Ciotti, L. & Marinacci, F. 2008, MNRAS, 387, 1117 Conway, J. T. 2000, MNRAS, 316, 540 Cuddeford, P. 1993, MNRAS, 262, 1076 Evans, N. W. & Collett, J. L. 1993, MNRAS, 264, 353 Evans, N. W. & de Zeeuw, P. T. 1992, MNRAS, 257, 152 Fraternali, F. & Binney, J. J. 2006, MNRAS, 366, 449 Fraternali, F., Oosterloo, T. A., Sancisi, R., & Swaters, R. 2005, in ASP Conf. Ser. 331: Extra- Planar Gas, 239 -- + Freeman, K. C. 1970, ApJ, 160, 811 Gonz´alez, G. A. & Reina, J. I. 2006, MNRAS, 371, 1873 Gradshteyn, I. S. & Ryzhik, I. M. 1994, Table of integrals, series and products (Academic Press, 5th ed. Edited by Jeffrey, A) Hunter, C. 1963, MNRAS, 126, 299 Hunter, J. H., Ball, R., & Gottesman, S. T. 1984, MNRAS, 208, 1 Hur´e, J.-M., Hersant, F., Carreau, C., & Busset, J.-P. 2008, A&A, 490, 477 Hur´e, J.-M. 2005, A&A, 434, 1 Kalnajs, A. J. 1971, ApJ, 166, 275 -- . 1972, ApJ, 175, 63 Kamphuis, P., Peletier, R. F., Dettmar, R.-J., van der Hulst, J. M., van der Kruit, P. C., & Allen, -- 14 -- R. J. 2007, A&A, 468, 951 Kuzmin, G. G. 1956, Astr.Zh., 33, 27 Lass, H. & Blitzer, L. 1983, Celestial Mechanics, 30, 225 Lynden-Bell, D. & Pineault, S. 1978, MNRAS, 185, 679 Lynden-Bell, D. 1989, MNRAS, 237, 1099 Meinel, R. 2001, in Exact Solutions and Scalar Fields in Gravity: Recent Developments (Edited by A Macias, J Cervantes-Cota, and C Lammerzahl; Kluwer Academic Publishers, New York), 69 -- 75 Mestel, L. 1963, MNRAS, 126, 553 Mihalas, D. & Routly, P. M. 1968, Galactic astronomy (W.H. Freeman and Company) Miyamoto, M. & Nagai, R. 1975, PASJ, 27, 533 Neugebauer, G. & Meinel, R. 1995, Physical Review Letters, 75, 3046 Oosterloo, T., Fraternali, F., & Sancisi, R. 2007, AJ, 134, 1019 Pedraza, J. F., Ramos-Caro, J., & Gonzalez, G. A. 2008, MNRAS, 390, 1587 Rand, R. J. 1997, ApJ, 474, 129 Satoh, C. 1980, PASJ, 32, 41 Schmidt, M. 1956, Bull. Astron. Inst. Netherlands, 13, 15 Swaters, R. A., Sancisi, R., & van der Hulst, J. M. 1997, ApJ, 491, 140 Toomre, A. 1963, ApJ, 138, 385 Vokrouhlicky, D. & Karas, V. 1998, MNRAS, 298, 53 This preprint was prepared with the AAS LATEX macros v5.2.
astro-ph/0002235
1
0002
2000-02-10T19:04:45
Problems encountered in the Hipparcos variable stars analysis
[ "astro-ph" ]
Among the 17 volumes of results from the Hipparcos space mission, two are dedicated to variable stars. These two volumes arose from the work of two groups, one at the Geneva Observatory and one at RGO (Royal Greenwich Observatory), on the 13 million photometric measurements produced by the satellite for 118204 stars. The analysis of photometric time series permitted us to identify several instrumental and mathematical problems: overestimation of the precision, offsets in the zero-points depending on the field of view, mispointing effects, image superpositions, trends in the magnitudes, binarity effects (spurious periods and amplitudes) and time-sampling effects. In this article we summarize some of the problems encountered by the Geneva group.
astro-ph
astro-ph
Delta Scuti and Related Stars ASP Conference Series, Vol. N , 2000 M. Breger & M.H. Montgomery, eds. Problems encountered in the Hipparcos variable stars analysis Laurent Eyer Instituut voor Sterrenkunde, Katholieke Universiteit Leuven, Celestijnenlaan 200 B, B-3001 Leuven, Belgie Michel Grenon Observatoire de Gen`eve, CH-1290 Sauverny, Suisse Abstract. Among the 17 volumes of results from the Hipparcos space mission, two are dedicated to variable stars. These two volumes arose from the work of two groups, one at the Geneva Observatory and one at RGO (Royal Greenwich Observatory), on the 13 million photometric measurements produced by the satellite for 118 204 stars. The analysis of photometric time series permitted us to identify sev- eral instrumental and mathematical problems: overestimation of the pre- cision, offsets in the zero-points depending on the field of view, mispoint- ing effects, image superpositions, trends in the magnitudes, binarity ef- fects (spurious periods and amplitudes) and time-sampling effects. In this article we summarize some of the problems encountered by the Geneva group. 1. Introduction The variable star analysis of the Hipparcos photometric data was an iterative process in interaction with the FAST and NDAC data reduction consortia. We started with data extracts from FAST, then from NDAC and finally worked on the whole merged data set, that is on the 118 204 time series. The result of variable star analysis is a beautiful by-product of the mission and it was clear that it had to be published at the same time as the astrometric results (ESA 1997). The time available for the photometric analysis was then short in order to match the deadlines. The teams were put under strong pressure. The approach was then to produce a robust analysis restricting the analysis to statistically well-confirmed variables, leaving suspected variables and ambiguous cases for further analysis. Because several instrumental problems were identified and solved, the vari- able star study definitely improved the overall quality of the Hipparcos photom- etry available now on the CD-ROMs. 147 148 Eyer & Grenon HIP 96647. Left: magnitudes before the data merging, Figure 1. open squares are for the following FOV, crosses for the preceding FOV. Right: the corrected final solution. 2. Hipparcos main-mission photometry Although the telescope diameter is small (29 cm), Hipparcos achieved a high pho- tometric precision in the wide Hp band (335 to 895 nm), thanks to the chosen time allocation strategy and to the frequent on-orbit photometric calibrations, making use of a large set of standard stars. The time allocated for a star ob- servation was adapted to its magnitude in order to homogenize the astrometric precision. The photometric reduction was made in time slices of 10 hours, called reduced great circles (RGC). During that time interval, the satellite scanned a closed strip in the sky, measuring about 2600 stars, among them 600 standard stars. The FAST and NDAC consortia independently reduced the photometry. They had to map the time evolution of the spatial and chromatic response of the detection chains for both fields of view (FOV), the preceding and the following. In Fig. 1, a zero-point problem between FOV magnitude scales is shown, before and after its correction. The light of the star was modulated by a grid for astrometric purposes. The transmitted signal was modeled by a Fourier series of 5 parameters. From this model two estimates of the intensity were done: one, measuring the integrated signal, the "DC mode", was robust to the duplicity but more dependent on the background, and the second, measuring the amplitude of the modulation, the "AC mode", was sensitive to the duplicity but not to the background (cf. van Leeuwen et al. 1997). These two estimates and their accuracies are given in the Epoch Photometry Annex (CD-ROM 2) and in the Epoch Photometry Annex Extension (CD-ROM 3). Problems encountered in the Hipparcos variable stars analysis 149 3. Noise and magnitude The precision of the magnitude is a function of the magnitude itself. In addition, for a star of constant magnitude, the errors may be variable. The data are then heteroscedastic. The correlation between the error and the magnitude, may generate some problems. For instance, the weighted mean cannot be used to estimate the central value of the magnitude distribution, if the amplitude is large as for the Miras. The global loss of precision as the satellite ages (Eyer & Grenon 2000) also needs to be considered. The usual period search algorithms are also sensitive to the inhomogeneity of the data. 3.1. Quoted transit errors During a single transit, a star was measured 9 times on average, the transit error σHp was derived in a first approximation from the spread of these measurements. However, the estimation did not include offsets which might have affected a whole transit, e.g., due to a mispointing or to a superposition of a star from the other FOV. In our first analysis, an empirical law was determined to correct the transit error underestimation, otherwise the number of candidate variable stars did not appear credible. During the phase of data merging, ad-hoc corrections were computed (Evans 1995). The errors were studied with different methods, comparing first the "average" error estimated from the σHp with the dispersion of the measurements on Hp. Another study by Eyer & Genton (1999) was made on the quoted errors using variograms; it showed a good general agreement with the Evans results, with some mild underestimations for faint magnitudes and some mild overestimations for the bright magnitudes in Evans' approach. 3.2. Time sampling The time sampling was determined by the satellite rotation speed and by the scanning law which were optimized to reach the most uniform astrometric pre- cision over the whole celestial sphere. The total number of transits per star is a function mainly of the ecliptic latitude. The time intervals between successive transits are 20-108-20-etc. . . minutes. The transits form groups which are sepa- rated by about one month, but the number of consecutive measurements as well as the time separation between groups of transits can vary strongly from one star to another. 3.3. Chromatic aging The irradiation by cosmic particles reduced the optical transmission with time. This aging was chromatic; it was worse than expected because the satellite had to cross the two van Allen Belts twice per orbit. Furthermore, the satellite was operational during a maximum of solar activity. For instance, the magnitude loss over 3.3 years was 0.8 mag for the bluest stars and only 0.15 mag for the reddest. The aging of the image dissector tubes were not uniform and distinct for both FOVs, therefore the aging corrections had to be calibrated as functions of the star location on the grid for each FOV. 150 Eyer & Grenon Figure 2. HIP 29862, an example of trend induced by aging Magnitude trends: An odd effect of the chromatic aging was the production of magnitude trends in the Hp time series. As the transmission loss was colour dependent, the magnitude correction had to be a function of the star colour. A colour index, monotonically growing with the effective wavelength of Hp band, had to be evaluated from heterogeneous sources. The precision of the equivalent V − I was highly variable. For stars with "bad" V − I colour, the magnitude correction was erroneous and produced a trend. A colour bluer than true gener- ates a spurious increase of the luminosity with the time. An example of a trend is given in Fig. 2. Stars like Be stars may also show quasi-linear trends over Selection of trends: the mission duration. The identification of spurious trends was iterative. LPVs, showing Gaussian residuals when modeled with a trend on top of their semi- periodic light-curve, were sorted first. But there was much more diversity in the data showing trends, true or spurious. So we used an Abbe test (Eyer & Grenon 2000) for a global detection. Stars with an Abbe test close to 1, or with a large trend or with very long periods, were flagged. Stars with possible envelopes were not retained. After visual inspection of the time series by Grenon, the number of stars selected by these different procedures was 2412. Correction of the star colour: The amplitude of the magnitude drift was used to correct the star colour. Indeed, if a time series shows a trend α which may be imputed to an incorrect initial colour, there is a possibility to recover the true star colour by the relation: (V − I)new = (V − I)old − 14290 α where α is expressed in magnitude per day. Problems encountered in the Hipparcos variable stars analysis 151 Two examples of series with outlying values. Left: Figure 3. HIP 35527 the case of a light pollution from the other FOV, where Sir- ius is the perturbing star (this case is not flagged); Right: HIP 57437 an example of mispointing effect with low values correctly flagged (open circles). Every selected case was investigated to decide whether the trend could be a consequence of an incorrect colour; 965 V − I indices were corrected this way with certainty and the origin of the errors on the colours was traced back. 4. Outlying values When studying variable stars, outlying values and anomalous data distributions are of great interest. Namely, it is important to distinguish outliers of instru- mental origin from those due to stellar physical phenomena. Some stars show luminosity changes on very short time scales. For Algol eclipsing binaries, the duration of the eclipse is short with respect to their period. With non-continuous time sampling, eclipses may appear as low luminosity points. UV Ceti stars show strong bursts in the U band on very short time scales. However, because of the width of the Hp band, the photospheric flux in the redder part of the band largely dominates that of the burst, with the result that no burst was detected with certainty in the M dwarfs. In Fig. 3 we present two cases of outlying values of instrumental origin. 4.1. Instrumental outliers: Mispointing effect The pointing precision of the satellite was normally better than 1 arcsec. How- ever after Earth or Moon eclipses and especially near the end of the mission when most gyroscopes were faulty, the problems of mispointing were more acute. The radius of the photocathode was 15 arcsec, with a lower sensitivity towards the 152 Eyer & Grenon edge. An inaccurate pointing was inducing a loss of counted photons, leading to dimmer points in the time series. The problems with extended objects were even worse, depending on their sizes. A similar situation happened with visual double systems when the sepa- ration was around 10 arcsec. In this case the target was either the primary or the photocenter of the system. From time to time the companion was on the edge or out of the FOV, diminishing the amount of collected light. Even when the two components were measured alternatively, the not-measured star might have sometimes entered in the FOV producing a luminosity excess mimicking a burst. 4.2. Instrumental outliers: Light pollution A neighbour star could contaminate the observed star, although most of the identified cases were rejected during the Input Catalogue compilation. The perturbing star was possibly a real neighbour or, more often, a star belonging to the other FOV. Several configurations are possible: • A star from the other FOV was added to the measured star. That is called a superposition effect. Stars in the Galactic plane were more often perturbed because of the higher star density. • The perturbing star was very bright and caused scattered light in the detection chain. This veiling glare could be felt even if the disturbing star was further than 15 arcsec. • When the separation was greater that 15 arcsec, the two effects of pollution and mispointing could produce high values of fluxes of the dim component. Fig. 4 shows the correlation between the asymmetry of the time series for double systems and the angular separation ρ. The asymmetry is positive for dimmer outlying values, and negative for pollution by the primary (bright out- lying values). 4.3. Selection of outliers The problems caused by the outliers were very acute at the beginning of the analysis; we had then to take drastic measures before searching for periods and amplitudes. We removed: • all measurements with a non-null flag. • the end of the mission, if the dispersion of the data was smaller than 0.3 before the day 8883. (the data for large amplitude variables were kept up to the end of the mission). • high luminosity values if there was a magnitude jump in consecutive tran- sits. • bad-quality measurements with transit errors higher than ǫ(DC) = 0.0005∗ 100.167 HpDC + 0.0014. Problems encountered in the Hipparcos variable stars analysis 153 Figure 4. a function of the angular separation ρ. Light pollution and mispointing effects for double stars as • temporarily one or two outliers to check their impact on the result of an analysis based on truncated time series. This removal represents a reduction by 6% of the number of measurements with non-zero flag. Suspect transits from the analysis of outliers were transmitted to the re- duction consortia, who flagged them according to the origin of the disturbance. 5. Alias and spurious periods The spectral window produces spurious periods when it is convoluted with the true spectrum. As a result, spurious periods around 0.09 d were frequently found for long periods or irregular variables as well as for stars showing magnitude trends. Periods of about 5 d for SR variables turned out to be nearly all spurious. With the Hipparcos time sampling the spectral window changes from one star to another and the alias effect had to be studied on a per star basis. A 58 day periodicity was found in many time series when applying the period search algorithm. This period corresponds to the time interval between consecutive measurements of double systems under the same angle with respect to the modulating grid (the modulated signal is higher when the components are parallel to the grid). 154 Eyer & Grenon 6. Advice about the use of the data We want to stress that caution should be taken in handling the epoch photom- etry, especially when the signal to noise ratio or when the number of retained measurements are small. The effects of multiperiodicity are generally very tricky. The spectral window should be investigated in detail and periods near sampling frequencies should be taken with care, in particular in the range 5 to 20 d where the Hipparcos photometry has the weakest detection capability. The selection of photometric data can be made according to their flags, the estimates of the transit errors and the background intensities. In case of doubt about outlying data, a look to the magnitude difference AC − DC will reveal problems related to duplicity and image superpositions since the amplitude of the modulated signal is reduced in the case of misaligned sources with respect to the grid orientation. The contents of the opposite FOV can be investigated thanks to their published positions. It is suggested to correct rather than to eliminate data since a loss of information might twist statistics. For an example of a successful selection procedure applied to the data of HIP 115510, see Lampens et al. (1999). 7. Conclusion Performing accuracy photometry in space is not free from problems. The same is true for the data analysis. Once the origin of the encountered problems is identified, it is possible to cope with them and determine precisely the domains of validity of the algorithms for search of periods and amplitudes. Globally the ratio quantity-quality generated by this mission for the study of variability has no equivalent up to now. References ESA 1997, The Hipparcos and Tycho catalogues, ESA SP-1200 Evans, D.W. 1995, Hipparcos Photometry Merging Report, RGO/NDAC 95.01 Eyer, L., and Genton, M.G. 1999, A&AS 136, 421 Eyer, L. and Grenon, M., 2000, in preparation Lampens, P., van Camp, M., Sinachopoulos, D. 1999 A&A, accepted van Leeuwen, F., Evans, D.W., Grenon, M., Grossmann, V., Mignard, F., Per- ryman, M.A.C. 1997, A&A323, L61
astro-ph/0210680
1
0210
2002-10-31T12:56:09
Interstellar extinction in the open clusters towards galactic longitude around 130 degrees
[ "astro-ph" ]
In this paper we present a detailed study of the intra-cluster reddening material in the young open clusters located around $l \sim 130^o$ using colour-excess diagrams and two-colour diagrams. The study supports the universality of the extinction curves for $\lambda \geq \lambda_J$, whereas for shorter wavelengths the curve depends upon the value of the $R_{cluster}$ (total-to-selective absorption in the cluster region). The value of $R_{cluster}$ in the case of NGC 654, NGC 869 and NGC 884 is found to be normal, whereas the value of $R_{cluster}$ in the cluster regions NGC 1502 and IC 1805 indicates an anomalous reddening law in these regions. It is also found that the extinction process in the $U$ band in the case of NGC 663 seems to be less efficient, whereas in the case of NGC 869 the process is more efficient.
astro-ph
astro-ph
Astronomy & Astrophysics manuscript no. MS2516 (DOI: will be inserted by hand later) November 3, 2018 2 0 0 2 t c O 1 3 1 v 0 8 6 0 1 2 0 / h p - o r t s a : v i X r a Interstellar extinction in the open clusters towards galactic longitude around 130o A. K. Pandey1, K. Upadhyay1, Y. Nakada2, K. Ogura3, 1 State Observatory, Manora Peak, Naini Tal, 263 129, Uttaranchal, India 2 Kiso Observatory, School of Science, University of Tokyo, Mitake-mura, Kiso-gun, Nagano 397-0101, Japan 3 Kokugakuin University, Higashi, Shibuya-ku, Tokyo 150-8440, Japan Received XXXX, 2002; accepted XXXX, In this paper we present a detailed study of the intra-cluster reddening material in the young open Abstract. clusters located around l ∼ 130o using colour-excess diagrams and two-colour diagrams. The study supports the universality of the extinction curves for λ ≥ λJ , whereas for shorter wavelengths the curve depends upon the value of the Rcluster (total-to-selective absorption in the cluster region). The value of Rcluster in the case of NGC 654, NGC 869 and NGC 884 is found to be normal, whereas the value of Rcluster in the cluster regions NGC 1502 and IC 1805 indicates an anomalous reddening law in these regions. In the case of NGC 1502 the value of Rcluster is found to be lower (2.57 ± 0.27) whereas in the case of IC 1805 it is higher (3.56 ± 0.29) than the normal value of 3.1. Although the intra-cluster material indicates a higher value of Rcluster in the NGC 663 region, the error in the estimation of Rcluster is too large to conclude anything. It is also found that the extinction process in the U band in the case of NGC 663 seems to be less efficient, whereas in the case of NGC 869 the process is more efficient. Key words. ISM: dust, extinction -- ISM: general -- open clusters and associations: general 1. Introduction Photometric studies are one of the most valuable and effi- cient tools for determining the physical properties of open clusters (e.g. distance, age, etc) and the interstellar matter (ISM) within the cluster as well as along the line of sight of the cluster. The accurate determination of the distance of star clusters is crucial for a wide range of astronomical studies. The interstellar extinction and the ratio of the total-to-selective extinction R = AV /E(B − V ) towards the cluster are important quantities that must be accu- rately known to determine the distances photometrically. Observations of star clusters in the near-infrared bands, especially of young clusters which are still embed- ded in remnants of their parental clouds, are very use- ful to study the extinction behaviour in these clusters. Observations of interstellar extinction due to the intra- cluster matter yield fundamental information on the op- tical properties of the particles responsible for extinction. The analysis of extinction curves in 20 OB stellar associ- ations indicates that big complexes are obscured by the same type of interstellar matter (Kiszkurno et al. 1984). On the other hand Kre lowski and Strobel (1983) com- pared the extinction law in two rich stellar aggregates, Send offprint requests to: A.K. Pandey ([email protected]) namely Per OB1 and Sco OB2, situated in different parts of the galactic disk. They found that the average extinc- tion curves of the two aggregates seem to differ substan- tially. The differences are large in the far-UV, which leads to the conclusion that an average extinction law cannot be applied to all associations. In an analysis of two different complexes of young stars (Sco-Ori region and Perseus re- gion) Kre lowski and Strobel (1987) found different shapes of extinction curves, supporting the assumption that the obscuring material in these two young star complexes have different physical properties. Consequently, in the case of young open clusters, the average extinction curve should not be used to correct spectral or photometric data for interstellar extinction. There is much evidence in the literature that indicates significant variations in the properties of the interstellar extinction and these refer mainly to high values of the total-to-selective extinction ratio R as compared to the normal value, 3.1, for the galactic diffuse medium (see e.g. Pandey et al. 2000 and references therein). Whittet (1977) reported that the value of R in the galactic plane can be represented by a sinusoidal function of the form R = 3.08+0.17 Sin(l +175o), which indicates a minimum value of R at l ≈ 95o. However, Tapia et al. (1991) found a value of R = 2.42 ± 0.09 for the cluster NGC 1502 (l = 143.7o), 2 Pandey et al.: interstellar extinction in open clusters Table 1. The details of the clusters used in the study Cluster l b Distance (pc) Log age E(B-V) Available (Yr) photometric data IC 1590 Be 62 NGC 436 NGC 457 NGC 581 Tr 1 NGC 637 NGC 654 NGC 663 Be 7 NGC 869 IC 1805 NGC 884 NGC 1502 123.13 123.99 126.07 126.56 128.02 128.22 128.55 129.09 129.46 130.13 134.63 134.74 135.08 143.65 -06.24 01.10 -03.91 -04.35 -01.76 -01.14 01.70 -00.35 -00.94 00.37 -03.72 00.92 -03.60 07.62 2940 2513 2942 2796 2241 2520 2372 2422 2284 2570 2115 2195 2487 900 6.54 7.22 7.78 7.15 7.13 7.43 6.96 7.08 7.13 6.60 7.10 6.67 7.15 7.03 0.32 0.85 0.48 0.48 0.44 0.63 0.70 0.85 0.82 0.80 0.58 0.80 0.58 0.74 which is significantly lower than the average galactic value of 3.1. The above discussions indicate that the interstellar ex- tinction shows a large range of variability from one line of sight to another. A precise knowledge of the spatial variability of interstellar extinction is important for the following reasons; (i) Since the extinction depends on the optical proper- ties of the dust grains, it can reveal information about the composition and size distribution of the grains. Consequently, variation of the extinction from one direc- tion to other may reveal the degree and nature of dust grain processing in the ISM (Fitzpatrick 1999). (ii) Since astronomical objects are viewed through in- terstellar dust, the wavelength dependence of extinction is required to remove the effects of dust obscuration from observed energy distributions. Uncertainties of extinction estimates limit the accu- racy of dereddened energy distributions. Such uncertain- ties might be acceptably small for very lightly reddened objects but can become important for modestly reddened objects. The intra-cluster extinction due to the remains of the star-forming molecular cloud decreases systematically with the age of the cluster (cf. Pandey et al. 1990). While analysing the U BV Ic data of NGC 663 we found that the V /(U − B) colour-magnitude diagram (CMD) for NGC 663 cannot be explained by a normal value of the extinction ratio E(U − B)/E(B − V ) = 0.72 (see Sect. 3). Some clusters in the direction of NGC 663 also appear to show abnormal extinction properties. This motivated us to study the interstellar extinction in detail towards the direction of the cluster NGC 663. U BV IcJHK U BV U BV Ic U BV Ic U BV Ic U BV U BV U BV RcIc U BV IcJHK U BV U BV RcIcJHK U BV RcIcJHK U BV RcIcJHK U BV IcJHK 2. Data online available database at The http://obswww.unige.ch/webda/ by Mermilliod (1995) provides an excellent compilation of the data on open clusters. We selected a range of 1200 < l < 1500 to study the intra-cluster extinction in open clusters. The online database lists 103 clusters in this longitude range. The parameters, such as distance, E(B − V ), age etc, are available for only 54 clusters. Seventeen clusters in this direction are quite young (log age < 7.5) and are suitable to study the behaviour of intra-cluster ISM. U BV CCD photometry for nine young clusters of the above sample is given by Phelps and Janes (1994). A casual look at the (U − B)/(B − V ) two-colour diagrams (TCDs) given by Phelps and Janes (1994, hereafter PJ94) indicates that the photometric data of 3 clusters, namely Be 62, NGC 637 and NGC 663, do not satisfy the main-sequence (MS) curve, where apparent U V excess can be seen for stars having spectral type later than A. PJ1994 concluded that this may be due to U V excess of pre-main-sequence (PMS) stars and this part of the (U − B)/(B − V ) TCD is not suitable for reddening estimation. We will discuss (U − B)/(B − V ) TCDs in ensuing sections. The present analysis has been carried out using photometric data of 14 clusters and details are given Table 1. 3. Extinction One of the main characterstics of the diffuse interstellar matter in the Galaxy is its irregular structure. This makes it difficult to map the extinction as a function of galac- tic longitude and distance. Young open clusters also show differential reddening due to the remains of the associated parental clouds. The differential reddening decreases with the age of the clusters (cf. Pandey et al. 1990). However, it does not show any correlation with the location of the cluster in the galactic disk. Pandey et al.: interstellar extinction in open clusters 3 Thus the extinction in star clusters arises due to two distinct sources; (i) The general interstellar medium (ISM) in the foreground of the cluster, and (ii) the localized cloud associated with the cluster. While for the former compo- nent a value of R = 3.1 is well accepted (Wegner 1993, Lida et al. 1995, Winkler 1997), for the intra-cluster re- gions the value of R varies from 2.42 (Tapia et al. 1991) to 4.9 (Pandey et al. 2000 and references therein) or even higher depending upon the conditions occurring in the re- gion. 3.1. Extinction Curve Extinction has often been analyzed using a two-colour nor- malizations of the form E(λ−V )/E(B −V ). In the present work the following methods are used to derive the ratio of E(λ − V )/E(B − V ). a) Colour excess diagrams (CEDs); method 'A' The colour excesses of the stars in the cluster region can be obtained by comparing the observed colours of the stars with their intrinsic colours derived from the MKK bi-dimensional spectral classification. For this pur- pose we used the data available in the online catalogue by Mermilliod (1995). When multiple data points are avail- able for a star, we used those selected by Mermilliod and given as MKS in the online catalogue. The colour excess in a colour index (λ − V ) is obtained from the relation E(λ − V ) = (λ − V ) − (λ − V )0, where (λ − V )0 is the intrinsic colour index and λ represents the magnitude in the U BV RIJHK pass bands. Intrinsic colours are ob- tained from the MKK spectral type-luminosity class rela- tion given by Schmidt-Kaler (1982) for U BV , by Johnson (1966) for V RI converted to the Cousin's system using the relation given by Bessel (1979), and by Koornneef (1983) for V JHK. The colour excesses E(U − V ), E(I − V ), E(J − V ), E(H − V ) and E(K − V ) are shown as a function of E(B − V ) in Figs. 1, 2, 3 and 4. The stars having Hα emission features and stars apparently lying away from the general distribution are not included in the analysis. A least-squares fit to the data is shown by a straight line which gives the ratio of E(λ − V )/E(B − V ) for the stars in the cluster region. The slope of the line representing the ratio E(λ − V )/E(B − V ) along with the error is given in Table 2. In general, the least-square errors are quite large. The reason for the large errors is mainly the small sam- ple and the small range in the E(B − V ). For comparison the colour excess ratios for the normal reddening law (cf. Johnson 1968, Dean et al. 1978) are also given in Table 2. b) Two colour diagrams (TCDs); method 'B' In most of the cases the MKK spectral classification is available only for a few stars of the cluster, which make the CEDs quite noisy (see e.g. the CEDs of NGC 654, NGC 663, NGC 869 and NGC 884). The TCDs of the form of (λ − V ) vs (B − V ), where λ is one of the wave- length of the broad band filters (R, I, J, H, K, L), provide 1.0 1.5 2.0 0.0 1.0 2.0 3.0 0.5 0.8 1.0 0.8 1.0 1.2 0.6 0.8 1.0 1.2 0.2 0.4 0.6 0.4 0.6 0.8 0.0 0.4 0.8 1.2 1.6 0.8 1.0 1.2 0.0 0.5 1.0 1.5 2.0 0.0 0.4 0.8 1.2 0.4 0.6 0.8 Fig. 1. E(U − V ) vs E(B − V ) colour-excess diagrams. Straight line shows a least-square fit to the data. 0.0 -0.5 -1.0 -1.5 -2.0 -0.4 -0.5 -0.6 -0.7 -0.8 0.0 0.4 0.8 1.2 1.6 0.3 0.4 0.5 0.6 Fig. 2. E(I − V ) vs E(B − V ) colour-excess diagrams. an effective method for separating the influence of the nor- mal extinction produced by the diffuse interstellar medium from that of the abnormal extinction arising within re- gions having a peculiar distribution of dust sizes (cf. Chini and Wargau 1990, Pandey et al. 2000). On these diagrams the unreddened MS and the normal reddening vector are practically parallel. This makes these diagrams useless for determining the amount of reddening, but instead, very useful for detecting anomalies in the reddening law. Chini and Wargau (1990) and Pandey et al. (2000) used TCDs to study the anomalous extinction law in the clusters M16 and NGC 3603 respectively. Figs. 5, 6, 7 and 8 show TCDs for the central region of the clusters. We used the data of the central region of the clusters to reduce the contami- nation due to field stars. The stars apparently lying away from the general distribution are not included in the anal- 4 Pandey et al.: interstellar extinction in open clusters Table 2. The colour excess ratios E(λ − V )/E(B − V ) obtained from the CEDs. Cluster E(U −V ) E(B−V ) E(I−V ) E(B−V ) E(J −V ) E(B−V ) E(H−V ) E(B−V ) E(K−V ) E(B−V ) NGC 654 NGC 663 NGC 869 NGC 1502 IC 1805 IC 1590 Normal -1.0 -2.0 -3.0 -4.0 -1.0 -2.0 -3.0 -4.0 -1.0 -2.0 -3.0 -4.0 −2.20 ± 0.26 −2.47 ± 0.26 −2.71 ± 0.27 −2.28 ± 0.25 −2.79 ± 0.30 −2.99 ± 0.30 2.12 ± 0.32 −1.20 ± 0.46 −1.75 ± 0.48 −1.59 ± 0.50 1.90 ± 0.13 −1.87 ± 0.16 −2.04 ± 0.23 −2.17 ± 0.22 1.74 ± 0.18 1.76 ± 0.08 −1.34 ± 0.04 −2.74 ± 0.30 −3.30 ± 0.37 −3.66 ± 0.38 1.56 ± 0.30 −1.33 ± 0.17 1.72 −2.58 −2.30 −1.25 −2.78 0.0 -1.0 -2.0 0.0 -1.0 -2.0 0.0 -1.0 -2.0 -1.0 -1.5 -2.0 -1.0 -1.5 -2.0 -1.0 -1.5 -2.0 -0.8 -1.2 -1.6 -0.8 -1.2 -1.6 -0.8 -1.2 -1.6 -0.8 -1.6 -2.4 -0.8 -1.6 -2.4 -0.8 -1.6 -2.4 -1.0 -2.0 -3.0 -1.0 -2.0 -3.0 -1.0 -2.0 -3.0 0.5 1.0 1.5 0.0 0.5 1.0 0.6 0.8 1.0 0.4 0.6 0.8 0.4 0.6 0.8 0.4 0.8 1.2 Fig. 3. E(J −V ), E(H −V ), E(K −V ) vs E(B−V ) colour- excess diagrams for the clusters IC 1805, NGC 1502 and NGC 663. Fig. 4. Same as Fig. 3 but for the clusters NGC 869, NGC 884 and NGC 654. ysis. The slopes of the distribution, mcluster, are given in in Table 3. The slopes of the theoretical MS, mnormal, on the TCDs, obtained for the stellar models by Bertelli et al. (1994) are also given in Table 3. The errors associated with the slopes are significantly smaller than the errors obtained in the CEDs. The values of the (λ − V )/(B − V ) can be converted to the ratio E(λ − V )/E(B − V ) using the following approximate relation; E(λ−V ) E(B−V ) = mcluster mnormal × [ E(λ−V ) E(B−V ) ]normal. c) (U − B)/(B − V ) two-colour diagram and colour- magnitude diagrams; method 'C' In absence of spectroscopic observations, the (U − B)/(B−V ) TCD and colour-magnitude diagrams (CMDs) are important tools to study the interstellar reddening to- wards the cluster as well as intra-cluster reddening. In the case of (U − B)/(B − V ) TCD a MS curve (e.g. given by Schmidt-Kaler 1982) is shifted along a reddening vector given by the ratio of E(U − B)/E(B − V ) ≡ X until a match between the MS and stellar distribution is found. In this method the reddening vector X plays an im- portant role. Over the years the observational as well as theoretical studies have used ( or 'misused', according to Turner 1994) a universal form for the mean galactic red- dening law. However, theoretical as well as observational estimates for the reddening vector X show a range from 0.62 to 0.80 (cf. Turner 1994). Recently DeGioia Eastwood et al. (2001) also preferred a value of X=0.64 in the case of Tr 14. The variability of X indicates the variations in the properties of the dust grains responsible for the extinc- tion. High values of X imply a dominance by dust grains of Pandey et al.: interstellar extinction in open clusters 5 Table 3. The slopes of the distribution of stars obtained from the (λ − V )/(B − V ) TCDs. Cluster (R−V ) (B−V ) (I−V ) (B−V ) (J −V ) (B−V ) (H−V ) (B−V ) (K−V ) (B−V ) NGC 654 NGC 663 NGC 869 NGC 884 NGC 1502 IC 1590 IC 1805 Normal −0.61 ± 0.02 −1.20 ± 0.04 −2.16 ± 0.25 −2.39 ± 0.23 −2.50 ± 0.25 −0.50 ± 0.02 −1.13 ± 0.03 −2.33 ± 0.30 −2.73 ± 0.35 −3.07 ± 0.43 −0.60 ± 0.02 −1.12 ± 0.02 −1.90 ± 0.10 −2.41 ± 0.10 −2.53 ± 0.11 −0.62 ± 0.02 −1.06 ± 0.02 −1.96 ± 0.08 −2.44 ± 0.08 −2.59 ± 0.10 −1.92 ± 0.24 −2.04 ± 0.24 −2.08 ± 0.24 −1.36 ± 0.09 −0.98 ± 0.04 −2.12 ± 0.15 −2.58 ± 0.17 −2.73 ± 0.18 −1.1 −2.42 −2.60 −0.55 −1.96 -0.5 -1.0 -1.5 0.0 1.0 2.0 3.0 0.0 1.0 2.0 3.0 0.0 -0.5 -1.0 -0.5 -1.0 -1.5 0.0 -1.0 -2.0 1.0 2.0 3.0 0.5 1.0 1.5 2.0 -0.2 -0.4 -0.6 -0.8 -0.5 -1.0 -1.5 -2.0 -0.2 -0.4 -0.6 -0.8 -1.0 -2.0 -3.0 0.3 0.6 0.9 0.3 0.6 0.9 0.7 1.4 2.1 -0.2 -0.4 -0.6 -1.0 -2.0 -3.0 0.0 1.0 2.0 0.0 0.2 0.5 1.0 2.0 3.0 Fig. 5. (R − V ) vs (B − V ) two-colour diagram. The data point shown by open circles is not included in the least-squares fit (see text). Fig. 6. (I − V ) vs (B − V ) two-colour diagram. small cross sections while the small values of X indicate a dominance of dust grains of larger cross section (cf. Turner 1994). Turner (1994) raised a question about the so called reddening-free parameter Q [= (U − B) − 0.72(B − V )]; how can Q be reddening-free when the reddening vector X is different from 0.72? As we have mentioned earlier, the distribution of stars in the V /(U − B) CMD of NGC 663 cannot be explained by a normal value of X (i.e. 0.72). The (U − B)/(B − V ) TCD and the V /(B − V ), V /(U − B) CMDs for the cen- tral region of the cluster NGC 663 are shown in Figs. 9 and 10 respectively. The data have been taken using the 105 cm Schmidt telescope of the Kiso Observatory, Japan (for details see Pandey et al. 2002). In Fig. 9, where the dashed curve shows the intrinsic MS by Schmidt-Kaler (1982) shifted along X = 0.72, we find a disagreement be- tween the observations and the MS at (B − V ) ∼ 0.90. PJ94 explained the disagreement due to the presence of pre-main-sequence (PMS) stars having U V excess. In Fig. 9 stars with V ≤ 16 are shown by filled circles. The appar- ent distance modulus for NGC 663 is (m − MV )=14.4 (cf. Pandey et al. 2002); the stars with V = 16 (i.e. MV =1.6) will have mass ∼ 2.5M⊙. Since the cluster has an age of ∼ 107 yr, the stars of ∼ 2.5M⊙ should have reached the MS and no longer be PMS stars. The TCD for these stars also does not support a normal value of X. We find that the MS shifted along a reddening vector of 0.60 and E(B − V )=0.68 explains the observations satisfactorily. The value of X in the NGC 663 cluster region can further be checked by comparing the theoretical zero-age- main-sequence (ZAMS) with the observed stellar distri- bution in the V /(B − V ) and V /(U − B) CMDs. Once the reddening is known from the (U − B)/(B − V ) CCD, the ZAMS is shifted to match the blue envelope of the ob- served stellar distribution in the V /(B −V ) and V /(U −B) CMDs. The ZAMS fitting for E(B − V )=0.68 and appar- ent distance modulus (m−MV )=14.4 is shown in Fig. 10a. Figure 10b shows V /(U − B) CMD where ZAMS, shifted for E(U − B)=0.49 (corresponding to the normal redden- ing vector X = 0.72) and (m − MV )=14.4, is shown by 6 Pandey et al.: interstellar extinction in open clusters -1.0 -2.0 -3.0 -1.0 -2.0 -3.0 -1.0 -2.0 -3.0 0.0 -0.5 -1.0 0.0 -0.5 -1.0 0.0 -0.5 -1.0 -0.5 -1.0 -1.5 -2.0 -0.5 -1.0 -1.5 -2.0 -0.5 -1.0 -1.5 -2.0 0.5 1.0 1.5 0.0 0.5 1.0 0.3 0.6 0.9 Fig. 7. (J − V ), (H − V ), (K − V ) vs (B − V ) two-colour diagrams for the clusters IC 1805, NGC 1502 and NGC 663. The data point shown by open circles are not included in the least-squares fit (see text). 0.0 -2.0 -4.0 0.0 -2.0 -4.0 0.0 -2.0 -4.0 0.0 -2.0 -4.0 -6.0 0.0 -2.0 -4.0 -6.0 0.0 -2.0 -4.0 -6.0 -1.0 -2.0 -3.0 -1.0 -2.0 -3.0 -1.0 -2.0 -3.0 0.0 1.0 2.0 3.0 0.0 1.0 2.0 3.0 0.0 0.5 1.0 Fig. 8. Same as Fig. 7 but for the clusters NGC 869, NGC 884 and NGC 654. -1.0 -0.5 0.0 0.5 1.0 1.5 0.5 1.0 1.5 Fig. 9. (U −B)/(B−V ) two-colour diagram for the cluster NGC 663 . The continuous curve represents the MS shifted along E(U − B)/E(B − V ) = 0.60, whereas dashed curve shows the MS shifted along a normal reddening vector. Star having V ≤ 16.0 are shown by filled circles. a dashed line, which clearly shows disagreement with the distribution of the stars. The ZAMS for E(U − B)=0.40 (corresponding to X = 0.60 and E(B − V ) = 0.68) nicely fits the blue envelope of the distribution. This supports an anomalous value X=0.60 for the slope in the NGC 663 cluster region. (U − B)/(B − V ) TCDs and V /(B − V ) CMDs were used to find out the value of X in all of the 14 clusters ex- amined in the present study. The (U − B)/(B − V ) TCD for 3 clusters namely Be 62, NGC 436 and NGC 637 is shown in Fig. 11. The values of the reddening vector X obtained from the TCD/CMDs are given in Table 4. The uncertainty in the estimated value of X arises due to un- certainties in the intrinsic colours (i.e. ZAMS), uncertain- ties associated with the observations and also uncertain- ties associated with the visual fit of the ZAMS to the ob- servations. The typical total uncertainty in the reported values of X is estimated to be ∼ 0.05. 4. The Value of R Table 4 indicates that the clusters Be 62, NGC 637 and NGC 663 show a smaller value whereas the clusters NGC 869 and NGC 436 show a higher value for X. The remain- ing 8 clusters show a normal value for X. The data given in Tables 2,3 and 4 are used to estimate the weighted (according to associated errors) mean value of the colour excess ratios E(λ − V )/E(B − V ) and the ratios are given in Table 5. Pandey et al.: interstellar extinction in open clusters 7 U band to K band are available,r is plotted as a function of λ−1 in Fig. 12. Fig. 12 indicates that the extinction in most of the cluster regions seems to be normal at λ > λI , except for the clusters NGC 1502 (where the colour excess ra- tios for λ > λI are less than the normal one) and NGC 1805 (where the colour excess ratio at λ ≥ λJ are higher than the normal one). Recent studies support a universal- ity of the extinction curves for λ > λI (see e.g. Cardelli et al. 1989, He et al. 1995). It is suggested that the nor- malization should be done using the E(V − K) instead of E(B − V ) (Tapia et al. 1991) because the E(V − K) does not depend on properties like chemical composition, shape, structure, degree of alignment of interstellar dust (cf. Mathis 1990 and references therein). Cardelli et al. (1989) found that the mean R dependent extinction law can be represented by the following relation Aλ AV = aλ + bλ R (1) where aλ and bλ can be obtained from the relations given by Cardelli et al. (1989). The above relation can be writ- ten, in terms of E(λ−V ) E(B−V ) , as E(λ−V ) E(B−V ) = R(aλ − 1) + bλ (2) The ratio of total-to-selective extinction towards the cluster direction 'Rcluster' is derived using the eqn. (2). The value of E(λ−V ) E(B−V ) (λ ≥ λJ ) given in Table 5 is used to estimate the value of Rcluster and the results estimate for Rcluster are given Table 6. The clusters that have a broad spectrum of data are discussed below. NGC 654 Sagar and Yu (1989) concluded that at wavelengths greater than 5500 A, the extinction is normal. The pres- ence of unusually well aligned interstellar grains indicated by the polarization measurements seems to increase the extinction in the U and B bands slightly (Sagar and Yu 1989). In the present work we find a rather normal extinc- tion law.The value of Rcluster is ∼ 2.97 ± 0.30(σ) which, within the error, is close to the normal value of 3.1. NGC 663 Using the colour excesses E(V − K) and E(B − V ) Tapia et al. (1991) found weak evidence for an anomalous reddening law with a value of Rcluster = 2.73±0.20, which is marginally lower than the normal value of 3.1. However, they felt that the scatter in their data is too large to con- clude about the value of Rcluster. Yadav and Sagar (2001) reported values for E(λ − V )/E(B − V ) ((λ ≥ λJ )) which are significantly smaller than the normal ones. The (weighted) mean value of Rcluster = 3.50±0.40(σ) suggests a marginally anomalous reddening law in the NGC 663 cluster region but in the opposite sense to that reported by Tapia et al. (1991) and Yadav and Sagar (2001). More near-IR data is needed to determine the -1.0 0.0 1.0 2.0 Fig. 10. V /B − V (the left panel), V /U − B (the right panel) CMDs for the cluster NGC 663. Star having V ≤ 16.0 are shown by filled circles. The continuous curve in left panel represents the ZAMS fitting for E(B − V )=0.68 and apparent distance modulus (m − MV )=14.4. The con- tinuous and dashed curves in right panel shows ZAMS shifted for E(U − B)=0.40 and E(U − B)=0.49 respec- tively. For details see text. Table 4. The value of the reddening vector X = E(U − B)/E(B − V ) obtained from the TCD/CMDs Cluster X IC 1590 Be 62 NGC 436 NGC 457 NGC 581 Tr 1 NGC 637 NGC 654 NGC 663 Be 7 NGC 869 IC 1805 NGC 884 NGC 1502 0.72 0.60 0.84 0.72 0.72 0.72 0.53 0.72 0.60 0.72 0.95 0.72 0.72 0.76 We define a parameter r which is the ratio of [E(λ − V )/E(B − V )]cluster (the ratio of colour excesses in the cluster region) and [E(λ−V )/E(B−V )]normal (the ratio of colour excesses for the normal reddening law). The values of E(λ − V )/E(B − V ) given in Table 5 are used to obtain the ratio r for the cluster region and resultant value of r is given in Table 6. For the clusters where the data from 8 Pandey et al.: interstellar extinction in open clusters Table 5. Weighted mean value of the colour excess ratios E(λ − V )/E(B − V ). Cluster E(U −V ) E(B−V ) E(R−V ) E(B−V ) E(I−V ) E(B−V ) E(J −V ) E(B−V ) E(H−V ) E(B−V ) E(K−V ) E(B−V ) NGC 654 NGC 663 NGC 869 NGC 884 NGC 1502 IC 1805 Normal 0.0 1.0 2.0 1.72 ± 0.07 −0.65 ± 0.03 −1.35 ± 0.07 −2.18 ± 0.25 −2.42 ± 0.24 −2.60 ± 0.26 1.60 ± 0.07 −0.55 ± 0.02 −1.28 ± 0.04 −2.43 ± 0.29 −2.84 ± 0.33 −3.08 ± 0.36 1.95 ± 0.07 −0.64 ± 0.03 −1.28 ± 0.02 −2.16 ± 0.16 −2.54 ± 0.15 −2.65 ± 0.17 1.72 ± 0.07 −0.66 ± 0.03 −1.21 ± 0.02 −2.29 ± 0.09 −2.61 ± 0.09 −2.77 ± 0.11 1.76 ± 0.07 −1.97 ± 0.19 −2.14 ± 0.24 −2.20 ± 0.24 −1.25 ± 0.05 −2.55 ± 0.22 −2.86 ± 0.23 −3.07 ± 0.24 1.72 ± 0.07 1.72 −1.25 −2.58 −0.60 −2.30 −2.78 0.0 1.0 2.0 2.0 0.0 1.0 2.0 2.0 0.5 1.0 1.5 0.5 1.0 1.5 0.5 1.0 1.5 2.0 Fig. 11. CCDs for the clusters Be 62, NGC 436 and NGC 637. The continuous curve represents the MS shifted along the reddening vector given in Table 4 and dashed curve shows the MS shifted along a normal reddening vector. Table 6. Mean value of the ratio r as a function of wavelength. λ−1(µm−1) Cluster NGC 654 NGC 663 NGC 869 NGC 884 NGC 1502 IC 1805 2.90 U 1.56 RC 1.25 IC 0.80 J 0.61 H 0.45 K Rcluster 1.00 ± 0.04 0.93 ± 0.04 1.19 ± 0.02 1.00 ± 0.04 1.02 ± 0.03 1.00 ± 0.03 1.08 ± 0.05 0.92 ± 0.04 1.07 ± 0.05 1.02 ± 0.05 1.08 ± 0.06 1.02 ± 0.03 1.02 ± 0.02 0.97 ± 0.02 1.00 ± 0.02 1.07 ± 0.14 1.07 ± 0.11 0.98 ± 0.05 0.97 ± 0.03 0.85 ± 0.05 1.12 ± 0.09 0.98 ± 0.09 1.09 ± 0.11 1.02 ± 0.05 0.99 ± 0.02 0.83 ± 0.05 1.14 ± 0.08 0.95 ± 0.03 1.10 ± 0.11 1.00 ± 0.04 0.98 ± 0.03 0.81 ± 0.04 1.16 ± 0.07 2.97 ± 0.30 3.50 ± 0.40 3.04 ± 0.20 3.19 ± 0.12 2.57 ± 0.27 3.56 ± 0.29 Rcluster in the NGC 663 cluster region. Here it is interest- ing to mention that the behaviour of the extinction curve towards UV also deviates from the normal one. Fig. 12 indicates a lower value for the E(U − V )/E(B − V ) ra- tio, whereas Yadav and Sagar (2001) reported a normal value for this ratio. They supplemented their data with the photometric spectral types which are based on the Q method, where they adopted E(U − B)/E(B − V )=0.72. Presumably a dominance of photometric spectral determi- nation forced the ratio of E(U − B)/E(B − V ) to a normal value. NGC 869 and 884 (h and χ Persei) From the extinction curve analysis Johnson (1965) found a value of Rcluster = 3.0 in the NGC 869 and NGC 884 cluster region. Tapia et al. (1984) also reported a nor- mal reddening law in the cluster region. Recently Yadav and Sagar (2001) found that the E(λ−V )/E(B −V ) ratios for λ ≥ λJ are smaller than the normal ones. The colour excess diagrams (method 'A') of NGC 869 indicates somewhat lower values for the ratios E(J − V )/E(B − V ) and E(K − V )/E(B − V ) but the errors are large, whereas in the case of NGC 884 the er- rors in the estimation of colour excess ratios are too large. The reason for the large errors is a small range in the E(B − V )(∼ 0.3 mag). Because of the large errors we have not used the colour excess ratios obtained from the Pandey et al.: interstellar extinction in open clusters 9 Various studies have been carried out to estimate the value of R in the cluster region of IC 1805, but the re- sults are not conclusive. Johnson (1968), Ishida (1969) and Kwon and Lee (1983) reported an anomalous red- dening law in the cluster region with values of Rcluster ∼ 5.7, ∼ 3.8 and ∼ 3.44 respectively. Kwon and Lee also reported a regional variation in the value of Rcluster, with a maximum value of Rcluster = 3.82 ± 0.15 for stars lo- cated in the outer region and the minimum of 3.06 ± 0.05 for stars located in the central region. Sagar and Yu (1990) found that the interstellar extinction law in the direction of most of the cluster members is normal. The colour ex- cess ratios E(λ − V )/E(B − V ) for λ ≥ λJ obtained in the present work indicate an anomalous reddening law in the cluster region of IC 1805. The value of Rcluster is es- timated as 3.56 ± 0.29, which also indicates an anomalous reddening behaviour in the cluster region. 5. The Aλ/AV curve Extinction has normally been analyzed using a two-colour normalization of the form E(λ − V )/E(B − V ). However, the true nature of the variability of observed extinction may be hidden by the choice of normalization. The quan- tity Aλ/AV reflects a more fundamental extinction be- haviour than the E(λ − V )/E(B − V ) (cf. Cardelli et al. 1989). The average colour excess ratios given in Table 5 can be used to estimate the quantity Aλ/AV in the fol- lowing manner, Aλ/AV = [E(λ − V )/E(B − V )Rcluster] + 1 where Rcluster is taken from Table 6. In Fig 13a the nor- malized extinction in the form Aλ/AV is plotted against λ−1 for the clusters NGC 654, NGC 663, NGC 869 and NGC 884 alongwith the average extinction law for Rcluster = 3.1 given by Cardelli et al. (1989). Fig 13a in- dicates that the agreement between observations and the extinction law by Cardelli et al. (1989) is good barring the AU /AV values for NGC 869 and NGC 663. In the case of NGC 869 the ratio AU /AV is higher than the normal one whereas in the case of NGC 663 it is lower than the normal one. The value of AU /AV = 1.46 in the case of NGC 663 supports Rcluster ∼ 3.5. Whereas in the case of NGC 869 the colour excess ratios indicate a perfectly normal red- dening law. It seems somewhat strange that in the case of these two clusters the ISM behaves at λU in a different way. The extinction law in the direction of two clusters IC 1502 and IC 1805 is found to be anomalous. The nor- malized extinction Aλ/AV for these two clusters along with the ρ Oph dark cloud (data taken from Martin and Whittet 1990) is plotted in Fig 13b. The effect of vary- ing Rcluster on the shape of the extinction curves is quite apparent at the shorter wavelengths for different environ- ments of the star forming regions. As we have discussed the extinction in the direction of star clusters arises due to the general ISM in the fore- Fig. 12. The ratio r as a function of λ−1. CEDs of NGC 869 in the further analysis. The TCDs (method 'B') indicate that the ratio of colour excesses E(λ − V )/E(B − V ) in both the clusters for λ ≥ λI are perfectly normal. It is interesting to note that the CED for NGC 869 yields E(U − V )/E(B − V ) = 1.90 ± 0.13. The V /(B −V ) and V /(U −B) CMDs also seem to support the ratio of E(U −V )/E(B −V ) = 1.9. In a recent study Keller et al. (2001) have adopted E(U − B)/E(B − V ) = 0.72 to fit the ZAMS to the stellar distribution on the V /(U − B) CMD of the stars in the NGC 869 and NGC 884 clus- ter region. We find that only in the case of NGC 884 the V /(U − B) CMD supports a normal value for the redden- ing vector X in the cluster region. NGC 1502 Tapia et al. (1991) found the ratio E(V − K)/E(B − V ) = 2.20 corresponding to Rcluster = 2.42 ± 0.09, which is significantly lower than the reddening value. Yadav and Sagar (2001) also found that the colour excess ratios E(λ − V )/E(B − V ) for λ ≥ λJ are significantly smaller than the normal ones. The colour excess ratios for λ ≥ λJ obtained in the present study are in good agreement with those reported by Tapia et al. (1991) and Yadav and Sagar (2001). The value of Rcluster is estimated as 2.57 ± 0.27 which is in good agreement with that obtained by Tapia et al. (1991). IC 1805 10 Pandey et al.: interstellar extinction in open clusters ground of the cluster and also due to the cloud associ- ated with the cluster. Various studies using OB type sin- gle stars support a value of R ∼ 3.1 for the general ISM (Wegner 1993, Lida et al. 1995, Winkler 1997). The min- imum reddening, E(B − V )min, towards the direction of the cluster is representative of reddening due to the fore- ground dust. The slopes of the distribution of stars having E(B − V ) ≤ E(B − V )min on the TCDs can give infor- mation about the foreground reddening law. In the case of IC 1805 (and NGC 654), where E(B − V )min is 0.65 (0.77), we used stars having E(B − V ) ≤ 0.80 (0.85) to estimate the foreground reddening presuming that star having 0.65(0.77) ≤ E(B − V ) ≤ 0.80(0.85) are not much affected by the anomalous reddening law in the cluster re- gion. The colour excess ratios E(J − V )/E(B − V ), E(H − V )/E(B − V ), and E(K − V )/E(B − V ) obtained are −2.00 ± 0.32(−1.98 ± 0.28), −2.58 ± 0.35(−2.47 ± 0.27) and −2.80 ± 0.39(−2.82 ± 0.36) respectively, which sup- port a normal reddening law in front of the cluster IC 1805. We further combined data of all the clusters hav- ing stars with reddening E(B − V ) ≤ 0.50. We feel that the limit of E(B − V ) ≤ 0.50 safely excludes the redden- ing due to intra-cluster matter as the smallest E(B − V ) for the NGC 869 and NGC 884 is ≈ 0.50 ( e.g. Uribe et al. 2002). A least-squares fit to 8 data points having 0.01 ≤ E(B − V ) ≤ 0.50 gives the colour excess ratio E(J −V )/E(B −V ) = −2.23±0.32, E(H −V )/E(B −V ) = −2.62 ± 0.22, and E(K − V )/E(B − V = −2.79 ± 0.26. These colour excess ratios also indicate a normal fore- ground reddening law towards the direction of the clusters used in the present study. Several studies have pointed to the apparent concen- tration of stars with high R-values in the vicinity of star forming regions. This effect has, for example, in the η Carina nebula (Forte 1978, Th´e and Groot 1983), M 16 (Chini and Wargau 1990) and M 17 (Chini et al. 1980, Chini and Wargau 1998) and it may be presumed to be characterstic of many more HII regions ( Winkler 1997). Winkler (1997) compared the value of R obtained for hottest stars in the Galaxy (spectral type O8 or earlier), which can be considered as indicators of regions with re- cent star formation, and he confirms that in the majority of the cases the stars with large R indeed seems to be in the vicinity of star forming regions. On the basis of the above discussions we presume that the anomalous extinction law in the direction of cluster IC 1805 is due to the intra-cluster material. 6. Conclusions In the present work we have carried out a detailed study of the intra-cluster material inside the young open clusters within a range of galactic longitude 1230 < l < 1440. We used three methods (cf. Sect. 3.1) to derive E(λ − V )/E(B−V ) and the results agree with each other in a few percent. It is found that the behaviour of extinction varies from cluster to cluster. The main results are as follows; 2 1 0 0 1 2 3 4 Fig. 13. The normalized extinction curves Aλ/AV . The values written along the curves represent the value of Rcluster . 1. The extinction curves at shorter wavelengths depend upon the Rcluster while they converge for λ ≥ λJ . 2. The extinction behaviour in the case of NGC 654, NGC 869 and NGC 884 is found normal, whereas in the case of NGC 663 there is some tendency for a higher value of Rcluster. It is interesting to mention that in the case of NGC 663 and NGC 869 the extinction at λU is found different from the normal one. In the case of NGC 663 the extinction process in the U band seems to be less efficient, whereas in the case of NGC 869 the process is more efficient. 3. The cluster regions of IC 1502 and IC 1805 show anomalous reddening laws. In the case of IC 1502 the Rcluster ∼ 2.57 ± 0.16 which is in agreement with the value (∼ 2.49 ± 0.09) obtained by Tapia et al. (1991), whereas in the case of IC 1805 the present study yields Rcluster ∼ 3.53 ± 0.25 which is in agreement with the value (∼ 3.44) given by Kwon and Lee (1983). More near-IR data is needed for the clusters NGC 663, NGC 1502 and IC 1805, where an anomalous value of Rcluster is found, to study the reddening law in detail in the these clusters. Acknowledgements. This work is partly supported by the DST (India) and the JSPS (Japan). AKP is thankful to the staff of KISO observatory for their help during his stay there. Authors are grateful to Prof. Ram Sagar for useful discussions. Thanks are also due to Dr. John Mathis for his comments which im- proved the contents of the paper. Pandey et al.: interstellar extinction in open clusters 11 References Bessel, M. S. 1979, PASP, 91, 589 Cardelli, J. A., Clayton, G. C., & Mathis, J. S. 1989, ApJ, 345, 245 Chini, R., Elsasser, H., & Neckel, T. 1980, A&A, 91,186 Chini, R.,& Wargau, W.F. 1990, A&A, 227, 213 Chini, R.,& Wargau, W.F. 1998, A&A, 329, 161 DeGioia Eastwood, K., Throop, H., Walker, G., & Cudworth, K. M. 2001, ApJ, 549, 578 Elmegreen, B. G., & Lada, C. J. 1977, ApJ, 214, 725 Fitzpatrick, E. L. PASP, 1999, 111, 63 Forte, J. C. 1978, AJ, 83, 1199 He, L., Whittet, D. C. B., Kilkenny, D., & Spencer Jones J. H. 1995, ApJS, 101, 335 Johnson, H. L. 1965, ApJ, 141, 923 Johnson H. L. 1966, ARA&A, 4, 193 Johnson, H. L., 1968, in Nebulae and Interstellar Matter, ed. B. M. Middlehurst, & L. M. Aller (University fo chicago press), chap 5 Ishida, K. 1969, MNRAS, 144, 55 Kiszkurno, E., Ko los, R., Kre lowski, J., & Strobel, A. 1984, A&A, 135, 337 Koornneef, J. 1983, A&A, 128, 84 Kre lowaski, J., & Strobel, A. 1983, A&A 127, 271 Kre lowaski, J., & Strobel, A. 1987, A&A 175, 186 Kwon, S. M.,& Lee, S. W. 1983, JKAS, 16, 7 Lida, H., Whittet, D. C. B., Kilkenny, D., & Spencer Jones, J. H. 1995, ApJS, 101, 335 Mathis, J. S. 1990, ARA&A 28, 37 Martin, P. G., & Whittet, D. C. B. 1990 ApJ, 357, 113 Mermilliod, J. C. 1995, in Information & On-line Data in Astronomy, ed. D. Egret, & M. A. Abrecht (Kluwar Academic Press), 227 Pandey, A. K., Mahra, H. S., & Sagar, R. 1990, AJ, 99, 617 Pandey, A. K., Ogura, K., & Sekiguchi, K. 2000, PASJ, 52, 847 Pandey, A. K., Upadhyay, K., & Ogura, K., et. al. 2002 (In preparation) Phelps R. L., & Janes K. A. 1994, ApJS, 90, 31 Sagar, R., & Yu, Q. Z. 1989, MNRAS, 240, 551 Schmidt-Kaler Th. 1982, in Landolt-Bornstein, Neue Serie Gr.VI, Vol 2b (Springer, Berlin, Heidelberg, New York) 19 Tapia, M., Roth, M., Costero, R., & Navarro, S. 1984, Rev Mex A&A 9, 65 Tapia, M., Costero, R., Echevarria, J., & Roth M. 1991, MNRAS, 253, 649 Th´e, P. S., & Groot, M. 1983, A&A, 125, 75 Turner, D. G. 1994, Rev. Mex. A&A 29, 163 Uribe, A., Garc´ia-Varela, J., Sabogal-Mart´inez, B., et al. 2002, PASP, 114, 233 Wegner, W. 1993, AcA, 43, 209 Winkler, H. 1997, MNRAS, 287, 481 Whittet, D. C. B. 1977, MNRAS, 180, 29 Yadav, R.K.S., & Sagar, R. 2001, MNRAS, 328, 370
astro-ph/9502102
1
9502
1995-02-27T17:19:31
AGAPE, an experiment to detect MACHO's in the direction of the Andromeda galaxy
[ "astro-ph" ]
The status of the Agape experiment to detect Machos in the direction of the andromeda galaxy is presented.
astro-ph
astro-ph
AGAPE, an experiment to detect MACHO's in the direction of the Andromeda galaxy R. Ansaria, M. Auri`ereb, P. Baillonc, A. Bouquetd, G. Coupinotb, C. Couturese, C. Ghesqui`eref , Y. Giraud-H´eraudf , P. Gondolod, J. Hecquetb, J. Kapland, A.L. Melchiord, M. Monieza, J.P. Picatb and G. Soucailb a LAL, Universit´e Paris Sud, Orsay, France, b Observatoire Midi-Pyr´en´ees Bagn`eres de Bigorre et Toulouse, France, c CERN, Gen`eve, Switzerland, d LPTHE, Universit´es Paris 6 et 7, France, e DAPNIA, CEN Saclay, France, f LPC Coll`ege de France, Paris, France. Presented at the 17th Texas Symposium on Relativistic Astrophysics, Munchen germany, December 94, by J. Kaplan. PAR-LPTHE 95-07 The M31 galaxy in Andromeda is the nearest large galaxy after the Small and Large Magellanic Clouds. It is a giant galaxy, roughly 2 times as large as our Milky Way, and has its own halo. As pointed by A. Crotts[1] and independantly by some of us[2] M31 provides a rich field of stars to search for MACHO's in galactic halos by gravitational microlensing[3]. M31 is a target complementary to the Large Magellanic Cloud and the galactic bulge wich are used by the three current experiments[4, 5, 6]. It is complementary in that it allows to probe the halo of our galaxy in a direction very different from that of the LMC. Moreover, the fact that M31 has its own halo and is tilted with respect to the line of sight provides a very interesting signature : assuming an approximately spherical halo for M31, the far side of the disk lies behind a larger amount of M31 dark matter, therefore more microlensing events are expected on the far side of the disk. Such an asymmetry could not be faked by variable stars[1]. In other words, M31 seems very appropriate to detect brown dwarfs through microlens- ing. However, as very few stars of M31 are resolved, we had to develop an approach to look for microlensing by monitoring the pixels of a CCD, rather than individual stars[2]. The AGAPE collaboration has set out to implement this idea. MONITORING PIXELS In the case of a crowded field such as M31, the light flux Fpixel on a pixel comes from the many stars in and around it, plus the sky background. The light flux of an individual star, Fstar, is spread among all pixels of the seeing spot and only a fraction of this light, Fpixel = {seeing fraction} × Fstar, reaches the central pixel. If the star luminosity is amplified by a factor A, the pixel flux increases by : ∆Fpixel = (A − 1) {seeing fraction} Fstar (1) The amplification of the star luminosity allows an event to be detected if the flux on the brightest pixel rises sufficiently high above its rms fluctuation σpixel : Typically, in our simulations, we require Q to be larger than 3 during 3 consecutive expo- sures and larger than 5 for at least one of them. ∆Fpixel > Q σpixel (2) 1 EXPECTED STATISTICS. We have performed numerical simulations using the above detection criterium. As an- ticipated, the number of events we expect to be able to detect depends strongly both on the stability of the pixel and on the average seeing. The numbers in Table 1 below assume a two meter telescope, a field of view of 60 by 20 arcminutes (planed for a second generation experiment) centered on the center of M31, 1 arcsecond pixels, 30 minutes exposures, and 120 consecutive nights. The brown dwarf mass is taken to be 0.08 M⊙, and we assume a standard halo (see reference[2] for instance) with a local dark matter density of 0.3 GeV/cm3 (0.0075 M⊙/pc3) and a core radius of 5 kpc. Table 1: Expected number of events under various observing conditions seeing σpixel/Fpixel 1" 2" 2% 1% 0.5% 2% 1% 0.5% number of galactic events 9 15 27 3 7 11 number of M31 events 15 29 50 4 10 21 total 24 44 77 7 17 32 These numbers have to be compared with the 2 events per year expected by the EROS collaboration for brown dwarfs with masses of order 0.1 M⊙. It is clear from Table 1 that the number of detectable events depends crucially on the rel- ative flux fluctuation on the pixel, σpixel/Fpixel. To study the feasibility of the experiment, we have analyzed these pixel fluctuations in three series of real data (Table 2) : • 1) 82 images of the Large Magellanic Cloud (LMC) taken by the EROS collaboration • 2) 26 images of M31 taken with the one meter telescope at Pic du Midi, in collaboration with F. Colas (Bureau des longitudes, Paris) and J. Lecacheux (DESPA, Meudon) Table 2: The mean relative fluctuation obtained for the three series of images listed above. The numbers in the first column refer to the numbers in the list. Images relative fluctuation LMC EROS (• 1) M31 Pic (• 2) M31 Pic (• 3) M31 Pic (• 3) mirror size (meter) 0.4 1 2 2 pixel size (arcsec) 1.15 0.7 0.25 1 ("superpixels") 3% 1% 0.7% 0.23% • 3) 4 images of M31 taken by E. Davoust (OMP, Toulouse) with the 2 m telescope at Pic du Midi. In this case, the angular size of the pixels is very small (0.23"), and we have also considered a rearrangement in 4×4 "superpixels". The results given in Table 2 clearly show that the required photometric stability of pixels can be reached. Moreover, the analysis of the third series of data shows that pixels small compared to the seeing allow an efficient matching between images, whereas, once images are matched, superpixels are more appropriate for a stable photometry. DISCRIMINATING AGAINST VARIABLE STARS AND OTHER VARI- ABILITIES Variable stars should be the main background. The usual tools to discriminate against this background are available (symmetry, unicity, achromaticity of the light curve). Still, some points particular to our approach are discussed below. 2 Achromaticity. At first sight, one would think that there is no achromaticity as a star rising above a background of a different color will cause a color variation of the pixels involved. However, it is easy to show that when a star rises above the background in two color bands (say red and blue) then the ratio (Fpixel − hFpixeli)red (Fpixel − hFpixeli)blue = Fstarred Fstarblue (3) is constant in time during a microlensing event. High amplifications. To rise above the background an unresolved star needs a rather high amplification (hAi ∼ 6), which will exclude most variable stars. On the other hand small secondary maxima indicating unstable stars with occasional strong flares will not be discriminated. Further study is required in this respect. THE FIRST RUN OF AGAPE We were given 57 half nights of observation on the 2 meter telescope "Bernard Lyot" at Observatoire du Pic du Midi in the French Pyr´en´ees, from September 29 to November 24 1994. The field was 8′ × 8′ only, covered by 4 exposures on a 800 × 800 thin Tektronix CCD camera with pixels 0.3" wide. The data of this prototype run are currently under treatment. A key step of this treat- ment is the alignement of successive images both in position and in photometry. The photometric alignement is performed by linearly transforming the light flux of one image in such a way that the mean flux and the variance of the transformed image matches those of some reference image. The result of this alignement between images is illustrated in figure 1. After alignement the dispersion of the relative difference between the two images is 1.6%. This preliminary result is very encouraging as our alignement procedures are not yet optimized. AKNOWLEDGMENTS We thank The EROS collaboration, and E. Davoust who allowed us to use their data, as well as F. Colas, and J. Lecacheux with whom we took data on the 1 meter telescope at Pic du Midi. The help of F. Colas during our first observation run has been particularly appreciated. References [1] Crotts A. P. S. 1992. ApJ. 399; L43. [2] Baillon P. Bouquet A. Giraud-H´eraud Y. & Kaplan J. 1993. A&A 277; 1. [3] Paczy´nski B. 1986. ApJ. 304; 1. [4] Alcock et al., 1993. Nature 365; 621, and these proceedings. [5] Aubourg E. et al. 1993. Nature 365; 623. [6] Udalski A. et al. 1993. Acta Astronomica 43; 289. 3 a b c Figure 1: photometric alignement of two different images of the same field. Histogramms of pixel flux of two images, a: before treatment, b: after photometric aligement. c: Histogram of the difference of pixel fluxes between the two images, after photometric alignement. 4
astro-ph/9908238
3
9908
1999-09-06T17:05:05
An HST Snapshot Survey of Proto-Planetary Nebulae Candidates: Two Types of Axisymmetric Reflection Nebulosities
[ "astro-ph" ]
We report the results from an optical imaging survey of proto-planetary nebula candidates using the HST. We exploited the high resolving power and wide dynamic range of HST and detected nebulosities in 21 of 27 sources. All detected reflection nebulosities show elongation, and the nebula morphology bifurcates depending on the degree of the central star obscuration. The Star-Obvious Low-level-Elongated (SOLE) nebulae show a bright central star embedded in a faint, extended nebulosity, whereas the DUst-Prominent Longitudinally-EXtended (DUPLEX) nebulae have remarkable bipolar structure with a completely or partially obscured central star. The intrinsic axisymmetry of these proto-planetary nebula reflection nebulosities demonstrates that the axisymmetry frequently found in planetary nebulae predates the proto-planetary nebula phase, confirming previous independent results. We suggest that axisymmetry in proto-planetary nebulae is created by an equatorially enhanced superwind at the end of the asymptotic giant branch phase. We discuss that the apparent morphological dichotomy is caused by a difference in the optical thickness of the circumstellar dust/gas shell with a differing equator-to-pole density contrast. Moreover, we show that SOLE and DUPLEX nebulae are physically distinct types of proto-planetary nebulae, with a suggestion that higher mass progenitor AGB stars are more likely to become DUPLEX proto-planetary nebulae.
astro-ph
astro-ph
An HST Snapshot Survey of Proto-Planetary Nebulae Candidates: Two Types of Axisymmetric Reflection Nebulosities Toshiya Ueta and Margaret Meixner Department of Astronomy, MC-221, University of Illinois at Urbana-Champaign, Urbana, IL 61801, [email protected], [email protected] Matthew Bobrowsky Orbital Sciences Corporation, 7500 Greenway Center Drive, #700, Greenbelt, MD 20770, [email protected] ABSTRACT We report the results from an optical imaging survey of proto-planetary nebula candidates using the Hubble Space Telescope (HST). The goals of the survey were to image low surface brightness optical reflection nebulosities around proto-planetary nebulae and to investigate the distribution of the circumstellar dust, which scatters the star light from the central post-asymptotic giant branch star and creates the optical reflection nebulosities. We exploited the high resolving power and wide dynamic range of HST and detected nebulosities in 21 of 27 sources. The reduced and deconvolved images are presented along with photometric and geometric measurements. All detected reflection nebulosities show elongation, and the nebula morphology bifurcates depending on the degree of the central star obscuration. The Star-Obvious Low-level-Elongated (SOLE) nebulae show a bright central star embedded in a faint, extended nebulosity, whereas the DUst-Prominent Longitudinally-EXtended (DUPLEX) nebulae have remarkable bipolar structure with a completely or partially obscured central star. The intrinsic axisymmetry of these proto-planetary nebula reflection nebulosities demonstrates that the axisymmetry frequently found in planetary nebulae predates the proto-planetary nebula phase, confirming previous independent results. We suggest that axisymmetry in proto-planetary nebulae is created by an equatorially enhanced superwind at the end of the asymptotic giant branch phase. We discuss that the apparent morphological dichotomy is caused by a difference in the optical thickness of the circumstellar dust/gas shell with a differing equator-to-pole density contrast. Moreover, we show that SOLE and DUPLEX nebulae are physically distinct types of proto-planetary nebulae, with a suggestion that higher mass progenitor AGB stars are more likely to become DUPLEX proto-planetary nebulae. Subject headings: stars: AGB and post-AGB -- stars: mass loss -- planetary nebulae: general -- reflection nebulae 1. Introduction Intermediate mass stars (initial main sequence mass of 0.8 − 8.0M⊙) evolve through a transitional proto-planetary nebula (PPN) phase between the asymptotic giant branch (AGB) and planetary nebula (PN) phases (Iben & Renzini 1983). A PPN consists of a cool (Teff <∼ 104 K) post-AGB stellar core and an extensive circumstellar shell of gas and dust, which is the former stellar envelope ejected through wind mass -- 2 -- loss1. In the PPN phase, AGB mass loss is assumed to have ceased but photoionization of circumstellar matter is considered not to have been initiated (cf. Kwok 1993). The PPN phase is still poorly studied because PPNe are statistically rare objects in the sky due to (1) their characteristically short time scale of evolution (∼ 103 years) with respect to a typical AGB evolutionary time scale (∼ 106 years) and (2) technological constraints imposed by PPNe's small angular scale and the presence of dust grains. Despite this rarity, approximately 100 PPN candidates have been identified from optical and infrared studies (e.g., Hrivnak, Kwok, & Volk 1989, Oudmaijer et al. 1992). One of the most significant changes that occurs to stars during the transition is the emergence of axisymmetry. It is observationally established that most OH/IR stars, whose stellar cores are AGB stars (Goldreich & Scoville 1976), show a high degree of spherical symmetry (cf. Habing & Blommaert 1993), while most (80%) planetary nebulae display either bipolar or elliptical symmetry2 (Zuckerman & Aller 1986). Therefore, the departure from spherical symmetry must take place somewhere along the evolutionary sequence between the two phases. The aspherical shaping of PNe has been qualitatively explained by the interacting stellar winds (ISW) model. In this framework, the fast wind (>∼ 103 km/s) is expected to "snow plow" slowly coasting (>∼ 10 km/s) circumstellar material which was ejected from the stellar envelope during the previous mass loss epoch (Kwok 1982). Then, axisymmetry can be imposed by introducing the notion of the equatorial density enhancement in the mass loss ejecta (Kahn & West 1985) and distinct morphological groups can be created by varying the degree of equatorial enhancement in red giant or AGB ejecta (e.g., Balick 1987, Habing et al. 1989, Mellema & Frank 1995). Although there has been a number of suggestions for the source of equatorial density enhancement (which includes magnetic fields and binary companions to name a few; e.g., Soker 1998, Mastrodemos & Morris 1999), there is no definite solution to the problem. Whichever the true scenario may be, a significant portion of the entire mass loss history is imprinted on the PPN circumstellar shell of gas and dust: the innermost edge defines the termination of mass loss and the mass loss history can be traced back in time as one probes outer regions of the circumstellar shell. Therefore, one can investigate when and how geometry of mass loss departs from spherical symmetry by sampling dust/gas distribution at various radial locations in a PPN circumstellar shell. One must, however, employ techniques that are sensitive to neutral gas (molecular line emission in radio) and dust (thermal emission in infrared and scattering of star light in visible) since no photoionization has taken place in the PPN circumstellar shell. In order to sample the most recent mass loss history one needs to probe the innermost regions of a PPN circumstellar shell. Previous ground-based investigations of PPN circumstellar shells include mid-infrared imaging of the thermal emission from warm (∼ 100K) dust grains (e.g., Meixner et al. 1999), optical imaging of the reflection nebulosity (e.g., Hrivnak et al. 1999b), and spectropolarimetry of the dust-scattered star light (Schmidt & Cohen 1981, Johnson & Jones 1991; Trammell, Dinerstein, & Goodrich 1994, hereafter TDG94). All of these studies have shown the axisymmetric nature of the innermost regions of PPN circumstellar shells. Recent work on HST imaging of PPNe (Egg Nebula, Sahai et al. 1998; IRAS 17150−3224, Kwok, Su, & Hrivnak 1998; IRAS 17441−2411, Su et al. 1998) have displayed spectacular images of bipolar reflection nebulosity but concentrated on a few PPNe that are associated with mostly or entirely obscured central stars. Our HST survey of PPNe covers a larger number of targets than any of the previous studies and, more importantly, includes PPNe candidates that are associated with bright central 1We reserve the word "shell" to refer to the circumstellar material that is physically detached from the central star to avoid confusion with the word "envelope," with which we refer to the mantle of a star. 2We consider bipolar and elliptical morphologies form mutually exclusive sets (see discussions below). -- 3 -- stars. Our main goal in this survey is to detect small and faint reflection nebulosities around PPN candidates at various evolutionary stages and to investigate if there exists any coherent morphological trend that will bridge gaps between the circumstellar shell morphologies in the AGB and PN phases. In this paper, we report results of our survey of optical reflection nebulosities around 27 PPN candidates, in which the high resolving power and wide dynamic range of HST are both exploited to the fullest extent. In the following sections, procedures of observations and data reduction are summarized (§2), results are presented (§3), and the physical nature of morphological groups that we find and the subsequent implications in the context of the PPN evolution are discussed (§4) with conclusions (§5). 2. Observations 2.1. Modes of Observation The 27 PPN candidates were observed with the Wide Field and Planetary Camera 2 (WFPC2) on-board HST between 1996 April and 1997 August (Program IDs 6364 and 6737) and were all acquired in the Planetary Camera (PC) chip (f/28.3, 0′′.0455 pixel−1). Observed coordinates for each object are listed in Tables 1, 2, and 3. To obtain high resolution images, each source was observed with a two- or three-point linear dithering pattern (the telescope is linearly shifted with a non-integer multiple of pixels). The dithering technique is now widely used in WFPC2 observations to fully exploit the high resolving power of HST from undersampled WFPC2 images. In conjunction with the dithering, the objects were observed with a set of different exposure times to cover the wide dynamic range that we required in our images. To image faint nebulosities with a sufficiently high signal-to-noise ratio, observations were made with intentionally long exposure times. For example, IRAS 22272+5435 was observed with 0.11, 1.2, 16, and 120 sec exposures. As a consequence, pixels in the vicinity of the central star were often saturated or contaminated by anomalies (e.g., bleeding, ghosts). The short exposure frames, which would be free of saturated pixels, were used to recover the loss of information in such pixels. In order to maximize the coverage of the dynamic range, we used a gain of 15 electrons per Data Number (DN) in our observations. We typically used two wide band filters, F555W and F814W (WFPC2 equivalent of the Johnson-Cousins bands, VJ and IC; e.g. Holtzman et al. 1995), for each source. However, F450W (Wide B) filter was used when a source had previously been observed with F555W. For extremely bright sources, we used medium and narrow band filters, F410M (Stromgren v), F547M (Stromgren y), and F469N (He II), to avoid saturation in the shortest exposure frames. In total, approximately a dozen raw images were obtained for each source and filter. 2.2. Data Reduction We used IRAF/STSDAS routines in data reduction. A standard HST pipeline calibration was performed with the latest reference files available at the time of data reduction. Duplicate frames of dithered images were combined into a single image by applying the variable-pixel linear reconstruction algorithm ("drizzle" package v1.2, Fruchter & Hook 1997), which would interlace each pixel in multi-point dithered frames according to the statistical significance of each pixel. The drizzled images were subpixelized -- 4 -- (0′′.0228 pixel−1) during the process and thus the high spatial resolution was recovered. Cosmic-rays were removed by the drizzling algorithm in the three-point dithered frames, while they were eliminated manually by replacing the contaminated pixels with a median of the neighboring pixels in the two-point dithered frames. In order to create non-saturated final source images, the saturated and unsaturated frames were combined by replacing saturated pixels with unsaturated ones from the shorter exposures, scaled by the exposure times. We have thus successfully obtained high dynamic range images of reflection nebulosities whose outer perimeters often seem to be sky-limited. Figures 1, 2, and 3 show the reduced images. The resulting nebular signal-to-noise ratio ranges from <∼ 2 for a faint nebula to > 100 for a bright nebula, while the star-to-nebula ratio, which quantifies the emission contrast between the central star and nebulosity as a measure of the dynamic range, ranges from 18 for a compact nebula up to 1.8 × 105 for an extended nebula. Some faint nebulae are barely distinguishable from the background sky, while some others are almost buried under the point spread function (PSF) of the central star itself. 2.3. Image Enhancements The reduced images, particularly those with a bright central star, were affected by the WFPC2 diffraction patterns that consist of linear spikes and circular wings. We employed image enhancement techniques to remove unwanted effects of the WFPC2 point spread function (PSF) so that the reflection nebulae would be more clearly seen. One could remove the effects of the WFPC2 PSF by either (1) subtracting a stellar PSF from a source image or (2) deconvolving a source image with a stellar PSF. A PSF subtraction can be done with either an observed PSF or a synthesized PSF. Although observed PSFs were often preferred (Krist et al. 1997), we used synthesized PSFs because there existed no PSF with a high enough dynamic range in proper filters for our observations. Model WFPC2 PSFs were generated by the code Tiny Tim (v4.4, Krist & Hook 1997) for a given mode of observation, but the PSF had to be scaled to account for the drizzling. Alternatively, a deconvolution technique can be used to eliminate the PSF effects in the reduced source images. We used the Richardson-Lucy (RL) algorithm and maximum entropy method and the former generally yielded better results than the latter. This suggests that the most significant noise in our images was due to photon shot noise because iterative solutions yielded by the RL algorithm are known to converge into the maximum likelihood solution in Poisson statistics (Shepp & Vardi 1982). In general, the contrast between the nebula and sky emission at the outer perimeter of a nebula was increased by a factor of >∼ 2; however, there seemed to be little improvement in a few very extended nebulosities where the nebula emission was comparable to the sky emission. Unlike the case of PSF subtraction, a deconvolution technique could be applied to all our images. One shortcoming in RL deconvolution is that the algorithm is known to amplify uncertainties and generate false depressions around pixels with unusually high DNs, and such "holes" indeed appeared in the deconvolved images. Therefore, we have to keep in mind that any interior structure in the deconvolved images should be regarded as suspect. Overall, we found that the RL deconvolved images provided the best removal of the PSF effects among all the methods we tried. Thus, we present only the RL deconvolved images alongside the original reduced images in Figures 1, 2, and 3. -- 5 -- 2.4. Measurements Photometric and geometric quantities were measured from the reduced source images. To give the total specific flux density (Fλ in ergs cm−2 s−1 A−1), the reduced images were flux calibrated by adopting the HST photometric calibration of SYNPHOT (v4.0, Simon 1997). First, we defined a photometric aperture that was large enough to encircle the entire source as well as the diffraction features. The total DN of a source was then determined by summing all DNs within the aperture. The background emission per pixel was estimated by calculating the averaged sky DN inside a 10-pixel-wide annulus that encloses the aperture but was separated from the aperture by a buffer zone. The background DN, which is the averaged sky DN multiplied by the number of pixels in the aperture, was subtracted from the total DN only when the averaged sky DN per pixel was greater than the root mean square of the sky DNs in the background annulus because otherwise the background subtraction would introduce an additional 1σ uncertainty to the results. We expect that an uncertainty due to the background subtraction is rather insignificant because considerably larger DNs in the emission core will dominate the total emission of the source and thus photon shot noise will dominate. The total source DN was then converted into Fλ and the WFPC2 system magnitude (STMAG). The extent of the nebulosities was estimated from the images by defining the "edge" of a nebula to be the outermost recognizable structure, in which the emission level turned out to be 1σ up to about 7σ of the sky depending on the quality of the image. The major and minor axes of the nebula were measured and the ellipticity of a nebula was derived by e = 1 − b/a (a and b are respectively major and minor axis lengths). With the edge of a nebula being defined, we can also measure the surface intensity (ergs cm−2 s−1 A−1 sr−1) at the peak and edge of the nebula, from which we can obtain the star-to-nebula surface intensity ratio as a measure of the width of the dynamic range covered by the image. When the central star is visible, the peak coincides with the location of the central star, but, when the central star is totally obscured, the peak is simply the local maximum in the emission region. All derived quantities are summarized in Tables 1, 2, and 3. 3. Results 3.1. Images of Reflection Nebulosity Of 27 PPN candidates, 21 were found with fascinating reflection nebulosities around the central stars and six did not seem to be associated with any nebulosity. By inspection, one immediately realizes the following: (1) all 21 nebulae show asphericity with varying degrees and (2) there clearly exist two types of axisymmetry among those aspherical nebulosities. One type of nebulosity is characterized by its very low surface brightness and multi-axis elongations which surround extremely bright central stars. The other type, however, is distinguished by the limb-brightened bipolar lobes with their partially or completely invisible central stars somewhere in the nebulae. Because this apparent bifurcation is so astonishing in the morphology of reflection nebulosity, we categorize the two types according to the traits in appearance described above and refer to the former type as the Star-Obvious Low-level-Elongated (SOLE) nebulae while the latter type as the DUst-Prominent Longitudinally-EXtended (DUPLEX) nebulae. Among 21 nebulosities, 11 and 10 are respectively found to be SOLE and DUPLEX nebulae in this classification. One of the key differences between SOLE and DUPLEX sources is the star-to-nebula surface intensity ratio (Table 1 and 2), which is a useful quantity to estimate the size of the dynamic range required to -- 6 -- observe SOLE nebulae. For sources with the visible central star (all SOLE and DUPLEX nebulae with the partially visible central star), the dynamic range varies from 18 (for the most compact SOLE source, IRAS 07430+1115) to 1.8 × 105 (for the most extended SOLE source, IRAS 19114+0002) with the average value of 1.3 × 104. On the other hand, for sources with the obscured central star (bipolar DUPLEX nebulae), the dynamic range is at most 330 (IRAS 20028+3910, whose northern lobe is barely detected) with the average value of 55. In recent subarcsecond optical imaging of PPNe by ground-based telescopes, four of our sources (IRAS 18095+2704, IRAS 19374+2359, IRAS 20028+3910, and IRAS 22574+6609) were observed to determine their morphology (Hrivnak et al. 1999b). However, they were unable to determine if IRAS 18095+2704 (the only SOLE source among the four) is extended partly due to the brightness of the star (V = 10.3) despite the suggestion of its extension from the FWHM of their V image. Because of the wide dynamic range we achieved (∼ 1000), our IRAS 18095+2704 images clearly show that it is indeed an extended source. Therefore, to detect and image faint, extended reflection nebulosities around a bright central star, i.e., the SOLE nebula, a very wide dynamic range must be used. Even the six orders of magnitude coverage barely detects the outermost structure in images of IRAS 19114+0002 and there may be even fainter, more extended nebulosity. Although we classify the objects mainly on morphological grounds, spectral energy distribution (SED) and two-color diagrams provide supplemental information to determine the morphological class of an object. The sources that are not considered to be associated with any nebulosities are referred to as stellar sources (see discussions below). We adopt these terms to address each morphological type hereafter and will discuss the morphological dichotomy in detail in the following subsections. Figures 1(a) to 1(c), 2(a) to 2(c), and 3 show SOLE, DUPLEX, and stellar sources, respectively3. 3.1.1. SOLE Nebulae The SOLE nebulae show the very bright central star embedded in a very low surface brightness nebulosity (Fig. 1). This type of reflection nebulae has been imaged for the first time by this survey: the central star is so bright that the object would always appear as a point source unless the observations are done with the high enough resolution (available with HST) and/or wide enough dynamic range (available with our method of multiple exposure times). The very eye-catching trident-like structures emanating from the central star are WFPC2 linear diffraction spikes and are not to be construed as real structure. The reader is encouraged to compare the images of the SOLE nebulae (Fig. 1) with those of the stellar sources (Fig. 3) to help the eye differentiate real structures from remnants of the PSF artifacts such as the diffraction spikes and circular halos. The morphology of the SOLE nebulae can be further subdivided into groups: a simple ellipse, multiple ellipses (more than one ellipse superposed onto one another with differing orientations of the major axes), an ellipse with embedded bipolar structure, and an ellipse with concentric shells. IRAS 07134+1005, IRAS 17436+5003, and IRAS 20462+3416 all have large (<∼ 4′′) and faint nebulosities. The size of the optical reflection nebulosity in IRAS 07134+1005 is the largest among the 21µm feature sources (e.g. Hrivnak & Kwok 1991a) and is comparable to mid-infrared images (Meixner et al. 1997). The extended nature of IRAS 17436+5003 was suspected from wings of 12CO (J=2-1) and 13CO (J=2-1) line profiles (Bujarrabal, Alcolea & Planesas 1992). IRAS 20462+3416, a young PN which has 3All fully reduced and deconvolved images are also available in FITS format on the World Wide Web's NCSA Astronomy Digital Image Library (ADIL, at http://imagelib.ncsa.uiuc.edu/document/99.TU.01/). -- 7 -- already started showing low extinction characteristics (Parthasarathy 1993, Smith & Lambert 1994), was observed to have experienced a brief period of an enhanced mass loss between 1993 and 1995 (Garc´ıa-Lario et al. 1997b). IRAS 02229+6208 and IRAS 07430+1115 have smaller (2′′ and 1′′) nebulae. Both of these sources, along with IRAS 05341+0852, have recently been observed by Hrivnak & Kwok (1999), but they were unable to determine if the sources are extended due to poor seeing. IRAS 04296+3429, IRAS 05341+0852, and IRAS 22272+5435 have two axes of elongation that are not perpendicular to each other. Despite a circular nebula prediction because of its shape of the SED (Hrivnak & Kwok 1991a, Hrivnak et al. 1999b), IRAS 04296+3429 shows a complex double-elongation structure: the secondary E-W elongation of IRAS 04296+3429 is close to but not aligned with a diffraction spike and is likely to be real. This is, however, consistent with a suggestion that the source is associated with axisymmetrically distributed, optically thin dust (TDG94). IRAS 05341+0852 shows a diffuse elongation in the NE-SW direction and there seems to be a secondary elongation inside of and tilted about 20◦ counter-clockwise from the primary one. IRAS 22272+5435, whose axisymmetric nature was already seen by spectropolarimetry (TDG94), has a bright, large core with the four elliptical tips which create an almost amoeba-like appearance for the nebula. The northern and southern elliptical tips are of equal brightness, but the western tip is 2.6 times fainter than the eastern tip, which is approximately 1.6 times brighter than the northern and southern tips: this suggests that the E-W elongation is tilted (with the eastern lobe being closer to us) but the N-S elongation is not tilted. IRAS 06530−0213 and IRAS 18095+2704 have rather peculiar structures. In addition to the well-defined elongation, both sources display an inner structure, which seems to be a limb-brightened bipolar lobes. IRAS 18095+2704 shows a similar spectropolarimetric trend as seen in IRAS 04296+3429 (TDG94), which may be related to a secondary jet-like structure that extends in the NE-SW direction. However, its very bright central star (VWFPC2 = 10.30) and poor seeing prevented Hrivnak et al. (1999) from resolving its extension. IRAS 19114+0002 shows rich structure: there are at least four inner concentric shells (11′′, 7′′, 4′′, and 3′′) with some protuberance and one very sharp elongation (8′′.5) about 15◦ East from North. The 12CO (J=2-1) map shows very extended structure, which seems to have been shifting its direction: 10% contour points to 72◦ East from North (37′′), while 50% contour points to 45◦ East from North (18′′). The protuberance seen in our images may have emanated from the same rotating point of origin. The axisymmetric nature has also been seen in polarimetric observations (TDG94). Although the sharp elongation suggests a rather large inclination angle, IRAS 19114+0002 is believed to be close to a pole-on orientation, which is evidenced by a hollow shell structure seen in both mid-infrared (Hawkins et al. 1995) and near-infrared polarimetric (Kastner & Weintraub 1995) imaging studies. 3.1.2. DUPLEX Nebulae The DUPLEX nebulae are recognized either by their magnificent bipolar nebulosities or by rather well-defined limb-brightened bipolar lobes (Fig. 2). They are usually outlined by a lower surface brightness halo. These nebulae differ from the SOLE nebulae in appearance primarily because their central stars are partially or completely obscured. The diffraction spikes are not usually an issue in the images of DUPLEX nebulae as the central stars are obscured from the direct view. The DUPLEX nebulae can also be further subdivided into two groups depending on the presence or absence of the central star. IRAS 16342−3814, IRAS 17150−3224, IRAS 17441−2411, IRAS 20028+3910, and IRAS 22574+6609 show multiple emission peaks without clear indications of the central star's whereabouts. Among those, IRAS 17150−3224 and IRAS 17441−2411 are found with comparable lobes both in size and brightness -- 8 -- and their lobes possess some inner structure that seems to be point symmetric. There are also thin concentric arcs extending beyond the perimeters of the lobes. The arcs appear to be created independently of the lobes because the arcs maintain the same emission level both in and out of the lobes and hence seem to be unaffected by the presence of the lobes. The intervals between arcs have been estimated to be too short to be caused by the consecutive AGB thermal pulses (Paczy´nski 1975). Our B and I band images of IRAS 17150−3224 and IRAS 17441−2411 confirm the findings by previous HST wide V band observations (Kwok, Su, & Hrivnak 1998, Su et al. 1998). Weintraub et al. (1998) obtained H2 emission profiles from these sources and confirmed their orientations suggested from the HST optical images. They even also found evidence of an expanding torus in IRAS 17150−3224, which was shown in a V -- I image constructed from ground-based observations (Kwok et al. 1996). IRAS 16342−3814 and IRAS 20028+3910 have unequal lobes in which some inner structure is recognized in the primary lobe. The bipolar nature of IRAS 16342−3814, an extreme AGB star, was revealed in H2O and OH maser observations (Likkel & Morris 1988), and recent VLA observations of OH maser determined its inclination angle to be about 40◦ (Sahai et al. 1999). IRAS 20028+3910 has recently been reported to be extended (2′′.2 × 2′′.0, Hrivnak et al. 1999b), but this only corresponds to the S lobe (which is about 25 times brighter than the N lobe) of this bi-lobal object. Interestingly, the deconvolved images of IRAS 20028+3910 show multiple-peaks within the S lobe. IRAS 22574+6609 is optically resolved for the first time and the images indicate the presence of more than two emission peaks, confirming an earlier suggestion of elongation (Hrivnak & Kwok 1999a). The emission level of the suspected third emission peak (0′′.2 north of the second peak) is almost the same as that of the background sky and thus its presence is inconclusive. This "third" peak may simply be a part of a clumpy second peak, in which case the overall appearance of the source resembles that of IRAS 17441−2411. Our V band photometry (VWFPC2 = 21.24) differs from that of Hrivnak & Kwok (1991a; VJ = 24). This difference is significant enough to mention, even though we are comparing magnitudes in slightly different systems. Their lower magnitude suggests that it may have been affected by the unusually poor seeing (Hrivnak & Kwok 1991a), or this star may have been experiencing a significant brightening. IRAS 08005−2356, IRAS 17423−1755, and IRAS 19374+2359 have the partially visible central star with limb-brightened bipolar lobes which appear as a pair of horseshoe structures facing each other along the bipolar axis. Slijkhuis et al. (1991) observed an unusually broad Hα line profile in IRAS 08005−2356 and attributed it to a fairly extended emission region. This interpretation is independently supported by spectropolarimetry, which shows an abrupt position angle shift suggesting an optically thick dust torus and optically thin reflection lobes (TDG94). Both of the above views are confirmed by the bipolar shape clearly seen in our images. Its SE lobe is approximately 3 times brighter than the NW lobe, which suggests that the SE lobe is tilted towards us so that the central star becomes partially visible within the conical opening angle of the lobe. IRAS 17423−1755 displays fascinating point symmetric jet-like structures extending 17′′ in the whole stretch. The NW lobe is more prominent (8 times brighter) than the SE lobe, whose presence can be traced with the help of the slightly visible, outer part of the limb. This suggests that the NW lobe is inclined towards us, again explaining the partial view of the central star. The way the SE lobe is obscured strongly suggests the presence of a dust torus between the lobes. This interpretation agrees with a model in an earlier multi-wavelengths study, in which fast, collimated jets punctured a detached shell causing a torus-like shell structure (Bobrowsky et al. 1995). Although less prominent, there are at least three knots in each of the point symmetric jets as seen previously (Bobrowsky et al. 1995, Riera et al. 1995, Borkowski, Blondin, & Harrington 1997). A hydrodynamic simulation shows a diverging outflow being focused into a narrow jet and the point symmetric structure can be explained by wobbling jets (Borkowski, Blondin, & Harrington 1997). IRAS 19374+2359 was observed by Hrivnak et al. (1999b) and a round extension (2′′.6) is seen. This corresponds to the outer halo in our images. Although our images of IRAS 19374+2359 have -- 9 -- smaller signal-to-noise ratio compared with other images, one can discern the star from the nebula in the northern lobe, which is 3 times brighter than the southern counterpart, suggesting that the northern lobe is pointing towards us. The two remaining DUPLEX sources, IRAS 09452+1330 (IRC+10216) and IRAS 23321+6545, do not neatly fit into either of the above subdivisions. IRAS 09452+1330 is the best studied C-rich AGB star in the Galaxy. The I band image was previously published (Skinner, Meixner, & Bobrowsky 1998) and is included in this survey for the sake of completeness. The aspherical appearance of IRAS 09452+1330 has been interpreted as a bipolar nebula whose southern lobe is pointed towards us, being separated from smaller northern lobes by a dust lane, and the bright point-like source in the southern lobe may be the central star (Skinner, Meixner, & Bobrowsky 1998). Thus, we classify IRAS 09452+1330 as a DUPLEX source because its reflection nebulosity appears quite similar to that of DUPLEX PPNe. We also observed the source with the wide B filter but did not detect anything. The optical counterpart to IRAS 23321+6545 is imaged for the first time. Its very small spatial extent suggests that this object is located relatively far away. However, the fact that this distant source appears extended alternatively indicates that the nebulosity has rather high surface brightness compared to the central star, which is a typical characteristic of DUPLEX nebulae. If IRAS 23321+6545 were a SOLE nebula, the PSF of the bright central star would have masked any structure of the fainter, compact nebula and the source would have appeared as a point source. Therefore, IRAS 23321+6545 must possess DUPLEX structure, possibly the one with the partially visible central star. 3.1.3. Stellar Sources Figure 3 shows the sources lacking clear indications of the presence of a nebulosity. IRAS 04386+5722, IRAS 20043+2653, and IRAS 22142+5206 only have the diffraction features and no evidence of extended emission regions. There is no deconvolved image displayed for both IRAS 20043+2653 and IRAS 22142+5206 because all of the frames were saturated and reconstruction of the non-saturated peaks was not possible. Ghosts appear in the images of IRAS 04386+5722 as the "double dots" about 2′′ east of the star: they are double because each ghost appeared at different chip locations in the dithered frames. IRAS 05113+1347 does not seem to have any extended nebula, however, the deconvolved images leave rather high residual DNs within 0′′.4 of the central star where a Tiny Tim PSF is able to simulate PSF effects rather well (Krist et al. 1997). Because of its small angular extent, it is inconclusive whether this is real or not. IRAS 10158−2844 and IRAS 15465+2818 are considered to have had little recent mass loss but are associated with very diffuse, extended circumstellar dust shells (Gillett at al. 1986, Waters et al. 1989). Our images are consistent with this picture of little mass loss by showing no apparent nebulosity. Because these stars are very bright (hence the use of a narrow band filter), the diffraction features are more prominent in these images than in other images. 3.2. Measured Quantities and Binary Companions Tables 1, 2, and 3 summarize the measured quantities for SOLE, DUPLEX, and stellar sources, respectively. The quantities are the total specific flux densities (Fλ in ergs s−1 cm−2 A−1), WFPC2 system magnitudes (STMAG), peak intensities, star-to-nebula intensity ratios, sizes, and ellipticity. Table 2 is subdivided into two sections according to the visibility of the central star: the peak intensity represents the intensity of the star when the location of the central star in the nebula is certain (top 4 objects) whereas -- 10 -- it represents the local intensity maximum in the emission region when the central star is completely or partially obscured (the rest). For images that show clear bipolar structure with halo, the size and ellipticity are measured for the entire halo as well as for each lobe. We only list photometric quantities for stellar sources (Table3). All VWFPC2 are generally in good agreement with the previously published VJ in the literature (δV ≈ 0.18 mag) where the STMAG closely resembles the Johnson system, except for two sources (IRAS 16342−3814 and IRAS 22574+6609). Comparison between IWFPC2 and available IC in the literature suggests that IWFPC2 is generally ∼ 1.3 dimmer than IC. This offset is due to the definition of STMAG system and the amount of offset is about equal to what is expected by definition in SYNPHOT (Simon 1997). Similarly, magnitudes obtained with other filters deviate from the values in the literature due to the definition of the STMAG system. Our VWFPC2 of IRAS 16342−3814 (15.64) agrees with another measurement independently made by Sahai et al. (1999) from the same image (15.7). However, these values are significantly dimmer than the value reported in a PPN photometric survey (13.65; Van der Veen, Habing & Geballe 1989, hereafter VHG89). A large photometric aperture used by them is suspected to have included nearby bright stars (Sahai et al. 1999). On the other hand, their near-infrared photometric values (J=12.17, H=10.75, K=9.61, done in 1986; VHG89) are significantly dimmer than those of yet another PPN photometric survey (J=9.29, H=8.32, K=7.71, done in 1993; Garc´ıa-Lario et al. 1997a). Because Garc´ıa-Lario et al. (1997) do not discuss their sources individually or give the observed coordinates, we are unable to assess the cause of the discrepancy. It is very unlikely that the central star of a PPN becomes dimmer in V and brighter in near-infrared. IRAS 22272+5435 is suspected to be a variable star with an almost 1 mag variation (Hrivnak & Kwok 1991b). Our measurement (VWFPC2 = 8.63) is consistent with its brightest magnitude. There seems to be no sign of any other unreported variability in our sources. The size of SOLE nebulae is the major and minor axis lengths of the elliptical elongation whereas the size of DUPLEX nebulae is the extent of the halo. Some SOLE nebulae have two axes of elongation. In such cases, we only list major axis lengths of the two elongations but not the minor axis lengths, though the ellipticity are given for each. The averaged ellipticity is rather high for both SOLE and DUPLEX nebulae (0.45 and 0.43, respectively). Here, the ellipticity is even larger for SOLE nebulae. This quantitatively confirms what we have seen in the images: reflection nebulosities are unquestionably elongated in their apparent shapes. We can also search for signs of binary companions in the vicinity of the sources, and there are several possible cases. IRAS 10158−2844 is seen with a star (WFPC2 He II mag = 13.70) which is about 2′′ east of the source. Although IRAS 10158−2844 is known to form a binary system of an orbital period of near 434 days with either a low-mass main sequence star or a white dwarf (Waelkens et al. 1991), this nearby star does not seem to be the companion of the binary system because the separation is too large (≈ 1500AU). Whether or not this nearby star is related to IRAS 10158−2844 is also not certain. IRAS 19114+0002 has a nearby star seen about 4.5′′ north of the source at the edge of the faintest nebulosity along one of the saturation spikes. However, the stars do not seem to be related due to rather large distance between the two. IRAS 19374+2359 has a nearby star (VWFPC2 = 20.39) inside the south lobe about 1′′ away from the central star, but it is likely to be a foreground or background star because of the low galactic latitude of the source (1◦). IRAS 22142+5206 also has a star (VWFPC2 = 20.36) about 2′′ east of the source, but the nature of the nearby star is not certain. IRAS 22142+5206 is now classified as a young stellar object embedded in a massive molecular cloud (∼ 7300M⊙; Dobashi et al. 1998). Their observations indicate that the source is associated with the most massive CO outflow (∼ 33M⊙) reported so far, and this may be because of the binarity of the source. -- 11 -- 4. Discussions 4.1. Axisymmetry: an Intrinsic Nature of PPNe We have detected optical reflection nebulosities in 21 sources out of 27 PPN candidates (78%) and all of these 21 PPN reflection nebulosities exhibit some type of axisymmetry with the averaged ellipticity of 0.44. Our direct imaging of reflection nebulosities confirms previously published results in a spectropolarimetric survey of post-AGB stars, which has indirectly shown that 24 of 31 sources (77%) are aspherical (TDG94). As we have discussed in the introduction, previous work in the literature has revealed that most PPN candidates possess axisymmetric nebulosities, and therefore, we conclude that the axisymmetry is an intrinsic trait of the PPN reflection nebulosities. Below, we discuss when this axisymmetry arises along the evolutionary track between the AGB and PN phases and how the two different types of PPN may arise. It is now generally accepted that the AGB phase is associated with two types of mass loss: an AGB wind (∼ 10 km/s) mass loss phase followed by a briefer but supposedly more violent superwind (∼ 20 km/s) mass loss phase (Renzini 1981). Because the termination of a superwind is considered to be the end of the AGB phase and no significant stellar wind is expected until the initiation of a fast wind (Kwok 1982), the PPN axisymmetry must arise just before the end of the AGB phase. This interpretation is also supported by the fact that mass loss is spherically symmetric in the beginning of the AGB phase (Habing & Blommaert 1993) and that some extreme AGB stars have already departed from spherically symmetric structure (e.g., IRAS 16342−3814). It is, therefore, very likely that a superwind is intrinsically axisymmetric and that the onset of a superwind initiates the morphological shift from spherical to axial symmetry in a PPN circumstellar dust/gas shell. Based on the results of our PPN survey with the above inference, we propose the following evolutionary scenario. In the AGB wind phase, an AGB star loses its mass through a dust-driven AGB wind (Salpeter 1974, Kwok 1975, Netzer & Elitzur 1993) in a largely spherically symmetric manner, creating a spherically symmetric circumstellar AGB wind shell. Towards the end of the AGB phase, some physical mechanism, albeit still unknown, comes to play and starts generating an equatorially density enhanced dust-driven wind, which we call an axisymmetric superwind. The axisymmetric superwind dumps the envelope material of the central AGB star preferentially on the equatorial plane, and a superwind shell with a torus-like density enhancement develops deep within the spherically symmetric AGB wind shell. The equatorial density enhancement in the superwind shell is further strengthened as the star evolves. At the end of the AGB phase, the superwind ceases and defines the inner boundary of a detached circumstellar dust/gas shell, which manifests itself as a mid- to far-infrared excess in the double-peaked SED of a PPN (Kwok 1993). This two-phased AGB mass loss scenario can be employed to explain an apparent "dust lane" obscuring the central star. In one-dimensional radiation transfer simulations, the superwind shell may be treated as a somewhat ad hoc addition to an otherwise spherically symmetric dust shell (e.g., Su et al. 1998). Meixner et al. (1997) have incorporated the two-phased mass loss in fully two-dimensional radiation transfer calculations and their synthesized mid-infrared images and SEDs agree with observations. However, the scenario used in these calculations is still a first order approximation: the transition from a spherically symmetric AGB wind to an axisymmetric superwind is assumed to take place abruptly. Instead, the transition is more likely to occur gradually because mass loss is essentially governed as a function of the fundamental stellar parameters, which do change gradually as the central star evolves if integrated over the -- 12 -- course of the entire AGB phase (e.g., Blocker 1995). Given that the details of the two-phased mass loss probably depend on the fundamental stellar parameters, the degree of equatorial density enhancement in a superwind may also be dependent upon them, and the physical environment will probably be distinct in each superwind shell. Figure 4 schematically describes how this can affect the structure of PPN circumstellar shells and may cause the morphological bifurcation of PPN reflection nebulosities. In a SOLE PPN (top), a marginal equatorial enhancement in the superwind shell can yield a dust torus that is optically thin (gray zone), and stellar photons can escape virtually in all directions (arrows). Hence an observer is able to see the bright central star embedded in an elliptically elongated nebula (dashed perimeter). In a DUPLEX PPN (bottom), on the other hand, a stronger equatorial enhancement in the superwind shell can result in a optically thick dust torus (black zone), and most photons are scattered off towards the biconical openings of the torus along its axis of symmetry (arrows), thereby generating a bipolar, dumbbell-like nebulosity (dashed perimeter). When the equatorial enhancement is exceptionally strong, the dust torus can be so flattened that it assumes the form of a thin disk. Therefore, we consider a disk to be an extremely equatorially enhanced torus. 4.2. SOLE vs. DUPLEX: Physically Distinct Nebulae The distinct appearances between SOLE and DUPLEX nebulae are characterized by the presence or absence of the central star and by the undisturbed elliptical or dumbbell-like outline of the nebulosity. We now discuss how the evident morphological dichotomy of the PPN reflection nebulosities can be caused by a physical difference in the circumstellar dust/gas shell and not by an inclination angle effect alone. More specifically, we propose that the optical morphology of PPN candidates bifurcates because the opacity in the circumstellar dust/gas shells varies due to differing degrees of equatorial density enhancement. Spectropolarimetric survey results of TDG94 show that there are two types of polarization position angle shift (a gradual shift in IRAS 04296+3429, a SOLE source and an abrupt shift in IRAS 08005−2356, a DUPLEX source) and that the presence of the optically thick and thin, two-component obscuring agent is suspected in PPNe with an abrupt position angle shift. This is consistent with the assumption of DUPLEX nebulae being associated with optically thick dust grains. In the following, we will present three additional pieces of evidence suggesting that the morphological dichotomy indeed corresponds to physically distinct nature of the circumstellar dust shell in PPN nebulae: the mid-infrared morphologies, SEDs, and IRAS/near-infrared colors. 4.2.1. Mid-Infrared Morphology of PPN Dust Shells Meixner et al. (1999) have recently observed 66 PPNe at mid-infrared wavelengths and directly imaged thermal dust emission regions in the circumstellar dust shell. The major discovery in the mid-infrared survey is the morphological bifurcation of dust emission regions: "core/elliptical" types have an extended low emission region surrounding a compact unresolved core, which is attributed to an optically thick equatorial density enhancement, while "toroidal" types show two emission peaks, which are interpreted as limb-brightened peaks of an optically thin, equatorial density enhancement. If we compare the mid-infrared and optical morphologies of PPNe, there is a one-to-one correspondence between the two morphologies -- 13 -- as is shown in Table 44. It appears that toroidals and core/ellipticals are strongly correlated with SOLE and DUPLEX nebulae, respectively. This correlation is consistent with the picture in which a dust optical thickness difference causes the PPN morphological bifurcation. That is, PPN candidates that are optically thin at visible wavelengths (SOLE nebulae) are also optically thin at mid-infrared wavelengths (toroidals), while PPN candidates that are optically thick at visible wavelengths (DUPLEX nebulae) are also optically thick at mid-infrared wavelengths (core/ellipticals). The ways in which the mid-infrared and optical images are spatially related in SOLE and DUPLEX nebulae differ, and hence, they also suggest a difference in dust shell optical thickness. By direct comparison between the mid-infrared and optical images of SOLE sources, the optical and mid-infrared nebulae are found to be spatially coincident and that the optical nebulosity is elongated perpendicularly with respect to the equatorial plane of the suspected dust torus, whose orientation is indicated by the two mid-infrared emission peaks. In Figure 5, for example, a resolved, deconvolved 11.8µm image of IRAS 07134+1005 shows limb-brightened dust emission peaks which are oriented in the east-west direction (Meixner et al. 1997), while its optical (Stromgren v) nebula is extended in the north-south direction (top left). A similar trend is seen in the composite image of IRAS 17436+5003 (top right; Skinner et al. 1994) as well. On the other hand, the mid-infrared and optical images of DUPLEX sources show elongation in the same direction, and the mid-infrared emission region is often completely embedded within the optical nebulosity. In Figure 5, for example, an I band image of IRAS 17150−3224 clearly displays the dust lane, which completely obscures the central star, between the two lobes and there is a very compact, unresolved mid-infrared emission core over the location of the dust lane (bottom left), and the unresolved dust emission core of IRAS 16342−3814 in 9.8µm is very compact with respect to the whole extent of the I band reflection nebulosity (bottom right). 4.2.2. Spectral Energy Distribution of PPNe One of the well-established characteristics of PPNe is that their SEDs have a "double-peaked" structure (cf. Kwok 1993). The shortward and longward peaks in the wavelength spectrum respectively correspond to the stellar and dust emission components. The morphological bifurcation between the SOLE and DUPLEX nebulae also manifest itself as a distinction between the SED shapes of these sources. A number of post-AGB stars have been classified into four classes based on the shape of the SED (VHG89): I. Flat spectrum between 4 and 25µm and a steep fall-off to shorter wavelengths, II. Maximum around 25µm and a gradual fall-off to shorter wavelengths, III. Maximum around 25µm and a steep fall-off to a plateau roughly between 1 and 4µm with a steep fall-off at shorter wavelengths, IV. Two distinct maxima; one around 25µm and a second between 1 and 2µm (IVa) or one around 25µm and a second < 1µm (IVb). Here, we adopt the VHG classification scheme for the SEDs of our sources. Because only five of our sources were studied in VHG89 with a partial coverage of the stellar component (i.e., > 1µm), we compiled new 4We included four other PPN candidates (one SOLE source, IRAS 21282+5050, and three DUPLEX sources, IRAS 04395+3601, Red Rectangle (IRAS 06176−1036), and Egg Nebula) whose mid-infrared and optical morphologies have been identified. -- 14 -- SEDs of our sources by adding our photometric measurements at optical wavelengths to the latest published data in the literature. Figures 6, 7, and 8 show updated SEDs for SOLE, DUPLEX, and stellar sources, respectively, with VHG class assignments indicated in each frame. The SED shapes of SOLE and DUPLEX nebulae are indeed very distinct from each other, but are very similar within each morphological type of the nebulosity. Stellar sources are of a mix of classes and are discussed in §4.2.5. Table 4 (column 5) summarizes a clear correlation between the morphological and SED classes. All SOLE nebulae have a double-peaked, class IV SED. The prominent central stars in SOLE nebulae appear in their SED as the well-defined, unobscured optical/near-infrared peak and the thermal emission from the circumstellar dust appears as an almost equal flux peak. There is a subdivision of the SED class among SOLE nebulae depending on the location of the optical/near-infrared peak. The difference stems from the degree of reddening of the central star due to its circumstellar dust/gas shell. Class IVb PPNe tend to have physically larger circumstellar shells and hence their column densities are lower, less dereddening the central star. Interestingly, mid-infrared images are resolved for those of class IVb (IRAS 07134+1005, IRAS 17436+5003, and IRAS 19114+0002), and this coincidence is consistent with the picture of more extended, optically thin dust shells of SOLE nebulae. IRAS 21282+5050 is a very young PN that we classify as a SOLE source based on morphology (Kwok, Hrivnak, & Langill 1993, Meixner et al. 1997), but its SED class appears to be III because the shorter wavelengths light from its hot central star (Of7, Cohen & Jones 1987) is more reddened by the dust than the star light from a typical PPN central star. DUPLEX SEDs are of class II (e.g., IRAS 19374+2359) or of class III (e.g., IRAS 17150−3224). Both classes II and III are characterized by a prominent far-infrared peak (30 to 50µm) with an optical/near- infrared excess that represents the central star or its associated reflection nebulosity. The difference between class II and III SEDs is the presence of a rather large near-infrared excess in class II, which is commonly attributed to either an ongoing mass loss episode or the presence of very compact circumstellar dust shell (VHG89). Interestingly, this SED class difference among DUPLEX sources corresponds to the visibility of the central star. That is, the SED will be of class II when the central star is partially visible whereas it will be of class III when the central star is completely obscured from the view. IRAS 09452+1330 is of class I but has a flat peak at <∼ 10µm caused by its lower Teff (∼ 2000K), which is consistent with its AGB stellar nature. Its very sharp drop into the shortward wavelengths resulted in non-detection in our B band image (Table 2). We classify IRAS 08005−2356 of class II because of its gradual fall-off in shorter wavelengths due to rather large optical/near-infrared flux. Our classification of IRAS 08005−2356 and IRAS 17423−1755 being DUPLEX sources is well supported by the resemblance in the shapes of the two SED classes. The differences in the SED shapes between SOLE and DUPLEX nebulae can be explained in the context of our hypothesis. In the case of a SOLE nebula, the circumstellar dust is optically thin and permits a clear, albeit reddened, view of the central star with a modest amount of dust emission, which is proportional to the column density of dust. Hence, we see two distinct, comparable peaks of stellar and dust emission in its SED. In the case of a DUPLEX nebula, on the other hand, the circumstellar dust is optically so thick that almost all of the star light is absorbed by the dust and is reradiated at mid- to far-infrared wavelengths; only a few optical photons escape through the biconical openings of the dust shell. Therefore, we see a prominent dust emission peak accompanied by an optical/near-infrared plateau in its SED. In the framework of our hypothesis, an inclination angle effect among DUPLEX sources manifests itself as the difference in their SED shapes. -- 15 -- 4.2.3. IRAS/Near-Infrared Two-Color Diagrams To demonstrate the differences between SOLE and DUPLEX sources, we use the J -- K vs. K -- [25] diagram, an IRAS/near-infrared two-color diagram (Fig. 9). Here, [25] is IRAS flux at 25µm in magnitude converted by [25] = −2.5 log(Fν /6.73) (IRAS Explanatory Suppliment 1988). Because K -- [25] color relates the heights of stellar and dust peaks whereas J -- K color describes the shape of the stellar component, the J -- K vs. K -- [25] diagram introduces characteristics of detached dust shells and incorporated the SED dichotomy into a diagram. The robustness of the J -- K vs. K -- [25] diagram is evident in the clear bifurcation between morphological groups. In J -- K color, SOLE sources are bluer (<∼ 1.45) than DUPLEX sources. This bluer color is consistent with the presence of an optically thinner circumstellar shell along the line of sight to the central star. In K -- [25] color, DUPLEX sources are spatially separated according to the visibility of the central star: DUPLEX sources with invisible central stars are redder than SOLE sources while those with partially visible central stars (IRAS 08005−2356 and IRAS 17423−1755) are bluer than SOLE sources. This bifurcation among DUPLEX sources follows the split of the SEDs into classes II and III: very high near-infrared excess in the class II DUPLEX sources makes their K -- [25] colors bluer than the class III DUPLEX sources and even bluer than SOLE sources. The diagram thus not only signifies the difference between SOLE and DUPLEX sources but also differentiates the partial/total obscuration of the central star in DUPLEX nebulae, and should be a very useful tool in identifying the nature of dust shell by near- to mid-infrared colors. With new all-sky near-infrared surveys becoming available (e.g. 2MASS and DENIS), near-infrared two-color diagrams will also be a valuable tool in discriminating a particular type of sources from a large data set. Whitelock (1985) presented near-infrared (JHK) photometry for 80 PNe and classified them into several types in terms of the visibility of the central star due to dust obscuration and of the location in the near-infrared two-color J -- H vs. H -- K diagram. Since the apparent morphological bifurcation is also partly based on the visibility of the central star, we adopt the classification scheme of Whitelock (1985) and make use of such two-color diagrams. Figure 10 is the J -- H vs. H -- K diagram with our sources that have published near-infrared photometric data. Also shown in the diagram are the regions according to the object classification for PN candidates: Nebula+Star, Nebula, Nebula+Dust, Star+Dust, and Miras (Whitelock 1985). Because of the dusty nature of PPNe, all of our sources are distributed over a linear diagonal region that corresponds to the regions of Miras and Star+Dust objects. There is a parallelism between the distributions of our targets and planetary nebulae in the diagram: SOLE nebulae correspond to PNe with prominently visible central nuclei (Nebula+Star), whereas DUPLEX nebulae correspond to dust enshrouded PNe (Nebula+Dust). There also seems to be a border that separates the region of SOLE sources from that of DUPLEX sources on the region of Miras (dashed line in Fig. 10). All SOLE nebulae are found on or blueward of the border and all DUPLEX nebulae are found on or redward of the border. The fact that DUPLEX sources are redder than SOLE sources suggests that DUPLEX nebulae are associated with a larger amount of obscuring dusts than SOLE nebulae are and this corroborates our hypothesis of the morphological bifurcation being induced by the differing optical thickness in the two types of sources. 4.2.4. The Inclination Angle Effect The inclination angle between the axis of symmetry with the line of sight can change the morphological appearance of an object. We have seriously considered if the inclination angle effect alone could explain all the observed differences between SOLE and DUPLEX nebulae with questions such as if SOLE nebulae are -- 16 -- nearly pole-on DUPLEX nebulae or not. In the following, we discuss that the evidence suggests otherwise. Suppose that the dust shell structure of all existing PPNe are of DUPLEX type, i.e., all dust shells have the same geometry of optically thick tori. When sources are oriented edge-on (90◦ inclination angle), most of the star light is blocked by the dust torus and we observe optical reflection nebulosities in the form of more or less well-balanced bipolar lobes with a dust lane (e.g. IRAS 17150−3224 and IRAS 17441−2411). However, in other cases when the inclination is in some intermediary angle, sources should appear as imbalanced nebulae either with a dust lane (with larger intermediate inclination angle) or without a dust lane (with smaller intermediate inclination angle), whose central stars are seen off center in the nebulosities when visible. The imbalance in the structure of nebulosities occurs because the far side of the nebulosity (which is pointing away from us) will be at least partially obscured by the near side of the optically thick dust torus, as we see in the images of IRAS 08005−2356 and IRAS 17423−1755, for example. This imbalance of brightness was also shown in simulated images of IRAS 17441−2411 with 4 different inclination angles presented by Su et al. (1998, their Fig. 5). According to their simulation, the central star is seen evidently off-centered even at 30◦ inclination angle. As we have seen in our images of SOLE nebulae, 11 of 21 nebulae appear as well-balanced, smooth, and symmetric low surface brightness nebulosities with their central stars located at centers of the nebulae. Following the discussion above, this is possible only when all 11 sources are oriented exactly pole-on or extremely close to pole-on. Thus, if this is the case one has to explain why nearly half the objects are oriented at zero or near-zero inclination angles with respect to us. For instance, IRAS 07134+1005 would be a prime example of a PPN viewed nearly exactly pole-on because it shows an extended, almost circular reflection nebulosity of uniform brightness with its central star at the center of the nebulosity. Nevertheless, the mid-infrared images of IRAS 07134+1005 clearly show a two-peaked, limb-brightened dust torus which suggests a non-zero inclination angle and radiative transfer calculations support a model of an equatorially enhanced dust torus viewed at an inclination angle of ∼ 45◦ (Meixner et al. 1997). If IRAS 07134+1005 were really a DUPLEX source viewed at a 45◦ inclination angle, it should have appeared as either an imbalanced nebula with the central star located off center or a bipolar nebula with one of the lobes partially obscured. However, that is not the case and thus the inclination angle effect can not simply explain the data. The most reasonable interpretation of the data is that the optical thickness along the line of sight is too low to cause any detectable difference, which is also supported by radiation transfer calculations (edge-on τ9.7µm ∼ 0.03, Meixner et al. 1997). As we have seen in the previous section, the distinction between SOLE and DUPLEX sources is obvious and the subdivision among DUPLEX sources is remarkable in the J -- K vs. K -- [25] diagram (Fig.10). One of the most peculiar aspects of the IRAS/near-infrared diagram is that the sources are not distributed linearly as in the J -- K vs. H -- K diagram. If the inclination angle effect were the main cause for the distinction between SOLE and DUPLEX sources, the sources would have been distributed linearly with the region of DUPLEX sources with obscured central stars being located between the regions of SOLE sources and DUPLEX sources without central stars. Alternatively, the absence of subdivision among SOLE sources corroborates our view of SOLE sources being associated with optically thin dust tori: no matter what the inclination angle is there is no partial obscuration in the SOLE sources and they cluster as a single group in the J -- K vs. K -- [25] diagram. It is of course possible that we come across a source whose orientation is exactly or very close to pole-on. In such cases, how the reflection nebulosities appear is not trivial. No matter to which morphological type a source belongs, it is incredibly difficult to detect reflection nebulosity when the source is viewed pole-on because the central star appears extremely prominent and the prominent PSF spikes are likely to -- 17 -- severely obscure the nebulosity. The appearance of IRAS 07430+1115, a SOLE PPN, does not fall into the morphological type of the DUPLEX nebulae, but the source still looks different from other SOLE nebulae. Although SED and two-color diagrams also suggest that this object is a SOLE source, it is possible that this is a DUPLEX source oriented at or very close to pole-on. However, we tentatively classify the source as a SOLE nebula and will not delve into the issue in this study. The pole-on cases must be further investigated with at least two-dimensional calculations. Radiative transfer calculations with fully axisymmetric models would clarify these inclination angle effects with more certainty and such calculations will be the focus of our future work. 4.2.5. Non-Association with Nebulosities There are six sources we tentatively classify as stellar sources because no reflection nebulosities are detected. Neither PSF subtraction nor deconvolution achieves the perfect removal of the WFPC2 diffraction features, and thus, there are always some residual diffraction artifacts which can mask a low surface brightness reflection nebula. In particular, our observations are insensitive to circular nebulosities with extremely low surface brightness, which can be easily confused with the WFPC2 circular diffraction wings. When viewed pole-on, SOLE sources would appear more or less like stellar sources due to a combined effect of the high star-to-nebula contrast and a confusion with the WFPC2 PSF artifact. For example, IRAS 04386+5722 and IRAS 05113+1347 may be such cases. The images of these sources barely show nebulosities and the deconvolved images are almost nebula-free for IRAS 04386+5722 whereas suggestive of a compact nebula for IRAS 05113+1347. However, their SEDs are of class IVa (Fig. 8) and their locations in two-color diagrams are within the region of SOLE sources (an asterisk and a filled star in Figures 9 and 10). The slight offset of IRAS 04386+5722 towards blue in K -- [25] color in the J -- K vs. K -- [25] diagram can be explained considering the pole-on inclination angle effect: the source is of SOLE type and hence the stellar emission peak is present in the SED irrespective of the angle of inclination and hence its J -- K color would not be very different from other SOLE sources, while its K -- [25] color can be bluer than other SOLE sources because the stellar emission would be more prominent than other SOLE sources with respect to the dust emission. IRAS 05113+1347 is located among other SOLE sources. IRAS 10158−2844 and IRAS 15465+2818 have very bright central stars, but the amount of far-infrared excess is smaller (Fig. 8). The lack of far-infrared excess in these sources makes the SED classification scheme of VHG89 inapplicable and also makes both have much bluer K -- [25] color than other sources (K -- [25] = 3.48 and 3.12, respectively). This is consistent with the evolutionary status of these R Coronae Borealis stars, in which they have had little recent mass loss but are associated with very diffuse, extended circumstellar dust shells (Gillett at al. 1986, Waters et al. 1989). However, if we consider the stellar component peaks alone (which peak at < 1µm) these objects will be found among DUPLEX sources (Fig. 10; filled stars in the region of DUPLEX sources) suggesting the presence of circumstellar dust rather close to the central star, on-going mass loss, or pole-on inclination angle effects. The nature of variability of IRAS 10158−2844 is suspected to be due to variable obscuration by circumstellar material along the line of sight and the inclination angle of this source is expected to be >∼ 50◦ (Waelkens et al. 1991). In fact, both of these sources show indications of a very recent mass loss (Clayton et al. 1997, Meixner et al. 1999). Hence, a possible explanation for the non-detection of any nebulosities is that the column density of dust near the star in these sources is much lower than in SOLE sources. IRAS 20043+2653 is classified as an OH/IR source (cf. Garc´ıa-Lario et al. 1997a), and hence, the -- 18 -- central star is likely be in the AGB or extreme AGB phase. This interpretation fits well with the SED shape and its very red H -- K color (both of which resemble to those of IRAS 09452+1330, an extreme AGB star). Therefore, the absence of any reflection nebulosity seems to suggest that IRAS 20043+2653 has not yet developed one. IRAS 22142+5206 is classified as a young stellar object embedded in a massive CO molecular cloud of ∼ 7300M⊙ (Dobashi et al. 1998). Its SED shows significant far-infrared excess without a stellar component (Fig. 8), which is a signature of a class I young stellar object (e.g., Wilking, Lada, & Young 1989). Its location in two-color diagrams is also consistent with the young stellar object interpretation. 4.3. Origins of the PPN Morphological Bifurcation The origin of axisymmetry in many astrophysical systems is always of a great importance and there have been numerous possible mechanisms for the creation of nebular morphologies (cf. Livio 1997). Instead of reviewing every possible means, we will focus on the origins of the differing equatorial density enhancement in the SOLE and DUPLEX nebulae. 4.3.1. Galactic Height and Progenitor AGB Stellar Mass One of the strongest pieces of circumstantial evidence that is related to the morphological bifurcation is probably that bipolar nebulae are preferentially found close to the plane of the Galaxy. With a large sample of PNe, Corradi & Schwartz (1995) found that bipolar PNe were distributed closer to the Galactic plane (scale height zh = 130 pc with z <∼ 850 pc) than elliptical PNe (zh = 325 pc with z <∼ 1300 pc). Following this finding, they suggested that bipolar PNe have evolved from more massive progenitors than elliptical PNe, adopting a lower limit of 1.5M⊙ for the bipolar PN progenitors. This correlation is also suggested by the galactic latitudes and the z values of our sources (Table 4). DUPLEX sources, having a mean height of 220 pc with a range of z <∼ 520, are more confined to the Galactic plane than the SOLE sources, which have a mean height of 470 pc with a range of z <∼ 2100. Although a direct comparison between our values (mean Galactic heights) to the values obtained by Corradi & Schwartz (1995; Galactic scale heights) is not possible, there certainly exists a parallelism in the ways these two types of PPNe and two types of PNe are distributed in the Galaxy. To test if the Galactic height distributions of the SOLE and DUPLEX sources are not exactly equal, we calculated the Kolmogorov -- Smirnov (K -- S) statistic (0.417) and its significance level (0.186). This means that there is 18.6% chance that the K -- S statistic, the greatest difference between the two cumulative distribution functions of the Galactic heights, will be smaller than 0.417, if both types of objects are from the same Galactic height distribution. Considering the fact that some SOLE sources do exist close to the Galactic plane where DUPLEX sources are populated, the outcome of the K -- S test is suggestive that SOLE and DUPLEX sources are not distributed in the exactly same manner. Therefore, it is likely that more massive progenitor AGB stars lead to DUPLEX PPNe and that bipolar PNe are the direct descendants of DUPLEX PPNe and elliptical PNe are the SOLE PPN offspring. In fact, this is a very reasonable result in the context of stellar evolution. The majority of AGB stars become white dwarfs of roughly 0.6 M⊙ (Schonberner 1981); the more massive the progenitor, the more material the star has to dump into the circumstellar environment. Therefore, PPNe from more massive progenitor AGB stars are likely to have more obscuring material in the circumstellar shell, which could completely block the central star from the observer's view as in DUPLEX nebulae. This is well demonstrated in Figure 10 as DUPLEX sources being redder than SOLE sources. Also consistent with the division of PPN into two classes is that -- 19 -- the degree of equatorial enhancement and the subsequent evolution depend on the fundamental parameters of the central star, especially, the stellar mass. 4.3.2. Circumstellar Chemistry Because PPNe can be put into two groups in terms of photospheric and/or circumstellar chemistry (C- or O-rich), we looked for a correlation between the circumstellar chemistry and morphological bifurcation. When referring to the chemical type of a PPN, one needs to be cautious because a PPN can have both C- and O-rich characteristics in the circumstellar environment above the photosphere (e.g., IRAS 08005−2356 and Red Rectangle). The circumstellar chemistry can be determined mainly by the presence of a certain molecular species (e.g., OH in an O-rich shell, Hu et al. 1994; HCN in a C-rich shell, Loup et al. 1993) or some infrared spectral feature (e.g., 9.7µm feature in an O-rich shell; 21µm feature in a C-rich shell, Kwok, Volk, & Hrivnak 1989). On the other hand, the photospheric chemistry can be determined by direct abundance measurements (e.g., IRAS 02229+6208, IRAS 07430+1115, Reddy, Bakker & Hrivnak 1999; IRAS 04296+3429, Decin et al. 1998; IRAS 05341+0852, Reddy et al. 1997; IRAS 06530−0213, Reddy 1999; IRAS 18095+2704, Klochkova 1995; IRAS 19114+0002, Reddy & Hrivnak 1999; IRAS 20462+3416, Garc´ıa-Lario et al. 1997a) or the presence of optical photospheric features of C2 or C3 molecules (e.g., Hrivnak 1995). Comparison between the PPN morphology and photospheric/circumstellar chemistry does not seem to yield any apparent correlation between the two. Table 4 (column 7) summarizes the non-correlation between morphology and chemistry. 4.3.3. Stellar Ages It may be possible to attribute the differing optical thickness in SOLE and DUPLEX PPNe to their ages. PPN dust shells expand with time and older shells tend to have smaller optical depth, and therefore, DUPLEX PPNe are expected to be younger than SOLE PPNe. If one insists on a single evolutionary channel, DUPLEX PPNe may be forerunners of SOLE PPNe. However, bifurcating PPNe into SOLE and DUPLEX types by their ages does not appear to be possible because of the spectral types of these stars. If DUPLEX PPNe were indeed younger than SOLE PPNe, DUPLEX PPNe would have had the latest spectral types possible (G to M types) and SOLE PPNe would have had the earliest spectral types (A to F types). However, SOLE PPNe include a number G types (e.g. IRAS 04296+3429) and DUPLEX PPNe include F types (e.g. IRAS 08005−2356 and Egg Nebula). Given the horizontal evolutionary track in the HR diagram during the PPN phase and its surpassingly short evolutionary time scale, it is not possible to conclude one type of PPNe is younger than the other. 4.4. From the Dual PPN Morphology to the PN Morphology During the PPN phase, the two-layered PPN shell keeps expanding around the central post-AGB star while the surface temperature of the star continues to rise. A fast wind initiates somewhere along the PPN phase and pushes the inner boundary of the PPN shell out to typical PN dimensions, while shaping the boundary geometry and increasing the boundary density (Kwok 1982). The central post-AGB star finally becomes hot enough to emit photoionizing photons, which illuminate the inner boundary of the circumstellar gas shell as a PN. Because we observe dust-scattered light in PPNe and ionized gas emission -- 20 -- in PNe, we can not trivially link the PPN and PN morphologies via a mere resemblance in the images. We can, nevertheless, interpolate the PPN and PN morphologies considering the circumstellar distribution of matter and see the PN morphological structures (round, elliptical, and butterfly classes; Balick 1987) in the dual PPN morphology. The bipolar shapes of DUPLEX PPNe (e.g., IRAS 08005−2356, IRAS 17150−3224) may be forerunners to the bipolar PNe (e.g., Hourglass Nebula, Sahai & Trauger 1995). The simply elongated SOLE nebulae (e.g., IRAS 17436+5003, IRAS 20462+3416) may be precursors of elliptical PNe (e.g., NGC 3132 (HST Heritage Team 1998), IC 3568 (Bond & Ciardullo 1997)). The multi-lobed SOLE nebulae (e.g., IRAS 22272+5435, IRAS 06530−0213) may be progenitors of complex PNe (e.g., Stingray Nebula (Bobrowsky et al. 1998), Cat's Eye Nebula (Harrington & Borkowski 1995)). We thus see that the development of the PPN axisymmetry in the superwind phase probably sets the stage for the emergence of axisymmetry in PNe. Jet-like structures, however, are rather rare in PPNe (e.g., IRAS 17423−1755, which is a young PN) and this may suggest that the formation of jet-like structures seen in PNe (e.g., NGC 5307, Bond & Ciardullo 1997) does not share the same generating mechanism as the PPN axisymmetric structures. 5. Conclusions After observing 27 PPN candidates with HST , we have found elongated low surface brightness reflection nebulosities around 21 sources. We have also found that an optical reflection nebulosity can manifest itself in the form of a faint, elliptically elongated shell in addition to the bipolar form. The PPN circumstellar shell seems to be intrinsically axisymmetric (ellipticity ∼ 0.44) and we argue that the axisymmetry emerges in the superwind phase, the latter of the two-phased AGB mass loss epoch. A morphological bifurcation exists among the PPN nebulosities: of 21 extended nebulae, 11 are SOLE nebulae (e.g., IRAS 07134+1005) and 10 are DUPLEX nebulae (e.g., IRAS 17150−3224). We discuss how the morphological dichotomy is caused by the difference in optical thickness of the PPN circumstellar dust shells: SOLE shells are optically thin whereas a DUPLEX shells are optically thick. The distinctness between SOLE and DUPLEX nebulae in terms of optical thickness of the dust shells is evidenced by the correlation between the mid-infrared morphology of dust emission regions and optical morphology of reflection nebulosities, the characteristic shapes of the SEDs, and the near- and IRAS/near-infrared two-color diagrams. We also discuss that the inclination alone may not be able to explain the well-balanced shape of reflection nebulosities with their central stars seen at the center. Although we find no correlation between the circumstellar chemistry and morphology, we do find that DUPLEX sources tend to be found closer to the Galactic plane than SOLE sources. This suggests that DUPLEX PPNe probably originate from higher mass AGB progenitor stars than SOLE PPNe. The origins of the apparent morphological bifurcation -- the equatorial density enhancement in the superwind -- remain inconclusive. In addition to optical imaging of reflection nebulosities, direct, high-resolution imaging of dust emission regions in mid-infrared wavelengths will be extremely important in studies of PPNe because optical images of reflection nebulosities alone are not sufficient to decipher the orientation of these objects. Similarly, future investigation will have to require at least two-dimensional, radiative transfer model calculations. This research is based on observations with the NASA/ESA Hubble Space Telescope, obtained at the Space Telescope Science Institute (STScI), which is operated by the Association of Universities for Research in Astronomy, Inc. under NASA contract No. NAS5-26555. Data reduction was done with routines provided in the Space Telescope Science Data Analysis System (STSDAS), which is a software package for -- 21 -- calibrating and analyzing data from HST . STSDAS includes the same calibration routines as are used in the routine data processing pipeline, as well as general-purpose tools and enhancements to the Image Analysis and Reduction Facility (IRAF) distributed by the National Optical Astronomy Observatories, which are operated by the Association of Universities for Research in Astronomy, Inc., under cooperative agreement with the National Science Foundation. This research also made use of the SIMBAD database, operated at CDS, Strasburg, France. Ueta and Meixner are supported by NASA Grants GO-06737.01-95A and GO-06364.02-95A and NSF Career Award Grant AST 97-33697. Bobrowsky is supported by STScI Grants GO-06364.01-94A and GO-006737.02-95A. We are thankful for the help that Chris Skinner gave us in setting up these observations. Chris J. Skinner, who died suddenly in October 1997, was a valued colleague and friend. We would also like to thank the STScI support staff for their guidance in setting up these observations and helping with the reduction of the data. Bakker, E. J., van Dishoeck, E. F., Waters, L. B. F. M., & Schoenmaker, T. 1997, A&A, 323, 469. REFERENCES Balick, B. 1987, AJ, 94, 671. Blocker, T. 1995, A&A, 297, 727. Blommaert, J. A. D. L., van der Veen, W. E. C. J., & Habing, H. J. 1993, A&A, 267, 39. Bobrowsky, M., Sahu, K. C., Parthasarathy, M., & Garc´ıa-Lario, P. 1998, Nature, 392, 469. Bobrowsky, M., Zijlstra, A. A., Grebel, E. K., Tinny, C. G., te Lintel Hekkert, P., Van de Steene, G. C., Likkel, L., & Bedding, T. R. 1995, ApJ, 446, L89. Bond, H. & Ciardullo, R. 1997, STScI-PRC, 97-38b. Borkowski, K. J., Blondin, J. M., & Harrington, J. P. 1997, ApJ, 482, L97. Bujarrabal, V., Alcolea, J., & Planesas, P. 1992, A&A,, 257, 701. Campbell, M. P., Elias, J. H., Gezari, D. Y., Harvey, P. M., Hoffmann, W. F., Hudson, H. S., Neugebauer, G., Soifer, B. T., Werner, M. W., & Westbrook, W. E. 1976, ApJ, 208, 396. Clayton, G. C., Bjorkman, K. S., Nordsieck, K. H., Zellner, N. E. B., & Schulte-Ladbeck, R. E. 1997, ApJ, 476, 870. Cohen, M., Anderson, C. M., Cowley, A., Coyne, G. V., Fawley, W., Gull, T. R., Harlan, E. A., Herbig, G. H., Holden, F., Hudson, H. S., Jakoubek, R. O., Johnson, H. M., Merrill, K. M., Schiffer, F. H., Soifer, B. T., & Zuckerman, B. 1975, ApJ, 196, 179. Cohen, M. & Jones, B.F. 1987, ApJ, 321, L151. Corradi, R. L. M. & Schwartz, H. E. 1995, A&A, 293, 871. Decin, L., Van Winckel, H., Waelkens, C., & Bakker, E. J. 1998, A&A, 332, 928. Dobashi, K., Yonekura, Y., Hayashi, Y., Sato, F., & Ogawa, H. 1998, ApJ, 115, 777. Feast, M. W., Carter, B. S., Roberts, G., Marang, F., & Catchpole, R. M. 1997, MNRAS, 285, 317. -- 22 -- Fruchter, A. S., Hook, R. N., Busko, I. C., & Mutchler, M. 1997, in 1997 HST Calibration Workshop, eds. S. Casertano, R. Jedrzejewski, C. D. Keyes, & M. Stevens (Baltimore: STScI), p.518. Garc´ıa-Lario, P., Manchado, A., Pottasch, S. R., Suso, J., & Olling, R. 1990, A&AS, 82, 497. Garc´ıa-Lario, P., Manchado, A., Pych, W., & Pottasch, S. R. 1997a, A&AS, 126, 479. Garc´ıa-Lario, P., Parthasarathy, M., de Martino, D., Sanz Fern´andez de C´ordoba, L., Monier, R., Manchado, A., & Pottasch, S. R. 1997b, A&A, 326, 1103. Geballe, T. R., Tielens, A. G. G. M., Kwok, S., & Hrivnak, B. J. 1992, ApJ, 387, L89. Geballe, T. R. & van der Veen, W. E. C. J. 1990, A&A, 235, L9. Gillett, F. C., Backman, D. E., Beichman, C., & Neugebauer, G. 1986, ApJ, 310, 842. Goldreich, P. & Scoville, N. 1976, ApJ, 205, 144. Gutler, J., Kompe, C., & Henning, Th. 1996, A&A, 305, 878. Habing, H., te Lintel Hekkert, P., & van der Veen, W. E. C. J. 1989, in IAU Symposium 131: Planetary Nebulae, ed. S. Torres-Peimbert (Dordrecht: Kluwer), p.359 Habing, H. & Blommaert, J. A. D. L. 1993, in IAU Symposium 155: Planetary Nebulae, eds. R. Weinberger & A. Acker (Dordrecht: Kluwer), p.243 Harrington, J. P. & Borkowski, K. J. 1995, 185th AAS meeting (Tucson, AZ). Hawkins, G. W., Skinner, C. J., Meixner, M., Jernigan, J. G., Arens, J. F., Keto, E., & Graham, J. R. 1995, ApJ, 452, 314. Holtzman, J. A., Burrows, C. J., Casertano, S, Hester, J. J., Trauger, J. T., Watson, A. M., Worthey, G. 1995, PASP, 107, 1065. Hrivnak, B. J. 1995, ApJ, 438, 341. Hrivnak, B. J. & Kwok, S. 1991a, ApJ, 368, 564. Hrivnak, B. J. & Kwok, S. 1991b, ApJ, 371, 631. Hrivnak, B. J. & Kwok, S. 1999a, ApJ, 513, 869. Hrivnak, B. J., Kwok, S., & Geballe, T. R. 1994, ApJ, 420, 783. Hrivnak, B. J., Kwok, S., & Volk, K. 1988, ApJ, 331, 832. Hrivnak, B. J., Kwok, S., & Volk, K. 1989, ApJ, 346, 265. Hrivnak, B. J., Langill, P. P., Su, K. L., & Kwok, S. 1999b, ApJ, 513, 421. HST Heritage Team 1998, STScI-PRC, 98-39. Hu, J. Y., Slijkhuis, S., de Jong, T., & Jiang, B. W. 1993a, A&AS, 100, 413. Hu, J. Y., Slijkhuis, S., Nguyen-Q-Rieu, & de Jong, T. 1993b, A&A, 273, 185. Hu, J. Y., te Lintel Hekkert, P., Slijkhuis, S., Baas, F., Sahai, R., & Wood, P. 1994, A&AS, 103, 301. -- 23 -- Humphreys, R. M. & Ney, E. P. 1974, ApJ, 190, 339. Iben, I. & Renzini, A. 1983, ARA&A, 21, 271. Iyengar, K. V. K. & Parthasarathy, M. 1997, A&AS, 121, 45. Johnson, J. J. & Jones, T. J. 1991, AJ, 101, 1735. Joint IRAS Science Working Group 1988, Infrared Astronomical Satellite Catalogs and Atlases, Version 2. Explanatory Suppliment, eds. C. A. Beichmen, G. Neugebauer, H. J. Habing, P. E. Clegg, & T. J. Chester. (Washington: NASA Reference Publication). Kahn, F. D. & West, K. A. 1985, MNRAS, 212, 837. Kastner, J. H. & Weintraub, D. A. 1995, ApJ, 452, 833. Klochkova, V. G. 1995, MNRAS, 272, 710. Krist, J. E., Burrows, C. J., Stapelfeldt, K. R., Ballester, G. E., Clarke, J. T., Crisp, D., Evans, R. W., Gallagher, J. S. III, Griffiths, R. E., Hester, J. J., Holtzman, J. A., Hoessel, J. G., Mould, J. R., Scowen, P. A., Trauger, J. T., Watson, A. M., & Westphal, J. A. 1997, ApJ, 481, 447. Krist, J. E., & Hook, R. N. 1997, The Tiny Tim User's Guide v4.4 (Baltimore: STScI). Kwok, S. 1975, ApJ, 198, 583. Kwok, S. 1982, ApJ, 258, 280. Kwok, S. 1993, ARA&A, 31, 63. Kwok, S., Hrivnak, B. J., & Boreiko, R. T. 1987, ApJ, 312, 303. Kwok, S., Hrivnak, B. J., & Geballe, T. R. 1990, ApJ, 360, L23. Kwok, S., Hrivnak, B. J., & Geballe, T. R. 1995, ApJ, 454, 394. Kwok, S., Hrivnak, B. J., & Langill, P. P. 1993, ApJ, 408, 586. Kwok, S., Hrivnak, B. J., Zhang, C. Y., & Langill, P. L. 1996, ApJ, 472, 287. Kwok, S., Su, K. Y. L., & Hrivnak, B. J. 1998, ApJ, 501, L117. Kwok, S., Volk, K., & Hrivnak, B. J. 1989, ApJ, 345, L51. Lawrence, G., Jones, T. J., & Gehrz, R. D. 1990, AJ, 99, 1232. Le Bertre, T. 1988, 203, 85. Likkel, L. & Morris, M. 1988, ApJ, 329, 914. Likkel, L., Morris, M., Omont, A., & Forveille, T. 1987, A&A, 173, L11. Livio, M. 1997, Sp. Sci. Reviews, 82, 389. Loup, C., Forveille, T., Omont, A. & Paul, J. F. 1993, A&AS, 99, 291. -- 24 -- Luck, R. E., Bond, H. E., & Lambert, D. L. 1990, ApJ, 357, 188. Manchado, A. Pottasch, S. R., Garc´ıa-Lario, P., Esteban, C., & Mampaso, A. 1989, A&A, 214, 139. Mastrodemos, N. & Morris, M. 1999, ApJ, in press. Meixner, M., Skinner, C. J., Graham, J. R., Keto, E., Jernigan, J. G. & Arens, J. F. 1997, ApJ, 482, 897. Meixner, M., Ueta, T., Dayal, A., Hora, J. L., Fazio, G., Hrivnak, B. J., Skinner, C. J., Hoffmann, W. F., & Deutsch, L. K. 1999, ApJS, 122, 221. Mellema, G & Frank, A. 1995, MNRAS, 273, 401. Netzer, N. & Elitzur, M. 1993, ApJ, 410, 701. Oudmaijer, R. D., van der Veen, W. E. C. J., Waters, L. B. F. M., Trams, N. R., Waelkens, C., & Engelsman, E. 1992, A&AS, 96, 625. Paczy´nski, B. 1975, ApJ, 202, 558. Parthasarathy, M. 1993, ApJ, 414, L109. Reddy, B. E. 1999, private communication Reddy, B. E. & Hrivnak, B. J. 1999a, AJ, 117, 1834. Reddy, B. E., Bakker, E. J. & Hrivnak, B. J. 1999b, submitted to ApJ. Reddy, B. E., Parathasarathy, M., Gonzalez, G., & Bakker, E. J. 1997, A&A, 328, 331. Reddy, B. E. & Parathasarathy, M. 1996, AJ, 1112, 2053. Renzini, A. 1981, in Physical Processes in Red Giants, ed. I. Iben & A. Renzini (Dordrecht: Reidel), p.431. Riera, A., Garc´ıa-Lario, P., Manchado, A., Pottasch, S. R., & Raga, A. C. 1995, A&A, 302, 137. Sahai, R., te Lintel Hekkert, P., Morris, M., Zijstra, A., & Likkel, L. 1999, ApJ, 514, L115. Sahai, R. & Trauger, J. T. 1996, STScI-PRC, 96-07. Sahai, R., Trauger, J. T., Watson, A. M., Stapelfeldt, K. R., Hester, J. J., Burrows, C. J., Ballister, G. E., Clarke, J. T., Crisp, D., Evans, R. W., Gallagher, J. S., III, Griffiths, R. E., Hoessel, J. G., Holtzman, J. A., Mould, J. R., Scowen, P. A., & Westphal, J. A. 1998, ApJ, 493, 301. Salpeter, E. E. 1974, ApJ, 193, 585. Schmidt, G. D. & Cohen, M. 1981, ApJ, 246, 444. Schonberner, D. 1981, A&A, 103, 119. Shepp, L. A. & Vardi, Y. 1982, IEEE Trans. Med. Imaging, MI-1, 113. Simon, B. 1997, Synphot User's Guide v4.0 (Baltimore: STScI). Skinner, C. J., Meixner, M., Bobrowsky, M. 1998, MNRAS, 300, L29. -- 25 -- Skinner, C. J., Meixner, M., Barlow, M. J., Collison, A. J., Justtanont, K., Blanco, P., Pina, R., Ball, J. R., Keto, E., Arens, J. F., & Jernigan, J. G. 1997, A&A, 328, 290. Skinner, C. J., Meixner, M., Hawkins, G. W., Keto, E., Jernigan, J. G., & Arens, J. F. 1994, ApJ, 423, L135. Slijkhuis, S., Hu, J. Y., & de Jong, T. 1991, A&A, 248, 547. Smith, V. V. & Lambert, D. L. 1994, ApJ, 424, L123. Soker, N. 1998, MNRAS, 299, 1242. Su, K. Y. L., Volk, K., Kwok, S., & Hrivnak, B. J. 1998, ApJ, 508, 744. te Lintel Hekkert, P. 1991, A&A, 248, 209. Trammell, S. R., Dinerstein, H. L., & Goodrich R. W. 1994, AJ, 108, 984 (TDG94). Van der Veen, W. E. C. J., Habing, H. J., & Geballe, T. R. 1989, A&A, 226, 108 (VHG89). Waelkens, C., Lamers, H. J. G. L. M., Waters, L. B. F. M., Rufener, F., Trams, N. R., Le Bertre, T., Ferlet, R., Vidal-Madjae, A. 1991, A&A, 242, 433. Waters, L. B. F. M., Lamers, H. J. G. L. M., Snow, T. P., Mathlener, E., Trams, N. R., van Hoof, P. A. M., Waelkens, C., Seab, C. G., & Stanga, R. 1989, A&A, 221, 208. Waters, L. B. F. M., Waelkens, C., van Winckel, H., Molster, F. J., Tielens, A. G. G. M., van Loon, J. Th., Morris, P. W., Cami, J., Bouwman, J., de Koter, A., de Jong, T., & de Graauw, Th. 1998, Nature, 391, 868. Weintraub, D. A., Huard, T., Kastner, J. H., & Gatley, I. 1998, ApJ, 509, 728. Westbrook, W.E., Becklin, E.E., Merrill, K.M., Neugebauer, G., Schmidt, M., Willner, S.P., & Wynn-Williams, C.G. 1975, ApJ, 202, 407. Whitelock, P. A. 1985, MNRAS, 213, 59. Whitmore, B. C. 1997, in 1997 HST Calibration Workshop, eds. S. Casertano et al (Baltimore: STScI), p.317. Wilking, B. J., Lada, C. J., & Young, E. T. 1989, ApJ, 340, 823. Zacs, L., Klochkova, V. G., & Panchuk, V. E. 1995, MNRAS, 275, 764. Zuckerman, B. & Aller, L. H. 1986, ApJ, 301, 772. This preprint was prepared with the AAS LATEX macros v4.0. -- 26 -- Fig. 1. -- (a) -- (c): Images of SOLE nebulae in the increasing order of their right ascension (north is up and east is left): the left-most frame shows the IRAS ID and scale of the object. The tick marks show relative offsets in arcseconds. The filter types are shown at the bottom of each frame with "+ RL" indicating Richardson-Lucy deconvolution. Wedges show the ranges of log-scaled flux density to help readers visually illustrate the emission contrast. See Table 1 for the star-to-nebula surface intensity ratio. Fig. 2. -- (a) -- (c): Images of DUPLEX nebulae. The displaying scheme follows Figure 1. See Table 2 for the star-to-nebula surface intensity ratio. Fig. 3. -- Selected Images of stellar sources. These images exemplify how PSF artifacts appear in both reduced and deconvolved images. The basic displaying scheme follows Figure 1. Fig. 4. -- A schematic diagram illustrating the essential difference between SOLE and DUPLEX sources. [Top] In the SOLE sources, the superwind mass loss creates marginally equatorially enhanced dust shell (gray area near the central star) within a spherically symmetric the AGB wind shell. Because the axisymmetric dust shell is optically thin, it permits star light to leak out in all directions (arrows), creating an elliptically elongated nebula (dashed line) with the bright central star. [Bottom] In the DUPLEX sources, the superwind mass creates highly equatorially enhanced dust shell (black area near the central star) within a spherically symmetric the AGB wind shell. Because the axisymmetric dust shell is optically thick, it permits starlight to leak out only along the biconical openings of the dust torus (arrows), creating a bipolar, dumbbell- like nebulosity (dashed perimeter) with the partially or completely obscured central star. Relative sizes of components are not to scale. Fig. 5. -- Composites of optical and mid-infrared images of four PPN candidates. Mid-infrared contours are overlaid onto an optical grayscale image. The top row shows examples of the SOLE-toroidal correlation (IRAS 07134+1005: F410M and 11.8µm and IRAS 17436+5003: F410M and 12.5µm) and the bottom row shows examples of the DUPLEX-core/elliptical correlation (IRAS 17150−3224: F814W and 12.5µm and IRAS 16342−3814: F814W and 9.8µm). Tickmarks show relative offsets from the center. Fig. 6. -- Spectral energy distributions of SOLE sources in λFλ (ergs s−1 cm−2) vs. λ (µm). The name of the source and its SED class assignment (Van der Veen, Habing & Geballe 1989; see text) are given in each frame. Photometric data are taken from our measurements and references listed in Table 4. Fig. 7. -- Spectral energy distributions of DUPLEX sources in λFλ (ergs s−1 cm−2) vs. λ (µm). Conventions follow those of Fig. 6. Fig. 8. -- Spectral energy distributions of stellar sources in λFλ (ergs s−1 cm−2) vs. λ (µm). Conventions follow those of Fig. 6. Fig. 9. -- IRAS/Near-infrared two-color (J -- K vs. K -- [25]) diagram for our sources. Circles, triangles, and stars are respectively SOLE, DUPLEX, and stellar sources. The asterisk is IRAS 04386+5722, a stellar source whose chemistry is undetermined. The numbers associated with symbols refer to the source numbers given in Table 4 (column 1). Two stellar sources, IRAS 10158−2844 and IRAS 15465+2818, and one DUPLEX source, IRAS 09452+1330, are located off the diagram: (K -- [25], J -- K) = (3.48, 1.59), (3.12, 2.11), and (10.21, 6.54), respectively. This diagram displays the color of the star vs. the ratio of the dust peak to the stellar peak in the SED. Sources are more clearly clustered into three groups: SOLE and DUPLEX sources with a totally obscured central star (w/o Star), and DUPLEX sources with a visible central star (w/ Star). Each subdivision is labeled in the diagram. The dashed line indicates a fiducial division between SOLE and DUPLEX sources. All sources above the line are of DUPLEX type and below the line are of SOLE type. -- 27 -- Fig. 10. -- Near-infrared two-color (J -- H vs. H -- K) diagram for our sources. Conventions follow those of Fig. 9. Two AGB stars, IRAS 09452+1330 (C-rich) and IRAS 20043+2653 (O-rich) are located off the diagram: (H -- K, J -- H) = (3.03, 3.51) and (2.79, 0.30), respectively. Whitelock's classification of planetary nebula (1985) is also shown in the diagram: the regions of Nebula+Star, Nebula+Dust, Star+Dust, Nebula, and Miras. Sources are clustered diagonally in the diagram, which indicates the black-body temperature decreasing to the upper right. The dashed line indicates a fiducial division between SOLE and DUPLEX sources. All sources on the right of the line are of DUPLEX type and on the left of the line are of SOLE type. 9 9 9 1 IRAS ID 02229+6208 04296+3429 05341+0852 06530−0213 07134+1005 (HD56126) 07430+1115 p e S 6 3 v 8 3 2 8 0 9 9 / h p - o r t s a : v i X r a 17436+5003 (HD161796) 18095+2704 (OH53.8+20.2) 19114+0002 (HD179821) 20462+3416 22272+5435 (HD235858) Observed and Derived Properties of SOLE Nebulae TABLE 1 Obs. Coord. (J2000) WFPC2 1 Fλ RA DEC Filter HST Mag2 Surface Intensity3 Peak3a Ratio3b Size (′′ × ′′ (σ)) Ellipticity4 (e = 1 − b/a) 02:26:41.9 +62:21:22 04:32:57.0 +34:36:13 05:36:55.0 +08:54:08 06:55:31.8 −02:17:29 07:16:10.3 +09:59:48 07:45:51.4 +11:08:20 17:44:55.4 +50:02:40 18:11:30.8 +27:05:15 19:13:58.6 +00:07:32 20:48:16.6 +34:27:25 22:29:10.4 +54:51:07 F555W 6.36 e−14 F814W 3.51 e−13 F555W 7.68 e−15 F814W 2.54 e−14 F555W 1.43 e−14 F814W 3.16 e−14 F555W 8.83 e−15 F814W 3.14 e−14 F410M 1.03 e−12 F547M 1.71 e−12 F555W 3.90 e−14 F814W 7.45 e−14 F410M 8.07 e−12 F547M 5.82 e−12 F555W 2.74 e−13 F814W 3.70 e−13 F410M 7.11 e−13 F547M 1.70 e−12 F555W 1.45 e−13 F814W 6.05 e−14 F555W 1.28 e−12 F814W 2.26 e−12 11.89 10.04 14.19 12.89 13.51 12.65 14.03 12.66 8.87 8.32 12.42 11.72 6.63 6.99 10.30 9.98 9.27 8.32 10.99 11.95 8.63 8.01 9.2 e−2 5.1 e−1 2.0 e−3 1.4 e−2 1.9 e−2 4.8 e−2 8.3 e−3 3.4 e−2 1.5 e+0 4.0 e+0 1.9 e−2 5.4 e−2 1.4 e+1 1.0 e+1 2.9 e−1 4.8 e−1 2.0 e+0 3.0 e+0 4.4 e−1 1.4 e−1 1.7 e+0 2.6 e+0 1000 1200 65 220 220 440 390 860 5300 11000 18 50 9400 7100 960 1100 180000 110000 17000 16000 3500 3700 2.09 × 1.42 (2.2) 2.14 × 1.52 (2.5) 1.88 × 1.57 (2.9) 1.69 × 1.59 (3.0) 1.12 × 0.81 (5) 1.14 × 0.78 (5) 2.36 × 1.11 (2.5) 2.33 × 0.92 (2.2) 4.73 × 4.15 (5.3) 4.70 × 3.71 (4.0) 1.01 × 0.90 (6) 0.98 × 0.86 (7) 4.34 × 2.46 (4.9) 4.49 × 2.53 (3.6) 1.89 × 1.27 (2.2) 1.82 × 1.04 (6) 10.69 × 8.30 (∼1) 10.82 × 8.52 (2) 4.10 × 3.23 (2.2) 3.99 × 2.94 (2.5) 3.53 × 3.47 (5.6) 3.47 × 3.37 (4.0) 0.32 0.29 0.61, 0.67 0.61, 0.66 0.72, 0.33 0.70, 0.31 0.53 0.61 0.12 0.21 0.11 0.12 0.43 0.44 0.80, 0.54 0.69, 0.60 0.22 0.21 0.21 0.26 0.43, 0.61 0.41, 0.62 1In units of ergs/s/cm2/A. 2HST WFPC2 system magnitudes (STMAG). 3In units of ergs/s/cm2/A/sr: 3a-peak intensity of the central source, 3b-star-to-nebula intensity ratio. 4a and b are respectively major and minor axis lengths. Observed and Derived Properties of DUPLEX Sources TABLE 2 IRAS ID Obs. Coord. (J2000) WFPC2 1 Fλ RA DEC Filter HST Mag2 Surface Intensity3 Peak3a Ratio3b Size (′′ × ′′ (σ)) Ellipticity4 (e = 1 − b/a) 9 9 9 1 p e S 6 3 v 8 3 2 8 0 9 9 / h p - o r t s a : v i X r a 08005−2356 08:02:40.8 −24:04:44 09452+1330 (IRC+10216) 17423−1755 (Hen3-1475) 19374+2359 09:47:57.4 +13:16:43 17:45:14.2 −17:56:47 19:39:35.6 +24:06:28 16342−3814 16:37:39.9 −38:20:17 Eastern Lobe Western Lobe Eastern Lobe Western Lobe 17150−3224 17:18:19.7 −32:27:21 Southeastern Lobe Northwestern Lobe Southeastern Lobe Northwestern Lobe 17441−2411 17:47:13.5 −24:12:50 Northern Lobe Southern Lobe Northern Lobe Southern Lobe 20028+3910 20:04:35.9 +39:18:45 Northern Lobe Southern Lobe Northern Lobe Southern Lobe 22574+6609 22:59:18.3 +66:25:47 23321+6545 23:34:23.1 +66:01:51 . . . . . . . . . . . . . . . . . . F555W 1.01 e−13 F814W 1.91 e−13 F450W . . . F814W 5.37 e−15 F555W 2.59 e−14 F814W 3.78 e−14 F555W 2.39 e−16 F814W 1.89 e−15 F555W 2.02 e−15 1.22 e−16 1.74 e−15 F814W 5.99 e−15 8.81 e−17 5.12 e−15 F450W 3.26 e−15 9.51 e−16 2.44 e−15 F814W 1.24 e−14 9.23 e−15 1.11 e−14 F450W 1.04 e−15 6.11 e−16 3.29 e−16 F814W 6.98 e−15 5.18 e−15 3.50 e−15 F555W 3.60 e−16 1.26 e−17 3.52 e−16 F814W 1.19 e−15 4.25 e−17 1.18 e−15 F555W 1.16 e−17 F814W 1.16 e−16 F555W 3.15 e−18 F814W 1.06 e−17 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.38 10.70 . . . 14.54 12.87 12.46 17.96 15.71 15.64 18.68 15.80 14.46 19.04 14.63 15.12 16.45 15.43 13.07 13.99 13.78 16.35 16.93 17.61 14.29 14.61 15.04 17.51 21.15 17.53 16.21 19.83 16.22 21.24 18.74 22.65 21.34 1.1 e−1 2.5 e−1 . . . 2.0 e−3 5.8 e−3 2.2 e−2 2.4 e−5 3.5 e−4 . . . 7.1 e−5 9.2 e−4 . . . 2.7 e−4 1.3 e−3 . . . 1.6 e−5 4.9 e−5 . . . 1.9 e−4 2.3 e−4 . . . 4.2 e−5 1.7 e−5 . . . 3.0 e−4 1.2 e−4 . . . 2.0 e−7 6.7 e−5 . . . 2.4 e−6 1.7 e−4 6.3 e−6 3.5 e−5 1.3 e−5 3.1 e−5 990 1200 . . . 250 1900 10000 32 130 9 120 8 50 10 32 170 210 5 2 28 11 1 330 1 70 6 12 9 12 2.68 × 1.42 (2.3) 2.58 × 1.13 (1.6) . . . 3.12 × 3.00 (∼1) 3.75 × 1.98 (5)i 3.70 × 2.19 (3)i 3.29 × 1.63 (2.6) 3.07 × 1.57 (4) 2.55 × 0.71 (3) 0.52 × 0.39 (3) 1.27 × 0.71 (3) 2.71 × 0.82 (3) 0.76 × 0.56 (2) 1.29 × 0.82 (2) 12.59 × 6.12 (5) 2.70 × 1.22 (10) 2.57 × 1.37 (10) 8.96 × 4.72 (3) 2.51 × 1.61 (5) 2.50 × 1.39 (5) 4.43 × 1.76 (2) 1.61 × 0.90 (3) 1.43 × 0.57 (3) 6.47 × 3.77 (5) 1.28 × 0.80 (10) 1.46 × 0.44 (10) 3.48 × 2.11 (3) 0.74 × 0.70 (∼ 1)ii 1.71 × 1.27 (7) 4.42 × 3.13 (3) 1.69 × 1.11 (5)ii 0.14 × 0.48 (5) 0.82 × 0.50 (4.7) 0.73 × 0.48 (2) 0.26 × 0.16 (8) 0.40 × 0.35 (5) 0.47 0.44 . . . 0.04 0.47 0.41 0 50 0.49 0.72 0.25 0.44 0.70 0.26 0.36 0.51 0.55 0.47 0.47 0.36 0.44 0.60 0.44 0.60 0.42 0.38 0.70 0.39 0.05 0.26 0.29 0.34 0.71 0.39 0.34 0.38 0.13 1In units of ergs/s/cm2/A. 2HST WFPC2 system magnitudes (STMAG). 3In units of ergs/s/cm2/A/sr: 3a-peak intensity of the central star or lobe, 3b-star- or peak-to-nebula intensity ratio. 4a and b are respectively major and minor axis lengths. iAssociated with point symmetric jets which extend 16′′.70 in V and 16′′.92 in I. iiThe northern lobe appears to have an oblate shape. 9 9 9 1 p e S 6 3 v 8 3 2 8 0 9 9 / h p - o r t s a : v i X r a Observed and Derived Properties of Stellar Sources TABLE 3 IRAS ID Obs. Coord. (J2000) WFPC2 1 Fλ RA DEC Filter HST Mag2 Surface Peak Intensity3 04386+5722 04:42:49.0 +57:27:47 05113+1347 05:14:07.8 +13:50:28 10:18:07.6 −28:59:32 F555W F814W F555W F814W F469N 1.63 e−14 2.09 e−13 3.68 e−14 9.15 e−14 2.44 e−11 13.37 10.60 12.49 11.50 5.43 5.1 e−2 4.2 e−1 4.8 e−2 1.3 e−1 6.7 e+1 10158−2844 (HR4049) 15465+2818 (RCrB) 20043+2653 15:48:34.4 +28:09:25 F469N 1.56 e−12 8.42 5.4 e+0 20:06:22.7 +27:02:32 22142+5206 22:16:10.1 +52:21:13 F555W > 7.39 e−16 < 16.73 F814W > 8.36 e−16 < 16.59 F555W > 1.83 e−15 < 15.74 F814W > 1.52 e−15 < 15.95 > 8.6 e−4 > 4.8 e−4 > 7.7 e−4 > 7.4 e−4 1In units of ergs/s/cm2/A. 2HST WFPC2 system magnitudes (STMAG). 3In units of ergs/s/cm2/A/sr. Optical vs. Mid-Infrared Morphologies and Other Properties of PPN Candidates TABLE 4 # IRAS ID Object Type Mid-IR Morph.i SED Type Spectral Class Chemii (*/CS) biii (deg) z iii (pc) Referencesiv SOLE SOURCES 1 2 5 6 9 7 9 8 9 16 1 18 p 19 e S 23 25 6 02229+6208 04296+3429 05341+0852 06530−0213 07134+1005 07430+1115 17436+5003 18095+2704 19114+0002 20462+3416 22272+5435 21282+5050 PPN PPN PPN PPN PPN PPN PPN PPN PPN? PN PPN PN DUPLEX SOURCES 3 v 9 8 10 3 13 2 14 8 0 15 9 17 9 20 / h 21 p 26 - 27 o r t s a : v i 3 X 4 r a 11 12 22 24 08005−2356 09452+1330 16342−3814 17150−3224 17423−1755 17441−2411 19374+2359 20028+3910 22574+6609 23321+6545 04395+3601 06176−1036 AFGL 2688 PPN AGB AGB PPN PN PPN PPN PPN PPN PPN PN PPN PPN STELLAR SOURCES PPN PPN RCrB RCrB AGB YSO 04386+5722 05113+1347 10158−2844 15465+2818 20043+2653 22142+5206 unknown unresolved unresolved unresolved toroidal unresolved toroidal unresolved toroidal . . . toroidal toroidal unresolved . . . core/elliptical core/elliptical . . . unknown core/elliptical unresolved unresolved unresolved core/elliptical core/elliptical core/elliptical unresolved unresolved unresolved unresolved unresolved unresolved IVa IVa IVa IVa IVb IVa IVb IVa IVb IVb IVa III II I III III II III II II/III II II/III III III III IVa IVa .. .. I II G8-K0Ia G0Ia G20-Ia F0Iab F5Iab G50-Ia F3Ib F3Ib G5Ia B1Ia G2Ia .. F5e C9.5 M G2 .. .. .. .. .. .. B0 B8V F5Iae M G8Ia B9.5Ib-II F0-F8pe .. .. C/C C/C C/C C/C? C/C C/C O/O O/O O/O O/.. C/C C/C C/O C/.. ../O ../O ../O ../C? ../O ../C C/C ../C C/C C/O C/C ../.. C/C C/.. C/.. O/.. C/.. 1.5 -9.1 -12.2 -0.1 10.0 17.1 30.9 20.2 -5.0 -5.8 -2.5 -0.1 3.6 45.1 5.8 3.0 5.8 2.2 1.0 4.1 6.0 4.3 -6.5 -11.8 -6.5 7.5 -14.3 22.9 51.0 -2.7 -3.6 60 650 2100 4 520 90 620 660 350 290 130 6 250 140 220 130 510 110 90 240 520 .. 190 70 140 .. 1200 270 1500 .. 280 22,48,52 1,8,12,19,20,23,31,37,44,46,47,53,62 1,12,15,22,46,48,53,54 12,28,29,48,50,53 1,5,23,25,33,35,36,37,44,46,47,62 12,22,23,31,48,52 1,5,11,23,25,30,43,44,45,48,56,62 5,12,23,24,26,33,34,40,44,48,62,63 5,18,23,25,32,44,51,62,63 12,13,49,59 1,12,14,19,21,23,37,44,46,47,62,63,67 47,69,70 11,19,29,31,58,62 1,6,40,44,57,62 12,17,42,48,55,63 12,17,27,28,29,38,39,44,48,53,63,66 12,61 3,4,12,28,29,38,44,48,53,60,63,66 26,29,34,40,48,63 26,29,44,48 20,26,29,48 29,44,48 20,74 68,73 71,72 22,48,52 1,9,11,12,19,23,36,48 1,48,62,64,65 7,10,48,62 46 9,46,48 iMarginally resolved sources have an inconclusive "unknown" morphology classification (Meixner et al. 1999). iiChemistry: * -- Photosphere, CS -- Circumstellar shell iiiGalactic latitude, b, is used with the averaged distance in the literature to estimate the Galactic height, z. ivReferences: Photometric data are obtained from references in boldface; 1.Bakker et al. 1997, 2.Blommaert et al. 1993, 3.Bobrowsky et al. 1995, 4.Borkowski et al. 1997, 5.Bujarrabal et al. 1992, 6.Campbell et al. 1976, 7.Clayton et al. 1997, 8.Decin et al. 1998, 9.Dobashi et al. 1999, 10.Feast et al. 1997, 11.Garc´ıa-Lario et al. 1990, 12.Garc´ıa-Lario et al. 1997a, 13.Garc´ıa-Lario et al. 1997b, 14.Geballe et al. 1992, 15.Geballe et al. 1990, 16.Gillett et al. 1986, 17.Gutler et al. 1996, 18.Hawkins et al. 1995, 19.Hrivnak 1995, 20.Hrivnak et al. 1991a, 21.Hrivnak et al. 1991b, 22.Hrivnak et al. 1999a, 23.Hrivnak et al. 1994, 24.Hrivnak et al. 1988, 25.Hrivnak et al. 1989, 26.Hrivnak et al. 1999b, 27.Hu et al. 1993a, 28.Hu et al. 1993b, 29.Hu et al. 1994, 30.Humphreys et al. 1974, 31.Iyengar et al. 1997, 32.Kastner et al. 1995, 33.Klochkova 1995, 34.Kwok et al. 1987, 35.Kwok et al. 1990, 36.Kwok et al. 1995, 37.Kwok et al. 1989, 38.Kwok et al. 1996, 39.Kwok et al. 1998, 40.Lawrence et al. 1990, 40.Le Bertre 1988, 42.Likkel et al. 1988, 43.Likkel et al. 1987, 44.Loup et al. 1993, 45.Luck et al. 1991, 46.Manchado et al. 1989 47.Meixner et al. 1997, 48.Meixner et al. 1999, 49.Parthasarathy 1993, 50.Reddy 1999, 51.Reddy et al. 1999a, 52.Reddy et al. 1999b, 53.Reddy et al. 1996, 54.Reddy et al. 1997, 55.Sahai et al. 1999, 56.Skinner et al. 1994, 57.Skinner et al. 1998, 58.Slijkhuis et al. 1991, 59.Smith et al. 1994, 60.Su et al. 1998, 61.te Lintel Hekkert 1991, 62.TDG94, 63.VHG89, 64.Waelkens et al. 1991, 65.Waters et al. 1989, 66.Weintraub et al. 1998, 67.Zacs et al. 1995, 68.Cohen et al. 1975, 69.Cohen et al. 1987, 70.Kwok et al. 1993, 71.Sahai et al. 1998, 72.Skinner et al. 1997, 73.Waters et al. 1998, 74.Westbrook et al. 1975
astro-ph/9902070
1
9902
1999-02-04T14:33:53
The zabs=zem Absorption Line Systems Toward QSO J2233-606 in the Hubble Deep Field South: NeVIII770,780 absorption and partial coverage
[ "astro-ph" ]
Results of a careful analysis of the highly ionized absorption systems, observed over the redshift range 2.198--2.2215 in the zem=2.24 HDFS-QSO J2233-606, are presented. Strong OVI and NeVIII absorptions are detected. Most of the lines show signature of partial coverage which varies from species to species. This can be understood if the clouds cover the continuum emission region completely and only a fraction of the broad emission line region. Using photo-ionization models we analyze in more detail the component at zabs = 2.198. Absolute abundances are close to solar but the [N/C] abundance ratio is larger than solar. This result, which is consistent with the analysis of high-z QSO broad emission-lines, confirms the physical association of the absorbing gas with the AGN. The observed column densities of NIV, NV and NeVIII favor a two-zone model for the absorbing region where NeVIII is predominantly produced in the highly ionized zone. It is most likely that in QSO J2233-606, the region producing the NeVIII absorption can not be a warm absorber. One of the Lyalpha absorption lines at zabs = 2.2215 has a flat bottom typical of saturated lines and non-zero residual intensity in the core, consistent with partial coverage. There is no metal-line from this Lyalpha cloud detectable in the spectrum which suggests either large chemical inhomogeneities in the gas or that the gas is very highly ionized. If the latter is true the cloud could have a total hydrogen column density consistent with that of X-ray absorbers. It is therefore of first importance to check whether or not there is an X-ray warm-absorber in front of this QSO.
astro-ph
astro-ph
A&A manuscript no. (will be inserted by hand later) Your thesaurus codes are: 11.17.1;11.17.4 J2233-606 ASTRONOMY AND ASTROPHYSICS 15.10.2018 9 9 9 1 b e F 4 1 v 0 7 0 2 0 9 9 / h p - o r t s a : v i X r a The zabs ∼ zem absorption line systems toward QSO J2233- 606 in the Hubble Deep Field South: Ne viiiλλ770,780 absorption and partial coverage ⋆ Patrick Petitjean1,2 and R. Srianand3 1Institut d'Astrophysique de Paris -- CNRS, 98bis Boulevard Arago, F-75014 Paris, France 2UA CNRS 173 -- DAEC, Observatoire de Paris-Meudon, F-92195 Meudon Cedex, France 3IUCAA, Post Bag 4, Ganesh Khind, Pune 411 007, India Abstract. Results of a careful analysis of the highly ion- ized absorption systems, observed over the redshift range 2.198 -- 2.2215 in the zem = 2.24 HDFS-QSO J2233-606, are presented. The strength and covering factor of the O vi and Ne viii absorption lines suggest that the gas is closely associated with the AGN. In addition, most of the lines show signature of partial coverage and the covering factor varies from species to species. This can be under- stood if the clouds cover the continuum emission region completely and only a fraction of the broad emission line region. Using photo-ionization models we analyze in more de- tail the component at zabs= 2.198, for which we can derive reliable estimates of column densities for H i and other species. Absolute abundances are close to solar but the [N/C] abundance ratio is larger than solar. This result, which is consistent with the analysis of high-z QSO broad emission-lines, confirms the physical association of the ab- sorbing gas with the AGN. The observed column densities of N iv, N v and Ne viii favor a two-zone model for the absorbing region where Ne viii is predominantly produced in the highly ionized zone. It is most likely that in QSO J2233-606, the region producing the Ne viii absorption can not be a warm absorber. One of the Lyα absorption lines at zabs = 2.2215 has a flat bottom typical of saturated lines and non-zero resid- ual intensity in the core, consistent with partial coverage. There is no metal-line from this Lyα cloud detectable in the spectrum which suggests either large chemical inhomo- geneities in the gas or that the gas is very highly ionized. Send offprint requests to: Patrick Petitjean ⋆ Based in part on observations obtained with the NASA/ESA Hubble Space Telescope by the Space Telescope Science Institute, which is operated by AURA, Inc., under NASA contract NAS 5 -- 26555; observations collected at the European Southern Observatory, La Silla, Chile; and observa- tions collected at the Anglo-Australian Observatory. If the latter is true the cloud could have a total hydrogen column density consistent with that of X-ray absorbers. It is therefore of first importance to check whether or not there is an X-ray warm-absorber in front of this QSO. Key words: quasars: absorption lines, quasars: individ- ual: J2233-606 1. Introduction The QSO J2233-606 (zem = 2.24) has been given tremen- dous interest as it is located in the middle of the STIS Hubble deep field south making this field an ideal target for studying the connection between the diffuse gaseous component of the universe and galaxies (Ferguson 1998). The spectrum of this QSO shows several associated sys- tems, (i.e. systems with zabs ≃ zem ), at zabs ∼ 2.2 with broad C iv and N v absorption lines (Sealey et al. 1998, Savaglio 1998, Outram et al. 1998). HST STIS spectra, together with the available ground-based data, provide different pieces of information about these systems over the rest-wavelength range 375-2800 A. Associated systems have been intensively studied in the past few years because they are believed to be inti- mately related to the central engines of AGNs as: (i) they frequently show absorption due to high-ionization lines; (ii) they have been convincingly shown to have metallic- ities of the order of or above solar (Petitjean et al. 1994, Savaglio et al. 1994, Hamann 1997); (iii) the absorbing gas does not cover the background emitting region com- pletely (Petitjean et al. 1994, Hamann et al. 1997b, Barlow & Sargent 1997). In a few cases it has been shown that the optical depth of the high-ionization lines varies with time on scales of a year (Hamann et al. 1997b) as is the case for some broad absorption line systems (e.g. Barlow 2 et al. 1992). This requires a recombination time-scale of less than a year and hence high particle density in the absorbing gas if the observed variability is caused by the change in the ionizing conditions. Soft X-ray spectra of an appreciable fraction of Seyfert- I galaxies show K-shell absorption edges of ionized oxygen (O vii and O viii; Reynolds 1997, George et al. 1998). Rapid variability of these absorption edges suggest that the absorbing gas is very close to the central engine. Moreover, there is a one-to-one correspondence between the presence of associated absorption systems and X-ray "warm absorbers" in Seyfert galaxies (Crenshaw et al. 1998). The case for a unified model for X-ray and UV absorbers, although attractive, is not completely convinc- ing yet however. Eventhough Mathur et al. (1994) showed that the X-ray and UV absorptions seen in 3C351 can be reproduced by a single-cloud model, there are cases where two different ionized zones are needed even to produce the optical depth ratios of O vii and O viii (see Reynolds 1997). Some Seyfert-I galaxies show signatures of the exis- tence of optically thin emitting clouds in the BLR. Shield, Ferland & Peterson (1995) have shown that the properties of the optically thin clouds in the inner BLR are consis- tent with that of warm absorbers (see also Porquet et al. 1998). Analysis of the broad Ne viiiλ774 emission line in QSOs shows that the Ne viii-emitting regions have ioniza- tion parameters in the range 5 -- 30, total hydrogen column densities of the order of 1022.5 cm−2 and average covering factors >30% for solar abundances and a nominal QSO spectrum (Hamann et al. 1997b). The ionization condi- tions in these emitting clouds are similar to those of warm absorbers. In order to investigate in more detail the nature of as- sociated systems and their possible connection to warm absorbers, absorptions from species with a wide range of excitation should be studied. Morevover column den- sities should be determined taking into account the ef- fect of partial coverage. In this prospect, absorption from Ne viii, if present, is crucial as its ionization potential, 207 eV, is much higher than the ionization potential of other easily observable species. To our knowledge the Ne viiiλλ770,780 absorption doublet has been detected in only two associated systems, in the line of sight to HS1700+6416 at zabs = 2.7126 (Petitjean et al. 1996; see the HST spectrum in Vogel & Reimers 1995) and in the line of sight to UM675 (Hamann et al. 1995, see the spec- trum in Hamann et al. 1997a) at zabs = 2.1340. Analysis of the latter system leads the authors to conclude that, al- though the detection of Ne viii provides strong evidence for a link between the associated system and the warm ab- sorber, the total hydrogen column density in the Ne viii phase is too small to produce the warm absorber phe- nomenon. Recently Telfer et al. (1998) have done a similar study (including detection of Ne viii) of the broad absorp- tion line system present in QSO SBS 1542+531. Note that in all previous studies Ne viii absorption is investigated with low dispersion FOS spectra. In the following we discuss in detail the associated sys- tems in the line of sight to QSO J2233-606 and especially the unambiguous detection of strong Ne viii absorption. We briefly present the method to derive column densi- ties in case of partial coverage in Section 2; describe the data in Section 3 and the individual absorption systems in Section 4, discuss the physical consequences of the obser- vations in Section 5 and draw our conclusions in Section 6. 2. Partial coverage When an absorbing cloud does not cover the background source completely, the observed residual intensity in the normalized spectrum, Rλ, can be written as, Rλ = (1 − fc) + fc × exp(−τλ), (1) where, τλ and fc are the optical depth and covering factor respectively. The latter is the ratio of the number of pho- tons produced by the region of the background source that is occulted by the absorbing cloud to the total number of photons (Srianand & Shankaranarayanan 1999). If two ab- sorption lines with rest-wavelengths λ1 and λ2 originate from the same ion (usually it will be doublets or multi- plets), their residual intensities, R1 and R2, at any veloc- ity v with respect to the centroid of the lines are related by, R2(v) = 1 − fc2 + fc2 ×(cid:18) R1(v) − 1 + fc1 fc1 (cid:19)γ (2) where fc1 and fc2 are the covering factors calculated for the two lines and, γ = (cid:18) f2λ2 f1λ1(cid:19) with f1 and f2 the oscillator strengths (see e.g. Petitjean 1999). The value of γ is close to 2 for doublets. Though our intuition says that fc1 = fc2 it need not be true in general (Srianand & Shankaranarayanan 1999). In- deed due the complex velocity structure in the BLR, pho- tons that could be absorbed by lines 1 and 2 (even in the case of doublets) originate from spatially distinct regions in the BLR. A cloud can thus be black in line 1 (covering the BLR region emitting at the corresponding wavelength λ1), but not in line 2 if it does not cover at the same time the region emitting photons with wavelength λ2. Also, the interpretation of the relative covering factors for various ions is not straightforward. Indeed, the covering factor of absorption lines seen over the wavelength range of the QSO spectrum that is dominated by the continuum may be larger than the covering factor of lines present on top of the QSO emission lines as the extension of the BLR is larger than that of the continuum emission region. Table 1. Spectroscopic data Coverage Inst./Tel. R S/N Ref. (A) per pixel 30000 E230M-STIS/HST 2275-3118 1126-1721 ∼1100 G140L-STIS/HST G230L-STIS/HST 1585-3173 ∼550 G430M-STIS/HST 3025-3565 ∼6000 35000 UCLES/AAT EMMI/NTT 21000 3530-4390 4386-8270 1.5 - 4 10 - 18 20 - 45 5 - 15 5 - 30 8 - 10 a a a a b c a Ferguson (1998), b Outram et al. (1998), c Savaglio (1998) 3. Data In the following analysis we use spectra of QSO J2233- 606 obtained with different instruments and telescopes. Table 1 gives the instrumentation, wavelength coverage, spectral resolution, typical signal-to-noise ratio and refer- ence for the data available. Ultra-violet and optical spec- tra were taken by the HST HDF-S STIS Team (Ferguson 1998) with the STIS E230M, G140L, G230L and G430M gratings on board of the Hubble Space Telescope. Optical data was obtained at the AAT and ESO/NTT by Outram et al. (1998) and Savaglio (1998) respectively. We have cor- rected the data for fluctuations in the zero-level by fitting low order polynomials to the bottom of the strong sat- urated Lyα lines. The echelle data have been smoothed using a gaussian filter of width 3 pixels. The continuum was fit with low-order polynomials and the spectrum was normalized. 4. Description of the associated system 4.1. Overview The absorption profiles produced by the associated ab- sorbers are shown for the most important transitions on a velocity scale in Fig. 1. Complementary identifications in the G230L spectrum of lower quality because of the pres- ence of the LLS break at ∼2700 A are given in Fig. 2. The two spectra obtained in 1997 (end of october) and 1998 (beginning of october) are superimposed on Fig. 2 to look for variability. It can be seen that the modest signal-to- noize ratio prevents any firm conclusion about the vari- ability of the Ne viii absorption lines. There are absorp- tion features near the expected position of Mg xλλ609,624 but, due to poor spectral resolution, this cannot be as- certained. On the contrary, O v and probably N iii are present. The maximum column density of He i found in the models discussed below is ∼1011 cm−2. We thus do not believe that the possible line at ∼1875 A seen in only one of the spectra can be He iλ584. The emission redshift of QSO J2233-606, derived from the high-ionization emission lines C iv, C iii] + Al iii, zem = 2.237, is smaller than the redshift derived from Mg ii, zem = 2.252, by about 1390 km s−1 (Sealey et al. 3 1998). Considering the Mg ii redshift as more represen- tative of the intrinsic redshift (Carswell et al. 1991), the associated absorptions, seen over the redshift range 2.198 -- 2.2215, have outflow velocities relative to the quasar of 2800 -- 5000 km s−1 which is modest compared to usual as- sociated or BAL outflows. -1000 -1000 0 0 1000 1000 2000 2000 -1000 0 1000 2000 Fig. 1. Absorption profiles of different transitions in the asso- ciated systems observed along the line of sight to J2233-606. The zero velocity is taken at z = 2.20. The vertical dashed lines mark redshifts 2.1982, 2.2052, 2.2075 and 2.2215 from the left to the right. Note the component at +950 km s−1 (zabs = 2.21) with O vi and Ne viii but no detectable H i absorptions. 4.2. zabs= 2.2215 There is a Lyα line at this redshift with flat bottom, con- sitent with line saturation, but with non-zero residual in- 4 1800 1900 2000 2100 2200 0 2200 2300 2400 2500 Observed Wavelength 2600 2700 Fig. 2. Possible identifications of lines from the associated sys- tem in the G230L spectrum. The vertical dashed lines mark the redshift range 2.198 -- 2.210. The two spectra obtained in 1997 (end of october) and 1998 (beginning of october) are superim- posed to look for variability. It can be seen that the modest S/N ratio prevents any firm conclusion about the variability of the Ne viii absorption lines. tensity. From the latter, it can be seen on Fig. 3 that the minimum covering factor for this line is fc ∼ 0.7. However before drawing any conclusions it is important to show that the feature is not due to blending of a few weaker lines. In Fig. 3 we plot the line profiles of the other de- tected Lyman series lines from this system. It can be seen that the Lyβ line is very strong. Moreover the residual intensity in the Lyβ line is smaller than the residual in- tensity in the Lyα line, consistent with saturation of the Lyα line. We first assume that the covering factor is the same for Lyα and Lyβ . In the middle panel we plot the observed Lyβ profile together with the predicted Lyβ profiles com- puted from the Lyα profile for three values of the covering factor: dotted, short -- dashed and long -- dashed lines are for covering factors 1, 0.8, 0.70 respectively. The predicted Lyβ profiles are inconsistent with the ob- served Lyβ profile and the latter seems too strong even for the minimum covering factor acceptable for the Lyα line (fc ∼ 0.7). The numerous saturated lines present in this part of the spectrum assures that the error in the zero level determination cannot explain the discrepancy. Although we cannot reject the presence of weak Lyα absorption lines superimposed with the Lyβ absorption, especially in the blue-wing, the good wavelength coincidence between Lyβ and Lyα seems to indicate that the contamination cannot be large. One way to explain the apparent strength of the Lyβ line is to assume that the covering factor for Lyβ is larger than for Lyα . Note that this is consitent with the QSO Lyα emission line to be stronger than the Lyβ emis- sion line. If true, then we can expect that the covering factor of the Lyγ line be even larger. In the top panel we plot the observed wavelength range of the Lyγ line and the predicted profile using the Lyβ profile for covering factors 1 (dotted line) and 0.9 (dashed lines). The best match is obtained for complete coverage though this does not reproduce the Lyγ very well. Smaller values of the covering factor predict too strong a Lyγ line. We conclude that, in order to understand the residual in- tensities in the different Lyman series absorption lines, it must be assumed that the covering factor increases from Lyα to Lyγ. It is thus likely that the absorbing cloud at zabs = 2.2215 completely covers the continuum source and partially covers the BLR. This system does not show absorption due to any de- tectable heavy element transitions either in the optical data (Outram et al 1998; Savaglio, 1998) or in the HST data. If partial coverage is a signature of physical associ- ation between the absorbing gas and the AGN, then the lack of metal lines in this system could suggest that there are large inhomogeneities in the chemical enrichment of the gas physically associated with central engines of QSOs. However, on the contrary, if we presume that the gas as- sociated with the central regions of the quasar is more or less uniformly enriched, then this system could correspond to very highly ionized gas. The ionization state should be such that all observable metal transitions are weak and undetectable. This condition demands log H i/H to be smaller than -8 (e.g. Hamann 1997) and log N (H)>22. Note that such a cloud could be related to the warm ab- sorbers. Another possibility is that the gas is extremely metal- poor and is produced by an intervening cloud with sizes less than the BLR (i.e. few pc). Not only this is much smaller than the dimensions derived for intervening Lyα clouds using adjacent lines of sight (e.g. Petitjean et al. 1998) but also these clouds would have been detected by previous surveys far away from the QSO. 4.3. zabs= 2.207 The C iv , N v and Lyα absorption lines produced by this system are shallow and broad (∼500 km s−1) like a minia- turised Broad Absorption Line system (BALs). Outram 1.2 1.2 1.2 1 1 1 0.8 0.8 0.8 0.6 0.6 0.6 0.4 0.4 0.4 0.2 0.2 0.2 1.20 0 0 1.2 1.2 1.2 1 1 1 1 0.8 0.8 0.8 0.8 0.6 0.6 0.6 0.6 0.4 0.4 0.4 0.4 0.2 0.2 0.2 0.2 1.20 0 0 0 1.2 1 1 0.8 0.8 0.6 0.6 0.4 0.4 0.2 0.2 0 0 -200 0 200 Fig. 3. Analysis of partial coverage in the zabs= 2.2215 system. The observed Lyα profile is plotted in the bottom panel. The middle panel shows the observed Lyβ profile (solid) together with the predicted profiles computed from Lyα assuming cov- ering factors fc = 1, 0.8 and 0.70 for the dotted, short-dashed and long-dashed lines respectively. The top panel shows the ob- served Lyγ profile with the predicted profiles computed from Lyβ assuming covering factors fc = 1 and 0.9 for the dotted and dashed lines respectively. et al. (1998) discuss the N v and Lyα absorptions from this system. They could not fit the N v doublet when as- suming 100% coverage. However, they managed to obtain a consistent fit after correcting the continuum by sub- tracting a Gaussian centered at 3938 A with FWHM = 670 km s−1 and maximum depth 19 percent of the origi- nal continuum level. They used a three-component model with large velocity dispersions. Savaglio (1998) observed the C iv absorption doublet from this system. She could fit the doublet with six components. The O vi and Ne viii doublets are detected in the G430M and E230M spec- tra respectively. Note that the O vi doublet is observed at slightly lower resolution than N v and C iv and the Ne viii doublet is found in a low S/N region blueward the Lyman limit of the moderately thick system at zabs= 1.87. Finally, strong O vλ629 is seen at ∼2020 A as ex- pected and possibly Mg xλ624 at λ2000 A. The stronger line of the C iv , N v , O vi and Ne viii doublets, together with the Lyα line, are plotted on Fig. 4 on a velocity scale. The dotted and dashed lines are the predicted velocity profiles computed from the profile of the second transition of the doublets for different values of the covering factor. Here again we assume identical covering 5 1 1 1 0.5 0.5 0.5 0 0 0 1 1 1 0.5 0.5 0.5 0 0 0 1 1 1 0.5 0.5 0.5 0 0 0 1 1 1 0.5 0.5 0.5 0 0 0 1 1 0.5 0.5 0 0 -200 0 200 Fig. 4. Analysis of partial coverage in the zabs= 2.207 system. The observed Lyα profile is plotted in the bottom panel. The solid curves in the other panels are the observed profiles of the stronger line of the doublets. The dotted and short-dashed lines are the predicted velocity profiles computed from the profile of the second transition of the doublet and assuming covering factors fc = 0.35 (dotted line) and 0.50 (dashed line). factors for both transitions. It can be seen that in order to reproduce the residual intensities of N v we need a cov- ering factor of the order of 0.35 at v ∼ +150 km s−1 and 0.50 at v ∼ -- 100 km s−1. The difference between the two models with fc = 0.35 and 0.50, is, at these places, of the order or larger than 0.1 (10% of the normalized contin- uum; see Fig. 4) when the rms deviation in the spectrum is ∼0.04. The O vi profiles are consistent with a covering factor ∼ 0.7 that cannot be reconciled with the values found for N v . Note that the O vi profiles could be affected by the relatively poor spectral resolution and the derived cov- ering factors should be considered as lower limits. The maximum error in the covering factor for the O vi lines is about 0.1, computed from the error in the residual in- tensity. The S/N ratio over the C iv doublet is not good enough and consistent residual intensities are obtained for a wide range of covering factors. The covering factor re- quired for the Ne viii lines is in the range 0.8-1.0 with 6 an error per pixel of 0.15. The covering factor derived from the Ne viii lines is therefore significantly larger than the one derived from the N v doublet. There seems to be an anti -- correlation between the covering factor and the ionization state or the wavelength. This again is consis- tent with clouds partially covering the BLR while covering most of the continuum emission region. Another interesting observation concerns the doubly- ionized species. It can be seen on Fig. 1 that O iiiλ832 is certainly blended with other lines; that there may be a shallow absorption at the expected position of C iiiλ977 although most of it should be Lyγ; and that there is a strong absorption at the expected position of N iiiλ989. This line cannot be Lyβ at zabs = 2.093 as there is no corresponding Lyα line. We cannot reject the hypoth- esis that this is an intervening Lyα line as the Lyβ range is of very poor S/N ratio. Interestingly enough, there is an absorption feature at the expected position of N iiiλλ684,685 (see Fig. 2). Although such a strong N iii absorption would be very surprizing (see next Section and Fig. 6), the presence of doubly-ionized species cannot be ruled out. The corresponding absorptions from triply ion- ized species, although weak and noisy, could be present (see in particular O ivλ787 and N ivλ765). If true, this would imply that the medium has two phases of low and high-ionization (see next Section). Since the possible strong N iiiλ989 line goes to zero, suggesting complete coverage; we note that there is a ten- dency for the absorption lines redshifted on top of the emission lines (here C iv and N v ) to have covering fac- tors smaller than lines redshifted in parts of the spectrum free from emission-lines (Ne viii and N iii). This further supports the conclusion that the gas covers the continuum emitting region but only part of the BLR. 4.4. zabs= 2.198 Lyα and N v absorptions produced by this system are de- tected by Outram et al. (1998) who already noted the incomplete coverage of the background source. They con- jectured that the absorbing cloud is larger than the contin- uum emitting region and smaller than the BLR. Savaglio (1998) has noted that single as well as two component fits to the C iv doublet result in a very poor fit, again suggesting partial coverage. O vi and Ne viii doublets are detected in the HST spectra. The stronger lines of the dou- blets together with Lyα are plotted on Fig. 5 on a velocity scale. The dotted and dashed lines are the predicted veloc- ity profiles computed from the second transition of dou- blets assuming different values of the covering factor (dot- ted, short-dashed and long-dashed lines are for fc = 0.7, 0.8 and 0.9 respectively). Here again we assume identical covering factors for both transitions. The N vλ1242 line of this system is blended with N vλ1238 of the system at 2.208. We have subtracted the contribution due to the N vλ1238 line before doing the analysis. The covering fac- 1 1 1 1 0.5 0.5 0.5 0.5 0 0 0 0 1 1 1 1 1 0.5 0.5 0.5 0.5 0.5 0 0 0 0 0 1 1 1 1 0.5 0.5 0.5 0.5 0 0 0 0 1 1 1 1 0.5 0.5 0.5 0.5 0 0 0 0 1 1 0.5 0.5 0 0 -200 -100 0 100 200 Fig. 5. Analysis of partial coverage in the zabs= 2.198 system. The observed Lyα profile is plotted in the lowest panel. The solid curves in other panels are the observed profiles of the strongest transition of the doublet. The dotted, short-dashed and long-dashed lines are the predicted velocity profiles com- puted using the profile of the weakest member of the doublets and assuming covering factor of fc = 0.7, 0.8 and 0.9 respec- tively. tor required to fit the N v doublet is ∼ 0.7±0.04. However the O vi profiles require values larger than 0.85±0.04. It is quite likely that the Ne viiiλ780 is blended with some other line in the blue wing. Also the Ne viiiλ770 profile is very noisy and is consistent with a wide range of covering factors. It is interesting to note that inspite of the poor signal to noise ratio, the C iv profiles clearly suggest that the covering factor is less than 0.7 and we have obtained a consistent fit for fc = 0.5. Thus like the zabs= 2.207 sys- tem, this system also shows different covering factor for different transitions; the covering factor beeing lower for C iv than for N v . 5. Physical conditions in the zabs = 2.198 system In this section we study the physical conditions in the zabs= 2.198 system in greater detail. The column density per unit velocity interval at any velocity, v, with respect to the centroid of the line is given by, N (v) = (mec/πe2) f λ τ (v), and from Eq. (1) the optical depth τ (v) is given by, τ (v) = − ln (cid:18) R(v) − 1 + fc fc (cid:19). (3) (4) Assuming equal covering factor for the two absorption lines of a doublet, fc1 = fc2 = fc, we obtain fc(v) for each doublet using Eq. (2). We then derive the column density per unit velocity interval from Eqs. (3) and (4). The total column density, N , is obtained by integrating N (v) over the velocity interval covered by the absorption line pro- file. For various species we give in Table 2 the absorption parameters (N and b) obtained from Voigt profile fitting (column #3 and #4 respectively) and the column density obtained by integration of the column density per unit ve- locity interval over the velocity profile (column #7). The velocity range (column #5), and the mean covering factor estimated over this range (column #6) are also given. In the case of single lines we use a conservative lower limit of 0.7 for the covering factor. Note that the column density obtained with the Voigt profile fits are lower limits as cor- rections are not incorporated to take into account partial coverage. Lyβ from this system is weak and we could not derive any bound on the covering factor from the Lyα line only. C iiiλ977 is not detected and we derive a two sigma upper limit from the continuum rms at the expected position of the line. Voigt profile fits to the C iv lines are taken from Savaglio (1998). As the estimated covering factor is low, the column density derived by fitting the profile is less by a factor of 5 compared to the estimated column density using the column density per unit velocity interval. N iii and Si iii are not detected whereas a line is present at the expected position of N ivλ765. This line clearly shows two components and the blue component is clearly absent in the N v profile. The further absence of N iii suggests that this component is not real. Thus we have fitted the N iv blue component as an intervening Lyα line and used only the column density obtained from the red component in our analysis. Although the resolution of the spectrum is not very high (∼ 50 km s−1), we have fitted the O vi lines using three components. There is a line at the expected position of O ivλ787. However this line could possibly be Lyγ from the system at zabs= 2.5907. By carefully fitting the Lyα , Lyβ and Lyδ lines in this system, we removed the contribution of the Lyγ and fit the residual as O ivλ787. As discussed before, the estimate of column densities can be somewhat uncertain due to limited S/N ratio (espe- cially for the Ne viii lines) and possible contamination by intervening lines. The derived column densities for H i, C iv , N v and O vi are characteristic of associated systems (Hamann 7 Ne VIII C IV N V MF87 O VI O IV N IV C III -0.5 0 0.5 1 O VI N IV O IV C III O III 0.5 N III 0 O VI O IV Ne VIII C IV N V 1 Ne VIII C IV N V 0 log U 0.5 1 -0.5 N III -0.5 N IV C III O III 16 15 14 13 12 11 16 14 12 16 15 14 13 12 11 Fig. 6. The logarithm of column densities for species indicated next to each curve is given versus the logarithm of the ioniza- tion parameter for ionizing spectra as given by Mathews & Ferland (1987; top panel); or power-laws with index α = -1.5 (middle panel) and -1 (bottom panel). 1997). We run photo-ionization models using Cloudy (Ferland 1996) to study the ionization structure of a plane parallel cloud with neutral hydrogen column density 1014 cm−2, solar metallicities, and illuminated by ionizing radiation fields with different spectra. Although a one- zone model is questionable in such medium, we believe that this is a reasonable representation of the absorbing cloud producing at least C iv, N v and H i (see below for O vi and Ne viii) because the kinematics of the lines are very similar (see Figs. 1 and 5) and the partial coverages discussed in the previous sections indicate small sizes for the cloud. Results for a typical AGN spectrum given by Mathews & Ferland (1987) and two power-law spectra are given in Fig. 6. In the framework of these models, it is apparent 8 Table 2. Parameters for the associated system at zabs=2.198 Ion fc v b z log N cm−2 (km s−1) (km s−1) log N cm−2 2.1982 13.74±0.01 43.99±1.01 -47, 53 H i C iii C iv N iv N v O iv O vi Ne viii ..... 2.1982 2.1980 2.1972 2.1982 2.1979 2.1987 2.1976 2.1981 2.1987 2.1976 2.1981 2.1987 ..... 13.77±0.05 13.73±0.04 13.12±0.05 14.57±0.03 14.43±0.06 13.94±0.08 14.39±0.16 14.87±0.24 14.63±0.14 14.63±0.27 14.79±0.10 14.68±0.09 ..... 19.80±2.40 30.43±1.35 32.51±4.06 31.46±0.58 49.33±2.74 37.70±4.53 a a a 39.11±9.70 25.39±3.85 32.33±3.20 ..... -27, 50 >0.7 <13.95 ..... <13.40 14.48 ∼0.5 >0.7 <14.34 -27, 53 ∼0.7 14.55 -47, 53 >0.7 <14.62 -46, 100 ∼0.8 15.38 -27, 53 ∼0.7 15.10 a Same as for Ne viii that the column densities are reproduced easily. The α = - 1 power-law ionizing spectrum is favored as it minimizes the N (C iv)/N (O vi) ratio. Although metallicities can- not be much less than solar (because of C iv and N v ), it would be difficult to argue for metallicities much larger than solar (especially for oxygen). However, contrary to what is observed, the predicted C iv column density is al- ways larger than that of N v . This suggests that nitrogen is over-abundant compared to carbon. It is interesting to note that the N v lines of the zabs= 2.207 system is also stronger than the lines of C iv suggesting a similar abun- dance pattern. Indeed, using N v /C iv emission line ra- tios, Hamann & Ferland (1992) have shown that nitrogen is over-abundant by a factor of ∼ 2 − 9 in high-redshift QSOs (z ≥ 2.0). They suggested rapid star-formation models to boost the nitrogen abundance through en- hanced secondary production in massive stars. Korista et al. (1996) have also noticed overabundance of nitro- gen with respect to carbon and oxygen in the well-studied BAL system in Q 0226-1024. They could obtain a much better fit of the lines after taking into account the abun- dance pattern due to rapid star-formation. The overabun- dance of nitrogen with respect to oxygen and/or carbon, does not seem to be seen in every associated system how- ever. Indeed, we have good data for the associated systems in Q 0207 -- 003 and Q 0138 -- 381 showing partial coverage. Although solar metallicities are needed to explain the line ratios, there is no indication of enhanced nitrogen abun- dance. From the observed N v to N iv column density ra- tio, it can be seen that log U < -- 0.5 (the ionization parameter U is the ratio of the ionizing photons den- sity to the total hydrogen density). However such a value for the ionization parameter implies that the gas pro- ducing N v and N iv can not account for the observed value of the Ne viii column density without an unreal- istically large neon abundance. Another and more likely explanation is that there are two distinct regions with different ionization parameters and that Ne viii predom- inantly originates from the more highly ionized region. This is supported by the fact that the Ne viii absorption is spread over a larger velocity range than the N v and O vi absorptions. We can check that, even in that case, C iiiλ977 would not be detectable. Indeed, the expected C iii column density is of the order of 1013 cm−2 (see Fig. 6), thus wobs(C iiiλ977) ∼ 0.2 A which is about twice below the detection limit at λ ∼ 3100 A. Since the low-ionization region by itself can account for the observed H i and O vi column densities, the high- ionization region should have N (Ne viii) much larger than N (O vi ) which means log U larger than 0.5 and most probably 1. If we suppose that this component is similar to warm absorbers detected by O vii and O viii edges in the X- rays, then the condition that the optical depth of these edges are larger than 0.1 implies log N (O vii) > 17.55 and log N (O viii) > 18.00. Photoionization models with solar abundances, fail to produce such high O vii and O viii column densities for H i column densities in the range 1013 to 1014 cm−2. Indeed, the ratios NeVIII/Ne, OVII/O and OVIII/O are all maximized over the range of ionization parameters U ∼ 10 -- 100 (see Hamann et al. 1995) where log HI/H ∼ -- 6. For log N (H i) = 14, this cor- responds to log N (H) ∼ 20 and, assuming solar abun- dances, implies that log N (O vii) and log N (O viii) are less than 17. Thus, it is most likely that in QSO J2233- 606, the region producing the Ne viii absorption can not be a warm absorber. It is thus of first importance to study the intrinsic spectral energy distribution of J2233-606 es- pecially in the X-rays. 9 Ferland, G. J. 1996, "HAZY a Brief Introduction to Cloudy", Univ. Kentucky, Dept. Physics & Astron.. Internal rep. George I.M., Turner T.J., Netzer H., et al., 1998, ApJS 114, 73 Korista, K., Hamann, F., Ferguson, J. & Ferland, G. 1996, ApJ, 461,641 Hamann F., 1997, ApJS 109, 279 Hamann F., Barlow T.A., Beaver E.A., 1995, ApJ 443, 606 Hamann F., Barlow T.A., Cohen R.D., Junkkarinen V., Bur- bidge E.M., 1997a, astro-ph/9704235 Hamann F., Barlow T.A., Junkkarinen V., Burbidge E.M., 1997b, ApJ 478, 80 Hamann F., Cohen R.D., Shields J.C., et al., 1998, ApJ 496, 761 Hamann F., Ferland G. J. 1992, ApJ, 391, L53 Mathews W.G., Ferland G.J., 1987, ApJ 323, 456 Mathur S., Wilkes B., Elvis M., Fiore F., 1994, ApJ 434, 493 Outram P. J., Boyle B. J., Carswell R. F., et al., 1998, MNRAS in press (astro-ph/9809404) Petitjean P., 1999, Proceedings of the Les Houches school "For- mation and Evolution of Galaxies"; O. Le Fevre and S. Charlot (eds.), Springer-Verlag, astro-ph/9810418 Petitjean P., Rauch M., Carswell R.F., 1994, A&A 291, 29 Petitjean P., Riediger R., Rauch M., 1996, A&A 307, 417 Petitjean P., Surdej J., Smette A., et al., 1998, A&A 334, L45 Porquet D., Dumont A.-M., Collin S., Mouchet M., 1999, astro- ph/9810333v2 Reynolds C.S., 1997, MNRAS 286, 513 Savaglio S., 1998, AJ 116, 1055 Savaglio S., D'Odorico S., Møoller P., 1994, A&A 281, 331 Shield et al. 1995, ApJ, 441, 507 Sealey M. K., Drinkwater J. M., Webb J. K., 1998, ApJL 499, 135 Srianand & Shankaranarayanan, 1999, astro-ph/9901091 Telfer R.C., Kriss G.A., Zheng W., Davidsen A.F., Green R., 1998, ApJ 509, 132 Vogel S., Reimers D., 1995, A&A 294, 377 6. Conclusion We have studied the associated absorption systems (zabs ∼ zem) in QSO J2233-606, seen over the redshift range 2.198 -- 2.2215, corresponding to outflow velocities relative to the quasar emission redshift (zem = 2.252 from Mg ii) of 2800 -- 5000 km s−1 which is modest compared to usual associated or BAL outflows. We have shown that the Lyα line at zabs= 2.2215 is saturated but has non-zero residual intensity. The cover- ing factor of the gas is of the order of 0.7. This absorption system is unusual in the sense that there is no detectable metal lines. This could reveal chemical inhomogeneities in the gas or, and more likely, this absorption could corre- spond to very highly ionized gas from which no absorption due to heavier element can be detected. If the latter is true, the ionization factor log HI/H could be smaller than -- 8 and the total column density in the cloud larger than log N (H) = 22. Such a cloud could be related to the warm absorbers. Over the redshift range 2.198 -- 2.21, conspicuous Ne viii, O vi, C iv and N v absorptions are seen whereas the H i absorption is weak. From analysing these ab- sorptions we conclude that (i) a two-phase medium of low and high-ionization respectively is required to ex- plain the complete set of column densities; (ii) the abundances are close to the solar value but nitrogen is enhanced with respect to carbon; (iii) although dif- ficult to ascertain there is a tendency for the absorp- tion lines redshifted on top of the emission lines to have covering factors smaller than lines redshifted in parts of the spectrum free from emission-lines; this suggests that the gas covers the continuum emitting region but only part of the BLR. The component at zabs = 2.21 (+950 km s−1 on Fig. 1) is remarkable as, though conspic- uous in Ne viii and O vi, it is undetected in H i, C iv and N v . The column densities derived for this subcomponent by Voigt-profile fitting are log N (Ne viii) = 14.57±0.10; log N (O vi) = 14.05±0.06 and log N (H i) < 13.30, con- sistent with what is expected from a high-ionization zone. Acknowledgements. We would like to thank all astronomers who have provided the data for this project and especially the HST HDF-S STIS team lead by H. Ferguson. We thank the referee, Fred Hamann, for a careful reading of the manuscript. References Barlow T.A., Junkkarinen V.T., Burbidge E.M., et al., 1992, ApJ 397, 81 Barlow T.A., Sargent W.L.W., 1997, AJ 113, 136 Carswell R.F., Mountain C.M., Robertson D.J., et al., 1991, ApJ 381, L5 Crenshaw D.M., Kraemer S.B., Boggess A., et al., 1998, astro- ph/9812265 Ferguson H.C., 1998, Space Telescope Science Institute Newsletter, Vol. 15, No. 3, 6 This article was processed by the author using Springer-Verlag LaTEX A&A style file L-AA version 3.
astro-ph/0507331
1
0507
2005-07-13T18:14:02
Anisotropy Studies Around the Galactic Center at EeV Energies with Auger Data
[ "astro-ph" ]
The Pierre Auger Observatory data have been analyzed to search for excesses of events near the direction of the galactic center in several energy ranges around EeV energies. In this region the statistics accumulated by the Observatory are already larger than that of any previous experiment. Using both the data sets from the surface detector and our hybrid data sets (events detected simultaneously by the surface detector and the fluorescence detector) we do not find any significant excess. At our present level of undestanding of the performance and properties of our detector, our results do not support the excesses reported by AGASA and SUGAR experiments. We set an upper bound on the flux of cosmic rays arriving within a few degrees from the galactic center in the energy range from 0.8-3.2 EeV. We also have searched for correlations of cosmic ray arrival directions with the galactic plane and with the super-galactic plane at energies in the range 1-5 EeV and above 5 EeV and have found no significant excess.
astro-ph
astro-ph
29th International Cosmic Ray Conference Pune (2005) 00, 101 -- 106 Anisotropy Studies Around the Galactic Center at EeV Energies with Auger Data The Pierre Auger Collaboration Presenter: A. Letessier-Selvon (Antoine.Letessier-Selvon@@in2p3.fr),fra-letessier-selvon-A-abs1-he14-oral The Pierre Auger Observatory data have been analyzed to search for excesses of events near the direction of the galactic center in several energy ranges around EeV energies. In this region the statistics accumulated by the Observatory are already larger than that of any previous experiment. Using both the data sets from the surface detector and our hybrid data sets (events detected simultaneously by the surface detector and the fluorescence detector) we do not find any significant excess. At our present level of undestanding of the performance and properties of our detector, our results do not support the excesses reported by AGASA and SUGAR experiments. We set an upper bound on the flux of cosmic rays arriving within a few degrees from the galactic center in the energy range from 0.8-3.2 EeV. We also have searched for correlations of cosmic ray arrival directions with the galactic plane and with the super-galactic plane at energies in the range 1-5 EeV and above 5 EeV and have found no significant excess. 1. Introduction The galactic centre (GC) region provides an attractive target for anisotropy studies with the Pierre Auger Observatory. On the one hand, there have been in the past observations by the AGASA experiment indicating a 4.5σ excess of cosmic rays (CRs) with energies in the range 1-2.5 EeV in a 20◦ radius region centered at (δ, α) ≃ (−17◦, 280◦) [1]. A later search near this region with a reanalysis of SUGAR data [2], though with smaller statistics, failed to corroborate these findings, but reported a 2.9σ excess flux of CRs with energies 0.8 <E<3.2 EeV in a region of 5.5◦ radius centered at (δ, α) = (−22◦, 274◦). On the other hand, since the GC harbors a very massive black hole, it provides a natural candidate for CR accelerator to very high energies. Following the recent high significance observation by HESS [3] of a TeV γ ray source near the location of Sagittarius A∗, predictions for neutron fluxes at Auger energies have been published [4]. Since neutrons emitted by such a source would go undeflected by galactic magnetic fields, they should appear as a point-like source, just spread by the angular resolution of the experiment (the neutron decay length becomes comparable to the distance to the GC at EeV energies). 2. Data Selection In this work we use Auger data from 1st January 2004 until 6th June 2005. We use the events from the surface detector (SD) [5] that passed the 3-fold or the 4-fold data acquisition triggers and satisfying our high level physics trigger (T4) and our quality trigger (T5) [6]. The T5 selection is independent of energy and ensures a better quality for the event reconstruction. This data set has an angular resolution better than 2.2◦ for all of the 3-fold events (regardless of the zenith angle considered) and better than 1.7◦ for all events with multiplicities > 3 SD stations [7]. In all our analyses we use a zenith angle cut at 60◦ like AGASA while SUGAR used all zenith angles. One concern about the use of Auger data at 1 EeV given the 1.5 km spacing of our SD stations could be the trigger efficiency. Various studies (using Monte Carlo simulation, our Hybrid data set, or the extrapolation of the spectral shape) have shown that our 3-fold trigger efficiency is better than 30% for proton induced showers 2 Pierre Auger Collaboration Data Set >0.1 EeV [0.8-3.2] EeV [1.0 - 5.0]EeV >5.0 EeV >10.0 EeV SD 122636 41792 Table 1. Events statistics 29773 1359 387 and better than 50% for Iron induced shower above 0.8 EeV. The 4-fold trigger reaches the same efficiency at about 2 EeV [8]. The efficiency ratio between Iron and proton is always below 1.6 above 0.8 EeV and this difference can be taken into account when setting upper limits on point sources. Regarding the hybrid events (i.e. those with both Fluorescence Detectors (FD) and SD signal) [9], despite the lower duty cycle of the fluorescence telescopes and consequently smaller statistics, they offer an excellent angular resolution of 0.4◦ [7] over the whole energy range and a much lower trigger threshold of 0.1 EeV. In Table 1 we show the statistics for the SD data-set in the various energy ranges we have used for our studies. This is the first search in the southern hemisphere since the SUGAR analysis and we have over 10 times more statistics. 3. AGASA and SUGAR Excesses To estimate the coverage map, needed to construct excess and excess probability maps, we tested two different techniques. The so-called "shuffling" technique where one uses the data to construct randomized isotropic data sets and, alternatively for the SD sample, a semi-analytical method described in [10]. Both techniques are equivalent for studying sources because the Poisson noise in our search window dominates over the uncertainty of our coverage map. This report is based on the shuffling technique. For the hybrid sample, we have defined 24 epochs each corresponding to a given telescope configuration. For the SD sample we have randomized, within 5 bands of zenith angle, the UTC hours and Julian days of the events and drawn the azimuth from a uniform distribution. We present in Fig. 1A the coverage map obtained from our SD sample in a region around the GC. Since we are well within the field of view of Auger there are no strong variation within this region. In Fig.1B,C,D we present the chance probability distributions (mapped to positive Gaussian significance for excesses and negative for deficits) in the same region for various filtering and energy cuts corresponding to our various searches. In these map the filtering is choosen as to maximize the appearance of eventual structures at a certain scale. 1.5◦ corresponds to our 2.2◦ angular resolution and therefore to point sources, 3.7◦ is similar to a 5◦ top-hat window and also corresponds to the SUGAR excess size, 13.3◦ is similar to a 20◦ top-hat and to the size of the excess reported by AGASA. In these maps the chance probability distributions are consistent with those expected as a result of statistical fluctuations from an isotropic sky. Regarding the region where the AGASA excess was reported, the results from the Auger Observatory are 1155 events observed, and 1160.7 expected (ratio 1.00±0.03) for the energy range [1.0-2.5] EeV. This is almost 3 times the event number of AGASA in this region due to the fact that the GC lies well within the field of view of Auger while it lies outside the one of AGASA. These results do not support the excess observed by AGASA, and in particular not at a level of 22% like the one they reported which would translate into a 7.5σ excess. In a worst case scenario where the source would be protons and the background much heavier (e.g. Iron), the difference in detection efficiency of the Auger trigger at 1 EeV would reduce the sensitivity to a source excess. However, using the Fe/proton efficiency ratio at 1 EeV (70%/50% = 1.44, an upper bound in the range [1-2.5] EeV) we would still expect to see a 5.2σ event excess in our data set. There may be systematic differences between the energy calibrations of the two experiments, so we have scanned the region maintaining the high/low energy ratio and moving the interval center by ±40%. We have also enlarged the interval to [0.8 -- 3.2] EeV. We found no significant excess in any of those cases. Regarding the Galactic Center Studies 3 Figure 1. Lambert projections of the galactic centre region, GC (cross), galactic plane (solid line), regions of excess of AGASA and SUGAR (circles), AGASA f.o.v. limit (dashed line). A) coverage map (same color scale as the significance maps, but in a range [0-1.0]). B) significance map in the range [0.8-3.2] EeV smoothed using the individual pointing resolution of the events and a 1.5◦ filter (Auger like excess), C) same smoothed at 3.7◦ (SUGAR like excess), D) in the range [1.0-2.5] EeV smoothed at 13.3◦ (AGASA like excess). excess claimed by SUGAR, we find in their angular/energy window 144 events observed, and 150.9 expected (ratio 0.95±0.08) , and hence with over an order of magnitude more statistics we are not able to confirm this claim. 4. Galactic Center We then searched for signals of a point-like source in the direction of the GC. Using a 1.5◦ Gaussian filter corresponding to the angular resolution of the SD [7]. In the energy range [0.8 -- 3.2] EeV, we obtain 24.3 events observed and, 23.9 expected (ratio 1.0±0.1). A 95% CL upper bound on the number of events coming from a point source in that window is ns(95%) = 6.3. This bound can be translated into a flux upper limit if we know how many events (ns) are expected for a given flux (Φs) integrated in this energy range. Since the detector efficiency is energy (and primary composition) dependent at EeV energies, the ratio Φs/ns will depend on the spectral shape and nature of the source. In the simplest case in which the source has a spectrum similar to the one of the overall CR spectrum (dN/dE ∝ E−3), we can relate the two with Φs = nsΦCR4πσ2/nexp where σ is the size of the Gaussian filter used. Using ΦCR(E) = 1.5 ξ(E/EeV )−3 × 10−12 (EeV−1 m−2 s−1 sr−1) where ξ ∈ [1, 2.5] denotes our uncertainty on the CR flux (ξ is around unity for Auger and 2.5 for AGASA), introducing ε the Iron/proton detection efficiency ratio (1 < ε < 1.6 for E ∈ [0.8, 3.2] EeV) and, integrating in that energy range we obtain : Φs < 2.5 ξ ε × 10−15 m−2s−1 @ 95% CL. Since the detector efficiency grows with energy, a harder source flux would lead to a stronger flux bound. In a worst case scenario, where both ξ and ε take their maximum value, the bound is Φs = 10.0 × 10−15 m−2s−1, and still excludes the neutron source scenario suggested in [1, 11] to account for the AGASA excess, or in [4] in connection with the HESS measurements. 4 Pierre Auger Collaboration Figure 2. Left, the GC region seen by the hybrid detector, the excess map is built using the individual pointing resolution of the events. Right, significance map from the SD data in galactic coordinates for events with 1 EeV < E < 5 EeV (top), in super-galactic coordinates for events with E>5 EeV (middle), idem for events with E>10 EeV (bottom). Using the better angular resolution of the hybrid reconstruction, we can set a point-like flux limit inside a 1◦ radius cone around the direction of the galactic center. Due to the limited statistics, all hybrid events with energies above 0.1 EeV (10589) are used in this analysis, we find in this window 4 events observed and, 3.4 expected , showing no significant excess. At 95% CL an upper limit for a neutron source at the GC above 0.1 EeV is: Φs < 1.2 ξ × 10−13m−2s−1 @ 95% CL. 5. Galactic/Super-Galactic Plane Studies It is expected that the origin of cosmic rays changes from galactic to extra galactic in the 1-10 EeV range. We have looked for an excess of events inside a ±10◦ band along the Galactic Plane in the energy range from 1 EeV to 5 EeV and along the Super Galactic Planes above 5 and 10 EeV. We observed 5077, 229, and 68 events respectively for expectations of 5083.3, 235.6, and 67.4 (ratios 1.00±0.01, 0.97±0.07, and 1.0±0.1) showing no significant deviations. On Fig. 3 we show our corresponding probability maps smoothed on a 10◦ scale. Again all chance probabilities are consistent with isotropy. References [1] N. Hayashida et al. (AGASA Collaboration) ICRC 1999, Salt Lake City, OG.1.3.04, [astro-ph/9906056]. [2] J. A. Bellido et al., Astropart. Phys. 15, 167 (2001) [astro-ph/0009039]. [3] F. Aharonian et al. (HESS Collaboration), [astro-ph/0408145]. [4] F.Aharonian and A.Neronov [astro-ph/0408303] [5] Auger Collaboration, "Performances of the Pierre Auger Observatory Surface Array," these proceedings. [6] Auger Collaboration, "Triggers of the Pierre Auger Observatory Surface Array," these proceedings. [7] Auger Collaboration, "Angular Resolution of the Pierre Auger Observatory ," these proceedings. [8] Auger Collaboration, "Acceptance of the Pierre Auger Observatory Surface Array," these proceedings. [9] Auger Collaboration, "The Hybrid Performance of the Pierre Auger Observatory," these proceedings. [10] Auger Collaboration, "Coverage and large scale anisotropies estimation methods...," these proceedings. [11] M. Bossa et al., J. Phys G 29 (2003) 1409.
astro-ph/9509001
1
9509
1995-09-01T10:02:06
Colour Gradients in the Optical and Near-IR
[ "astro-ph" ]
For many years broadband colours have been used to obtain insight into the contents of galaxies, in particular to estimate stellar and dust content. Broadband colours are easy to obtain for large samples of objects, making them ideal for statistical studies. In this paper I use the radial distribution of the colours in galaxies, which gives more insight into the local processes driving the global colour differences than integrated colours. Almost all galaxies in my sample of 86 face-on galaxies become systematically bluer with increasing radius. The radial photometry is compared to new dust extinction models and stellar population synthesis models. This comparison shows that the colour gradients in face-on galaxies are best explained by age and metallicity gradients in the stellar populations and that dust reddening plays a minor role. The colour gradients imply $M/L$ gradients, making the `missing light' problem as derived from rotation curve fitting even worse.
astro-ph
astro-ph
Colour Gradients in the Optical and Near-IR 5 9 9 1 p e S 1 1 v 1 0 0 9 0 5 9 / h p - o r t s a : v i X r a Roelof S. de Jong12 1 Univ. of Durham, Dept. of Physics, South Road, Durham DH1 3LE, UK 2 Kapteyn Institute, P.O.box 800, 9700 AV Groningen, The Netherlands Abstract. For many years broadband colours have been used to obtain insight into the contents of galaxies, in particular to estimate stellar and dust content. Broad- band colours are easy to obtain for large samples of objects, making them ideal for statistical studies. In this paper I use the radial distribution of the colours in galaxies, which gives more insight into the local processes driving the global colour differences than integrated colours. Almost all galaxies in my sample of 86 face-on galaxies become systematically bluer with increasing radius. The radial photometry is compared to new dust extinction models and stellar population synthesis mod- els. This comparison shows that the colour gradients in face-on galaxies are best explained by age and metallicity gradients in the stellar populations and that dust reddening plays a minor role. The colour gradients imply M/L gradients, making the 'missing light' problem as derived from rotation curve fitting even worse. 1 The colour gradients A sample of 86 spiral galaxies was imaged in the B, V , R, I, H and K passbands to study light and colour distributions as a function of radius. Full details of sample selection and data reduction are described in de Jong & van der Kruit (1994). The galaxies were selected to be face-on and to have a diameter of at least 2′. The sample is statistically complete and can be corrected for selection effects. It can therefore be used to analyze the nature of the Freeman law (Freeman 1970) and this analysis has been reported elsewhere (de Jong 1995a, 1995b). The luminosity profiles were determined in the usual way by measuring the average surface brightness on annuli of increasing radius. Radial colour profiles were created by combining profiles in different passbands. The run of colour as function of radius is put on a common scale for all galaxies in Fig. 1, where the average B -- K colour at each radius is plotted as function of the average R surface brightness at this radius. 2 Roelof S. de Jong Fig. 1. The average B -- K colour at each radius as function of the average R surface brightness at this radius. The galaxies are divided into 4 morphological RC3 T-type bins. The dashed lines are provided to have a common reference among the bins. Two observations can readily be made from this diagram. Firstly, all galaxies become bluer going radially outward, correlating strongly with the average surface brightness at each radius. Secondly, even at the same sur- face brightness, late type spiral galaxies are bluer than earlier types (use the dashed lines to guide the eye). Furthermore it should be noted that there is a smooth transition in colour from the bulge to the disk region. The colours of bulges are nearly identical to the colours of inner disk regions (see also Peletier, these proceedings). 2 The dust and stellar population synthesis models A possible explanation for the colour gradients is reddening due to dust extinction. As galaxies are intricate mixtures of stars and dust, we cannot describe the reddening by a simple extinction law, but have to calculate the separate contributions of absorption and scattering. To predict reddening profiles due to absorption and scattering full 3D Monte Carlo simulations were made (de Jong 1995a) of galaxies with smooth exponential dust and stellar distributions in both radial and vertical directions. The main free parameters are the dust to stellar scaleheight (zd/zs) and scalelength (hd/hs) ratios, the central optical depth (τ0,V ) and the properties of the dust particles. The (poorly determined) Galactic dust properties were assumed, since extra- galactic albedo and phase scattering functions have never been measured. Colour Gradients in the Optical and Near-IR 3 Fig. 2. The radial colour-colour reddening profiles resulting from the Monte Carlo simulations of stars and dust in exponential disks for different zd/zs and τ0,V values. The arbitrary unreddened colours are indicated with filled circles for each model, the galaxy centres indicated by open circles. Figure 2 shows a number of model colour-colour reddening profiles for different zd/zs and τ0,V values. The positions of the models in the diagram are arbitrary (depending on the colours of the underlying population), but the shape of the reddening profiles are determined by the distribution and the properties of the dust. Note that all reddening profiles point in about the same direction, independent of the dust configuration, and that this direction is different from the standard screen model extinction vector (arrow). I have used two sets of stellar population synthesis models in the com- parison with the data. The Solar metallicity models of Bruzual & Charlot (1993) are used to study the colour changes of populations due to different star formation histories. Two extreme cases are considered, a single star burst model and a constant star formation model. The colours of these populations are inspected after 8 and 17 Gyr. The models of Worthey (1994) are used to study the effects that age and metallicity have on the colours of a population. 3 The comparison between models and data Figure 3 shows the colour-colour diagrams of the models and the galaxies, again divided into four T-type bins. Note that the same model should fit the data in all colour combinations, thus in both Figs. 3a & 3b. The thin lines represent the galaxy data; the central galaxy colours are indicated by the open circles, the lines show the run of colours as function of radius. All galaxies with type T<6 are confined to a small region in these diagrams, only the later types show a considerable larger spread. 4 Roelof S. de Jong Fig. 3a. The run of B -- V versus R -- K colour as function of radius for the galaxies (thin lines, centre indicated by open circle). Dust models in the top left corner, Bruzual & Charlot (1993) models after 8 and 17 Gyr thick dashed lines, Worthey (1994) models for indicated metallicities thick solid lines in the two bottom panels. Some dust model profiles are indicated in the top-left corner of the panels. As mentioned in Sect. 2, these profiles can be placed anywhere in the diagrams and their direction depends mainly on the dust properties, not on the relative distribution of dust and stars. Clearly, the colour gradients cannot be caused by reddening alone, assuming that the dust properties used are correct. In the 5≤T<6 panels the colours predicted by Worthey's models after 12 Gyr are shown for a range of indicated metallicities. The metallicity-colour trend runs in the same direction for other ages and apparently, a metallicity gradient alone cannot explain the observed colour gradients. The effects of different star formation histories are indicated by the two dashed lines in the centre of the panels. The red ends of these lines indicate the colours of a single burst population, the blue end of a population of constant star formation rate. Both the position and the direction of these solar metallicity models seem to agree reasonably well with the data, and the most simple explanation for the colour gradients would be age gradients across the disks of spiral galaxies. Still this cannot be the whole story, as galaxies still have star formation in their central regions, which means that a single burst is a bad approximation. Furthermore it is known from measurement of HII regions that spiral galaxies have metallicity gradients in their gas content, Colour Gradients in the Optical and Near-IR 5 Fig. 3b. As Fig. 3a, but for B -- I versus V -- K. which most likely is partly reflected in the stellar component. So the most consistent picture is one where colour gradients are caused by both age and metallicity gradients, with the central regions of galaxies being on average quite old and having a range of metallicities, whereas the outer parts are young and have low metallicities. The large spread in colours of the late type galaxies can be explained by stellar population changes as well. The single-burst age evolution is indicated by the thick, solid lines in the 6≤T≤10 panels, for the metallicities indicated in the 5≤T<6 panels. The colours indicate that the stellar population in some of these galaxies are on average very young and have a low metallicity. The colour gradients imply large M/L gradients for the optical passbands, making the 'missing light' problem as derived from rotation curve fitting even worse, irrespective whether they are caused by dust or population changes. References Bruzual G.A., Charlot S. 1993, ApJ 405, 538 de Jong R.S. 1995a, Ph.D. Thesis, Univ. of Groningen, The Netherlands de Jong R.S. 1995b, A&A, submitted de Jong R.S., van der Kruit P.C. 1994, A&AS 106, 451 Freeman K.C. 1970, ApJ 160, 811 Worthey G. 1994, ApJS 95, 107
astro-ph/9906373
1
9906
1999-06-24T06:47:19
High-Energy Gamma-Ray Observations of Two Young, Energetic Radio Pulsars
[ "astro-ph" ]
We present results of Compton Gamma-Ray Observatory EGRET observations of the unidentified high-energy gamma-ray sources 2EG J1049-5847 (GEV J1047-5840, 3EG J1048-5840) and 2EG J1103-6106 (3EG J1102-6103). These sources are spatially coincident with the young, energetic radio pulsars PSRs B1046-58 and J1105-6107, respectively. We find evidence for an association between PSR B1046-58 and 2EG J1049-5847. The gamma-ray pulse profile, obtained by folding time-tagged photons having energies above 400 MeV using contemporaneous radio ephemerides, has probability of arising by chance of 1.2E-4 according to the binning-independent H-test. A spatial analysis of the on-pulse photons reveals a point source of equivalent significance 10.2 sigma. Off-pulse, the significance drops to 5.8 sigma. Archival ASCA data show that the only hard X-ray point source in the 95% confidence error box of the gamma-ray source is spatially coincident with the pulsar within the 1' uncertainty (Pivovaroff, Kaspi & Gotthelf 1999). The double peaked gamma-ray pulse morphology and leading radio pulse are similar to those seen for other gamma-ray pulsars and are well-explained in models in which the gamma-ray emission is produced in charge-depleted gaps in the outer magnetosphere. The inferred pulsed gamma-ray flux above 400 MeV, (2.5 +/- 0.6) x 10E-10 erg/cm^2/s, represents 0.011 +/- 0.003 of the pulsar's spin-down luminosity, for a distance of 3 kpc and 1 sr beaming. For PSR J1105-6107, light curves obtained by folding EGRET photons using contemporaneous radio ephemerides show no significant features. We conclude that this pulsar converts less than 0.014 of its spin-down luminosity into E > 100 MeV gamma-rays beaming in our direction (99% confidence), assuming a distance of 7 kpc, 1 sr beaming and a duty cycle of 0.5.
astro-ph
astro-ph
High-Energy Gamma-Ray Observations of Two Young, Energetic Radio Pulsars V. M. Kaspi1 and J. R. Lackey2 Department of Physics and Center for Space Research, Massachusetts Institute of Technology, 70 Vassar St., Cambridge, MA 02139 Astronomy Department, Boston University, 725 Commonwealth Ave., Boston, MA 02215 J. Mattox3 Australia Telescope National Facility, Epping, NSW, 2121 Australia R. N. Manchester4 Astrophysics and Supercomputing, Swinburne University of Technology, Mail 31, PO Box M. Bailes5 218, Hawthorn, Victoria, 3122 Australia R. Pace6 Physics Department, University of Adelaide, Adelaide, SA, Australia ABSTRACT results We present of Compton Gamma-Ray Observatory/EGRET observations of the unidentified high-energy γ-ray sources 2EG J1049−5847 (GEV J1047−5840, 3EG J1048−5840) and 2EG J1103−6106 (3EG J1102−6103). These sources are spatially coincident with the young, energetic radio pulsars PSRs B1046−58 and J1105−6107, respectively. We find evidence for an association between PSR B1046−58 and 2EG J1049−5847. The γ-ray pulse profile, obtained by folding time-tagged photons having energies above 400 MeV using contem- poraneous radio ephemerides, has probability of arising by chance of 1.2 × 10−4 according to the binning-independent H-test. A spatial analysis of the on-pulse 9 9 9 1 n u J 4 2 1 v 3 7 3 6 0 9 9 / h p - o r t s a : v i X r a 1Alfred P. Sloan Research Fellow; [email protected] [email protected] [email protected] [email protected] [email protected] – 2 – photons reveals a point source of equivalent significance 10.2σ. Off-pulse, the sig- nificance drops to 5.8σ. Archival ASCA data show that the only hard X-ray point source in the 95% confidence error box of the γ-ray source is spatially coincident with the pulsar within the 1′ uncertainty (Pivovaroff, Kaspi, & Gotthelf 1999). The double peaked γ-ray pulse morphology and leading radio pulse are similar to those seen for other γ-ray pulsars and are well-explained in models in which the γ-ray emission is produced in charge-depleted gaps in the outer magnetosphere. The inferred pulsed γ-ray flux above 400 MeV, (2.5 ± 0.6) × 10−10 erg cm−2 s−1, represents 0.011±0.003 of the pulsar's spin-down luminosity, for a distance of 3 kpc and 1 sr beaming. For PSR J1105−6107, light curves obtained by folding EGRET photons using contemporaneous radio ephemerides show no significant features. We conclude that this pulsar converts less than 0.014 of its spin-down luminosity into E > 100 MeV γ-rays beaming in our direction (99% confidence), assuming a distance of 7 kpc, 1 sr beaming and a duty cycle of 0.5. Subject headings: gamma rays: observations - pulsars: general - pulsars: in- dividual: PSR B1046−58, PSR J1105−6107 1. Introduction An outstanding unknown in γ-ray astronomy is the nature of a population of high-energy γ-ray sources at low Galactic latitude which were first identified using the COS-B satellite (Swanenburg et al. 1981). More recently, these sources were studied in greater detail by the Energetic Gamma Ray Experiment Telescope (EGRET) aboard the Compton Gamma-Ray Observatory (Kanbach et al. 1996). In contrast to the high-energy γ-ray sources at high latitudes, which are typically associated with active galaxies, several dozen apparent point sources of high-energy γ-rays at latitudes b < ∼ 10◦ have as yet not been identified with astronomical objects. However, a handful of low-latitude sources has been identified with young, energetic radio pulsars (see Thompson 1996 for a review). This suggests that many of the unidentified sources are also young pulsars. This hypothesis has been considered in various studies, many of which suggest that the majority, if not all, of the unidentified sources are pulsars (Yadigaroglu & Romani 1995, Kaaret & Cottam 1996, Yadigaroglu & Romani 1997, Cheng & Zhang 1998). Sturner & Dermer (1995) suggested that many of the sources are not pulsars but rather are a result of cosmic ray interactions with gas in supernova remnants. With the EGRET spatial resolution generally limited to tens of arcminutes, identification of discrete high-energy γ-ray sources is difficult using spatial coincidence alone, explaining why this problem has remained open so long. – 3 – Rotation-powered pulsars permit a straightforward determination of the nature of a γ- ray source using the signature pulsed emission. Indeed the Crab and Vela pulsars are long- known to be strong sources of pulsed γ-rays. EGRET has provided unambiguous and/or strong evidence for pulsations from five other rotation-powered pulsars (see Table 1). Five E/d2, where the spin-down of the six pulsars having the highest values of the parameter luminosity E ≡ 4π2I P /P 3 (P is the pulse period and P its rate of increase) and d is the distance to the source, are established sources of high-energy γ-ray pulsations. Finding rotation-powered pulsars at γ-ray energies is important because those detected thus far have their maximum luminosities in this energy range. Thus γ-rays may hold the key to the poorly understood pulsar energy budget and emission mechanism, in which mechanical energy of rotation is efficiently converted into non-thermal high-energy radiation. The origin of the γ-rays – whether at the neutron star polar cap (e.g. Daugherty & Harding 1996) or in the outer magnetosphere (e.g. Romani & Yadigaroglu 1995) – is still debated. In this paper we examine high-energy γ-ray data from the direction of two young, energetic pulsars which are good candidates for observable γ-ray pulsations. PSR B1046−58 is a 124 ms radio pulsar having characteristic age 20 kyr and spin-down luminosity E = 2.0 × 1036 erg s−1. It was discovered in a survey of the Galactic plane for radio pulsars (Johnston et al. 1992), and has an unremarkable radio average pulse profile, characterized by a single narrow peak having width ∼8 ms at 1.4 GHz (Figure 1). Its pulsed radio emission is highly linearly polarized and shows a slow position angle variation across the pulse, suggestive of conal emission (Qiao et al. 1995). The dispersion measure of 129.01±0.01 pc cm−3 toward this pulsar implies a distance of 2.98±0.35 kpc (Taylor & Cordes 1993). In the rank-ordered E/d2 list, PSR B1046−58 falls ninth, with six of the eight above it known γ-ray pulsars, five of those six being high-energy γ-ray sources (see Table 1). The pulsar's position lies near the 95% confidence contour for the Second EGRET catalog source 2EG J1049−5847 (Thompson et al. 1995). Thompson et al. (1994) reported a pulsed flux 99.9% confidence upper limit for energies above 100 MeV of 4.5×10−7 photons cm−2 s−1 for a 50% duty cycle pulse. They noted, however, a "hint" of a pulsation for PSR B1046−58 above 1 GeV, having probability of chance occurrence of 14%. As reported in a companion paper, Pivovaroff, Kaspi & Gotthelf (1999) have detected X-rays from the direction of the pulsar; that emission is not pulsed and is probably due to a synchrotron nebula. PSR J1105−6107 is a 63 ms pulsar having characteristic age 63 kyr and E = 2.5 × 1036 erg s−1 (Kaspi et al. 1997). It exhibits a narrow double peaked radio pulse profile having width 3.4 ms at 1.6 GHz. Its radio emission is highly linearly polarized, showing a slow positional angle variation suggestive of conal emission (Crawford et al. 1999). The pulsar's dispersion measure of 270.55±0.14 pc cm−3 implies a distance of 7.1±0.9 kpc (Taylor & Cordes 1993). Ranking by E/d2, it is 23rd, well above the known γ-ray pulsar PSR B1055−52 – 4 – (whose distance may well be overestimated; see Thompson et al. 1999 and references therein) though well below several sources which have not been detected by EGRET (Thompson et al. 1994). PSR J1105−6107 is spatially coincident with the unidentified EGRET source 2EG J1103−6106 (Thompson et al. 1995, Kaspi et al. 1997). This γ-ray source has pulsar-like properties: it has a hard spectrum and is not time variable. However the Second EGRET catalog noted marginal evidence for the source being extended. PSR J1105−6107 has been detected as an unpulsed X-ray source (Gotthelf & Kaspi 1998, Steinberger, Kaspi, & Gotthelf 1999); this emission is best explained as being due to nebular synchrotron emission powered by the pulsar wind. Given the importance of identifying γ-ray pulsations in radio pulsars as well as es- tablishing the nature of the unidentified EGRET sources, additional EGRET exposure of the Eta Carina region was obtained. We report here on an analysis of EGRET data for PSR B1046−58 obtained between 1991 May and 1997 October, and for PSR J1105−6107 between 1993 July and 1997 October. In §2 we summarize the EGRET observations used in our analysis. In §3 we report our findings, showing evidence for γ-ray pulsations from PSR B1046−58, but none from PSR J1105−6107. We discuss our results and their implica- tions in §4. 2. Observations and Analysis The Energetic Gamma Ray Experiment Telescope (EGRET) aboard the Compton Gamma Ray Observatory is sensitive to photons in the energy range from approximately 30 MeV to 30 GeV, using a multi-level spark chamber system to detect electron-positron pairs resulting from γ-rays. A NaI calorimeter provides energy resolution of 20-25%, and a plastic scintil- lator anti-coincidence dome prevents triggering on events not associated with high-energy photons. The detector thus has negligible background. The detector is described in more detail in Thompson et al. (1993) and references therein. The data sets used in this analysis consist of archival data, as well as PI data for Cycles 5 and 6. Data sets included event UTC arrival times accurate to 100 µs, energy, measured direction of origin, satellite location, and other information. A summary of the observations of PSRs B1046−58 and J1105−6107 is presented in Tables 2 and 3. – 5 – 2.1. Timing Analysis The timing analysis was done using the interactive software PULSAR which has been described elsewhere (e.g. Fierro 1995) and is based on the TEMPO software package (e.g. Taylor & Weisberg 1989). Event arrival times are first reduced to the solar system barycenter and then folded using a contemporaneous ephemeris derived from radio monitoring. Young pulsars are notorious for displaying timing irregularities, namely glitches and long-term timing "noise" (see Lyne 1996 for a review). The pulsars in question are no exception (Johnston et al. 1995, Kaspi et al. 1997). Timing noise alone can result in phase deviations as large as a full cycle over a year or two, while glitches can result in so large a phase deviation that only unusually dense observations can permit unambiguous phase determination. Such deviations cannot be tolerated when searching for γ-ray pulsations over many years, particularly when expecting weak pulsations, as even small deviations could broaden the signal beyond detectability. For this reason, we have used contemporaneous radio ephemerides determined by timing observations made at the 64-m radio telescope at Parkes, Australia. Most radio observations were done at a central frequency near 1500 MHz, with some at 430 and 660 MHz. Prior to 1995, all data were obtained using filter-bank timing systems (2 × 64 × 5 MHz at 1520 MHz and 2 × 256 × 0.125 MHz at 430 and 660 MHz) that have been described elsewhere (e.g. Bailes et al. 1994). Most data from after 1995 were obtained using the Caltech correlator- based pulsar timing machine (Navarro 1994), which has 2× 128 lags across 128 MHz in each of two separate frequency bands. Typically, correlator observations were made at central fre- quencies of 1420 and 1650 MHz simultaneously. Filter-bank data were recorded on tape and folded off-line; correlator data were folded on-line. Resulting folded profiles were convolved with high signal-to-noise templates to yield topocentric pulse arrival times. Resulting arrival times were analyzed using the standard TEMPO pulsar timing software package6 together with the JPL DE200 ephemeris (Standish 1982). Ephemerides for folding the γ-ray data were de- termined piecewise using a minimum of some two dozen arrival times per valid interval, with RMS residuals well under the γ-ray binning resolution in all cases. The ephemerides used are provided for each EGRET viewing period in publicly accessible machine-readable format at http://www.atnf.csiro.au/research/pulsar/psr/archive/. Since the range of radio frequencies observed at each epoch was not generally sufficiently large to determine dispersion measures with high precision, we assume that the dispersion measures do not vary significantly. Substantial variations would result in changes in the 6http://pulsar.princeton.edu/tempo – 6 – pulse phase extrapolated from the radio ephemeris to the effectively infinite γ-ray frequency according to the cold plasma dispersion law. This could cause smearing of the pulse profile. We can test the plausibility of our assumption of a constant dispersion measure (DM) using the results of Backer et al. (1993) who empirically characterized variations in DM for sev- eral pulsars. As neither pulsar considered here lies within a detectable supernova remnant (Johnston et al. 1995, Kaspi et al. 1996, Gotthelf & Kaspi 1998, Pivovaroff, Kaspi, & Got- thelf 1999), the very large DM variations observed for the Crab and Vela pulsars, usually attributed to their unusual environments, are unlikely to be present. From the empirical findings of Backer et al. (1993), for PSR B1046−58, the maximum DM variation per year expected is ∼0.01 pc cm−3. Over the ∼8 yr duration of the experiment described here, this amounts to a variation in pulse phase of ∼1.2×10−3, a factor of ∼40 smaller than our binning resolution. For PSR J1105−6107, the DM variation could be an order of magnitude larger, however the experiment in this case spans only ∼4 yr. The phase variation then could be as large as ∼1.1×10−2, smaller, but only by a factor of ∼4, than our binning resolution. We conclude that DM variations are certainly negligible for PSR B1046−58, and probably negligible for PSR J1105−6107. 2.2. Spatial Analysis The energy dependence of the point-spread function of the EGRET detector can be characterized by the function θ67(E) = 5◦.85 × (E/100)−0.534, (1) where θ67 is the half-angle of the cone encircling the actual source direction and which con- tains 67% of the events having energy E in MeV (Thompson et al. 1993). In our timing analysis, we used all events within θ67. The source directions are (J2000) RA 10h 48m 12s.6(3), DEC −58◦ 32′ 03′′.75(1) for PSR B1046−58 (Stappers et al. 1999), and (J2000) RA 11h 05m 26s.07(7), DEC −61◦ 07′ 52′′.1(4) for PSR J1105−6107 (Kaspi et al. 1997), where the numbers in brackets are 1σ uncertainties. The total number of counts folded is given for each viewing period in Tables 2 and 3 for PSRs B1046−58 and J1105−6107, respectively. Spatial analysis of γ-ray data is done using the likelihood ratio test (Mattox et al. 1996). The significance of a point source is found using a test statistic Ts that compares the likelihood of the distribution of counts (which is governed by Poisson statistics) having occurred under the null hypothesis of no point source, with that with a point source present, given the instrumental point-spread function (PSF). Mattox et al. have shown that a point- – 7 – source significance in units of the standard deviation is given by √Ts where Ts ≡ 2(ln L1 − ln L0), (2) where L1 is the likelihood under the assumption of a point source and L0 is that under the null hypothesis. The likelihood itself is given by L = Y ij θnij ij e−θij nij! , (3) where the product is over pixel ij that contains nij counts, and the model prediction is θij. 3. Results 3.1. PSR B1046−58 3.1.1. Timing Analysis In Figure 1, we present the folded light curve for PSR B1046−58 for energies above 400 MeV. The significance of the apparent modulation can only be established using ap- propriate statistical tests. The optimal test is the H-test (de Jager 1994), which provides a figure of merit for a folded light curve that is independent of the number of bins or of the phase of the putative signal. For this reason, the H-test is superior to the standard χ2 test (e.g. Leahy et al. 1983). Further, the H statistic has been shown to be optimal when con- sidering a light curve having unknown pulse morphology. In this sense, it is superior to the Z 2 N test (Buccheri et al. 1983) which can be done for specified numbers of harmonics only. For the light curve in Figure 1, the H statistic is 22.5, which has a probability of 1.2 × 10−4 of having occurred by random chance, with 5 harmonics indicated. This corresponds to an equivalent normal distribution deviation of 3.9σ from the null hypothesis of no pulsations. For reference, χ2 = 26.3 for 10 bins (9 degrees of freedom); the probability of this value or higher having occurred by chance is 1.8 × 10−3 (3.1σ). The Z 2 4 statistic has value 33.8; the probability of this value or higher having occurred by chance is 4.3 × 10−5 (4.1σ). Note that there is effectively only a single trial folding period. Since we considered two pulsars, these probabilities should be multiplied by a factor of two. Although we did not search independent energy bands, the choice of 400 MeV lower energy cutoff was made to optimize the significance; we discuss this further below. We have checked the robustness of our result in several ways. First, we divided the total data set into two halves of approximately equal size and verified that the signal is seen in – 8 – both, albeit at reduced significance, as expected. For the first half data set (MJDs 48386– 49361), we find H=16.7, which has probability 1.3 × 10−3, while for the second half data set (MJDs 49361–50728) we found H=15.0, having probability 2.4 × 10−3. These statistics are for E > 400 MeV. Second, we verified that the signal persisted when the energy-dependent cone opening angle (Eq. 1) was increased to include 99% of the counts from the known source direction; in this case we found H=12.5, which has probability 6.8 × 10−3. Note that zero is suppressed in the light curve shown in Figure 1. The observed pulsed fraction for E > 400 MeV is 0.12±0.02. The off-pulse γ-rays may originate as diffuse Galactic emission - the pulsar is located in the complicated Eta Carina region, tangent to the Carina spiral arm, which is known to contain significant diffuse emission and many discrete sources. In §3.1.2 we describe a spatial analysis that (Bertsch et al. 1993, Hunter et al. 1997). indicates that a small fraction of the off-pulse could be emission from PSR B1046−58 itself. One way to check for the presence of contaminating diffuse emission is to utilize the different expected spectra of diffuse and pulsed signals. In Figure 2 we show, in the lower panel, results of three statistical tests as a function of lower cutoff energy. The probability shown is that of finding the value of the statistical test, or higher, for a profile consisting of folded photons having energies above the cutoff. The upper panel shows the number of photons included in the analysis, here done for counts within θ67. Qualitatively, the behavior is as expected: the diffuse emission has a softer spectrum than those of the known γ-ray pulsars, so as the lower energy bound is increased, the significance of the pulse should increase until the pulsar signal is too photon-starved to be detected. Furthermore, the EGRET PSF narrows with increasing energy (Eq. 1), so diffuse emission and nearby confusing point sources are preferentially removed as the lower energy bound is increased given advance knowledge of the pulsar position (Lamb & Macomb 1997). However, the sudden decrease in significance when the lower energy cut changes from 400 to 600 MeV is unexpected, given the smooth decrease in the number of counts. This could indicate a spectral cutoff in the pulsed emission, or could simply be a result of the poor statistics for the pulse profile; it could also suggest the apparently significant modulation is a statistical fluctuation. The detection of pulsed high-energy γ-ray emission from PSR B1951+32 by Ramana- murthy et al. (1995) was done by considering events selected from a cone of fixed angle with respect to the pulsar position, as a compromise between maximizing signal from the source and minimizing background contamination. We have carried out the same analysis for PSR B1046−58, using a fixed cone of 2◦ half-angle. We chose this size cone as it resulted in a comparable total number of events as for the energy-dependent cone. For the fixed cone, we find H= 23.9, corresponding to a probability of 6.9 × 10−5 (4.0σ), an improvement over the energy-dependent-cone results. This supports the reality of the pulsations. However, – 9 – not all the profile statistics improve for the fixed cone; for example, the χ2 statistic drops to 25.3, corresponding to a probability of 2.7 × 10−3 (3.0σ), still statistically significant, but less so than with the energy-dependent cone. Ramanamurthy et al. (1996) reported high-energy γ-ray pulsations from the radio pulsar PSR B0656+14 using photons that had been "weighted" according to their energy and direction, and the known telescope response. We have employed a similar algorithm to theirs in considering the apparent modulation seen in the unweighted data. Each event is assigned a weight which is equal to the probability that a specific γ-ray originated from the point source (rather than as diffuse Galactic emission), w = PSF(r, E) PSF(r, E) + B S (E) , (4) where PSF(r, E) is the energy-dependent EGRET point spread function, the probability per steradian for a photon of energy E to be detected at a measured angle r from the source. B/S is the ratio of diffuse Galactic counts per steradian to source counts at energy E. As the combinations of weights within phase bins are, unlike the unweighted data, not described by Poisson statistics, the same statistical tests do not apply. Instead, we considered the simple statistical standard deviation of the binned values, xi, PN i=1(xi−x)2/(N−1). We used Monte Carlo simulations of profiles to determine the probability of chance occurrence of a profile having this statistic. Simulated profiles of weighted data were made by randomly varying the phases. In this way, for E > 400 MeV, we found that the probability of obtaining a profile having the same modified χ2 statistic or higher for the folded, weighted data was 0.026. Thus the significance of the detection decreases using the weights. This was not expected. However, given the evidence below in favor of the association, it does not eliminate the possibility that pulsed flux is being detected from PSR B1046−58; the decrease in significance could be due to the poorly known pulsed γ-ray spectrum (which was assumed to be a power law of index 1.97±0.07 – see §4.1.2) or to a statistical fluctuation. 3.1.2. Spatial Analysis We have conducted a likelihood analysis (Mattox et al. 1996) of the EGRET data for PSR B1046−58 (see §2.2) for E > 400 MeV events. In the analysis, the model for the diffuse Galactic flux of Hunter et al. (1997) was used, and the flux of nearby sources in the Second EGRET catalog (Thompson et al. 1995) was estimated simultaneously with that from the position of PSR B1046−58. A point source at the pulsar position was detected with a significance of √Ts = 10.8σ. We estimate 347±40 γ-ray events from this point source, corresponding to a flux of (1.6 ± 0.2)×10−7 photons cm−2s−1 (E > 400 MeV). The – 10 – distribution of EGRET events is consistent with the EGRET PSF; we find a reduced χ2 of 1.3 for the sum of all events for Cycles 1–6 of the EGRET mission. A similar analysis was done for different rotational phases. Phase-resolved event maps for E > 400 MeV were made for peak 1 (phases 0.300–0.475), interpeak 1 (0.475–0.675), peak 2 (0.675–0.825), and interpeak 2 (0.825–0.300). The peak 1 and peak 2 maps were combined to form an "on-pulse" map. The interpeak 1 and interpeak 2 maps were combined to form an "off-pulse" map. For the on-pulse map, a point source at the position of PSR B1046−58 was detected with a significance of √Ts = 10.2σ, corresponding to an estimated 203±25 γ-ray events. For the off-pulse map, a point source at the position of PSR B1046−58 was detected with a significance of √Ts = 5.8σ corresponding to an estimated 144±29 γ-ray events. The distri- bution of EGRET events was consistent with the PSF for both the on- and off-pulse phase selections. The EGRET position estimate was also consistent with that of PSR B1046−58 for both the on- and off-pulse phase selections. Thus, significant flux is detected during both the on- and off-pulse phases, and it appears that both the pulsar and diffuse Galactic γ-ray emission contribute to the emission represented by the suppressed zero in Figure 1. We also subdivided the on-pulse map into two, by making separate maps for peak 1 and peak 2 data. Both peak 1 and peak 2 show a point source at the position of the pulsar, with significances of 6.0σ and 8.4σ, respectively. Maps of "residual" flux were also made for both the on- and off-pulse phase selections by subtracting the counts distribution for nearby sources, counts corresponding to an isotropic flux of 0.295 × 10−5 photons cm−2 s−1 sr−1, and counts corresponding to diffuse Galactic flux (with an empirically determined scaling factor of 0.242 for the model of Hunter et al. 1997 for E > 100 MeV flux). The residual counts map was then converted to a residual flux map by dividing by the exposure. The difference between the residual fluxes for the on- and off-pulse maps shows the expected distribution for the EGRET PSF with a source coincident with PSR B1046−58. 3.2. PSR J1105−6107 3.2.1. Timing Analysis In Figure 3, we show the folded light curve for PSR J1105−6107 for energies above 400 MeV. The H statistic is 3.6, which has a probability of 0.24 of having occurred by chance. We thus find no evidence for γ-ray pulsations from PSR J1105−6107. We also tried folding – 11 – photons from within a cone encircling the source direction that contains 99% of the events; again no evidence for pulsations was found. We also searched for pulsations in different energy ranges to take advantage of the expected increased pulsed fraction with increasing low energy cutoff, but found no significant signals. Finally, we considered data weighted according to the algorithm described in §3.1, but again found no significant pulsations. The 99% confidence level upper limit on the pulsed fraction of 2EG J1103−6106 at the PSR J1105−6107 spin period for E > 400 MeV, using the H test and assuming a duty cycle of 0.5 is 0.33. For a duty cycle of 0.1, it is 0.13. For E > 100 MeV photons, the analogous limits are 0.09 and 0.04 for duty cycles 0.5 and 0.1, respectively. 4. Discussion 4.1. PSR B1046−58 The folded γ-ray profile for PSR B1046−58 and the spatial analysis of the γ-ray data provide evidence for the association between PSR B1046−58 and 2EG J1049−5847. However additional support for this conclusion is desirable, because of the difficulty in interpreting the pulse significance as a function of energy and the drop in significance when photon weighting is used. We consider here evidence that provides additional support for the association between PSR B1046−58 and 2EG J1049−5847. 4.1.1. ASCA X-Ray Observations of the 2EG J1049−5847 Field Additional evidence in favor of associating PSR B1046−58 with 2EG J1049−5847 are the results of ASCA X-ray observations of the 2EG J1049−5847 field, which are described in a companion paper (Pivovaroff, Kaspi, & Gotthelf 1999). There, archival ASCA ob- servations are presented that cover the full 95% confidence error circle of the counterpart (1999) in the recently published to 2EG J1049−5847 as determined by Hartman et al. show, 3EG catalog (where the source name is 3EG J1048−5840). As Pivovaroff et al. the only X-ray source within the 95% confidence contour of the γ-ray source is coincident with PSR B1046−58 within the < 1′ X-ray positional uncertainty. The source is unpulsed and is likely to be a synchrotron nebula. However, there is an additional nearby hard- spectrum X-ray source outside the 95% confidence contour (but within the 99% contour) that could also be the counterpart, a possibility we cannot rule out without an improved γ-ray source position. Nevertheless, the most likely counterpart from the X-ray observations is PSR B1046−58. – 12 – 4.1.2. Spectrum and Energetics Merck et al. (1996) found that 2EG J1049−5847 had a spectral index of 2.0±0.1, and noted that the γ-ray source had properties consistent with the known γ-ray pulsar population, as well as the positional coincidence with PSR B1046−58. Bertsch et al. (in preparation) find a spectral index of 1.97±0.09 for the 3EG counterpart of 2EG J1049−5847, 3EG J1048−5840, and an integrated flux for E > 100 MeV of (62±7)×10−8 photon cm−2 s−1 (R. Hartman, personal communication). We have done an independent pulse-phase-resolved spectral analysis by considering the fluxes and spectra for peaks 1 and 2 separately, using γ-rays arriving during the off-peak 2 interval as a background estimate. We find spectra for both peaks that are consistent with that reported above as well as with each other. Using the results of the on-pulse spatial analysis (§3.1.2) together with the reported 3EG point source flux (Hartman et al. 1999) and spectrum (Bertsch et al. 1999), we derive a pulsed flux for E > 400 MeV of (2.5 ± 0.6) × 10−10 erg cm−2 s−1, which corresponds to an efficiency for conversion of spin-down luminosity into pulsed E > 400 MeV γ-rays of 0.011±0.03 for 1 sr beaming, assuming d = 3 kpc. Using the derived total pulsed fraction (§3.1.1), the total number of counts and the exposure yields consistent results. The observed flux is consistent with previously published upper limits (Thompson et al. 1994). 4.1.3. Absence of Variability McLaughlin et al. (1996) quantified the variability of EGRET sources and showed that known γ-ray pulsars are not variable, in contrast to all known γ-ray-loud active galaxies for which accurate γ-ray flux measurements have been made. They further analyzed the variabil- ity of the unidentified EGRET sources and concluded that 2EG J1049−5847 is non-variable, consistent with its identification as an energetic rotation-powered pulsar. An analysis of the larger data set used to produce the 3EG catalog reveals no evidence for variability from the source, consistent with its identification as a rotation-powered pulsar (Hartman et al. 1999). 4.1.4. Pulse Profile Morphology and Relative Radio Phase The morphology of the γ-ray profile shown in Figure 1 lends some support to its in- terpretation as pulsations from PSR B1046−58. Note that in the profile, the difference in height between the two peaks is not statistically significant, and the bin resolution is op- timal. The two peaks, having separation 0.36±0.13 in pulse phase, is reminiscent of the γ-ray profile of the Vela pulsar (e.g. Kanbach et al. 1994), which has comparable spin-down – 13 – parameters. Similar γ-ray pulse profiles are also observed for the Crab pulsar, Geminga, and PSR B1951+32 (Thompson 1996 and references therein). Hard X-ray observations of the Vela pulsar reveal a pulse profile that is similar to that at γ-ray energies (Strickman & Harding 1998); this suggests that hard X-ray observations of PSR B1046−58 could confirm its identification as a source of pulsed γ-rays. It must be kept in mind, however, that the profile of other well-established high-energy γ-ray pulsars is not as clearly double peaked. For example, PSR B1706−44, whose spin-down parameters most closely resemble those of PSR B1046−58, has a γ-ray profile whose morphology could be interpreted, within the limited statistics, as either single or double peaked with small peak separation. Therefore the double peaked profile in Figure 1 does not lend unambiguous support to the proposed association. The γ-ray pulse's offset from the single radio pulse (Fig. 1) is similar to those seen in the known high-energy γ-ray pulsars. The offset in time between the peak of the radio pulse and the approximate centroid of the first of the two γ-ray peaks shown in Figure 1 is 0.19±0.12 in pulse phase, with the uncertainty dominated by the poor profile statistics. The phase offset to the approximate midpoint between the two apparent pulses is 0.37±0.13. The uncertainty in these numbers due to the uncertainty of 0.01 pc cm−3 in the single-epoch DM measurement (§1) is negligible. This offset is similar to those of the other γ-ray pulsars having similar γ-ray pulse morphologies, namely the Crab (considering the radio precursor as the analog of the PSR B1046−58 radio pulse), Vela, and PSR B1951+32 (Thompson 1996 and references therein). These show an approximate empirical trend of radio–γ-ray phase offset that increases with characteristic age (e.g. Thompson 1996). The observed radio/γ- ray offset for PSR B1046−58 is in agreement with this trend, which predicts an offset in the approximate range 0.32–0.38. 4.1.5. Previous Discussion of the PSR B1046−58 / 2EG J1049−5847 Association In addition to Thompson et al. (1994) noting a hint of a γ-ray pulsation from PSR B1046−58, Fierro (1995) discussed a possible association between the pulsar and 2EG J1049−5847 in detail. He noted the positional coincidence as well as the plausibility of the implied spectral and energy properties of the putative γ-ray pulsar. He also noted the absence of variabil- ity in the source. Yadigaroglu & Romani (1997) used a statistical test that quantifies the probability for positional coincidence of unidentified EGRET sources with nearby young objects. They identified the PSR B1046−58 / 2EG J1049−5847 association using this anal- ysis. Zhang & Cheng (1998) argued that the PSR B1046−58 / 2EG J1049−5847 association was plausible on the basis of predictions from a version of the outer gap model of γ-ray – 14 – production in pulsar magnetospheres. 4.1.6. Implications If the association between PSR B1046−58 and 2EG J1049−5847 is correct, then this is only the eighth pulsar to show evidence for high-energy γ-ray pulsations. The poor statistics in the folded profile and the presence of significant off-pulse emission preclude detailed anal- ysis of the γ-ray emission properties. Nevertheless, the pulsar's apparent γ-ray properties resemble those of other known γ-ray pulsars, particularly the Crab and Vela pulsars and PSR B1951+32 (radio and γ-ray pulse morphologies and phase offset). PSR B1046−58 fur- ther shares a property common to all other known "older" γ-ray pulsars (that is, excluding the Crab and PSR B1509−58), namely that its maximum luminosity appears to be in the high-energy γ-ray range. By contrast, Pivovaroff et al. (1999) show that PSR B1046−58's efficiency for conversion of E into X-rays in the ASCA range is only ∼ 1 × 10−4. We now consider what the implications of the possible γ-ray pulsations from PSR B1046−58 are for models of γ-ray emission from rotation-powered pulsars. If the pulsations are real, the observed phase offset of the radio pulse is difficult to understand in the context of models in which the γ-rays are produced by particles which are accelerated along open field lines near the neutron star by electric fields (e.g. Dougherty & Harding 1982, Dermer & Sturner 1994). In these models, the radio emission is thought to arise in the same polar cap region, so the radio pulse should be observed between the γ-ray peaks, which result from emission from a hollow cone. The conal geometry results from the smaller radius of curvature of magnetic field lines near the polar cap rim (see Harding 1996 for a review). This radio/γ-ray phase problem has been noted for other γ-ray pulsars as well, e.g. Vela (Daugherty & Harding 1996); the results reported here argue that the problem is general. Also, emission beams in the polar cap model tend to be narrow and aligned with the magnetic and spin axes (e.g. Sturner & Dermer 1994, Harding & Muslimov 1998); this can pose a problem for the pulsar birth rate. However, the radio phase offset problem, as argued by Harding & Muslimov (1998), seems general to both the polar cap γ-ray emission model as well as to geometric models of thermal X-ray pulsations, while the relative phases of the latter two are consistent. Polar cap models therefore make a testable prediction regarding the phase of the peak of any thermal X-ray emission that may one day be detected from PSR B1046−58. The observed γ-ray pulse morphology and radio phase offset for PSR B1046−58 are, by contrast, well explained in the outer gap model. In this model, the γ-ray emission is produced in the charge-depleted region between the closed magnetic field lines and the velocity of light cylinder (e.g. Cheng, Ho, & Ruderman 1986, Chiang & Romani 1994, Romani & Yadigaroglu – 15 – 1995, Zhang & Cheng 1998). The double peaked γ-ray pulse morphology in this model arises from a single pole's outer gap that has uniform emissivity along its field lines but in which the radiation is subject to significant relativistic aberration and time-of-flight delays through the magnetosphere; the expected pulse morphology for most viewing geometries is double peaked, as long as the magnetic inclination is not small. The morphology and phase are also consistent with the geometry proposed by Manchester (1996) in which the radio emission is produced in the outer gap as well. The outer gap model predicts bridge emission between the two γ-ray peaks, whose outer edges should be sharp; this could be tested with an improved γ-ray pulse profile. This model results in large beaming fractions that are consistent with the pulsar birth rate; this implies that most, if not all, of the unidentified EGRET low- latitude sources that have constant fluxes are young rotation-powered pulsars (Yadigaroglu & Romani 1997), although most of them will not be detectable in radio waves from Earth because of smaller radio beaming fractions. We note, however, that recently reported very tentative evidence for γ-ray emission from the millisecond pulsar PSR J0218+4232 (Kuiper et al. 1998) could pose a problem for the outer gap model, which predicts that millisecond pulsars should not be detectable in high-energy γ-rays (Romani & Yadigaroglu 1995). 4.1.7. Unpulsed γ-ray Emission from PSR B1046−58? The spatial analysis of the EGRET data presented in §3.1.2 showed evidence for off- pulse emission consistent with coming from a point source coincident with the pulsar. For E > 400 MeV, this emission has flux (1.9 ± 0.6) × 10−10 erg cm−2 s−1, which corresponds E, for isotropic emission. The poor spatial resolution of EGRET obviously to ∼ 10% of precludes a firm association between the pulsar and this unpulsed emission, although we note that off-pulse γ-ray emission has been seen for the Crab, Vela and Geminga pulsars, with that of the Crab possibly originating in the surrounding nebula (Fierro et al. 1998). 4.2. PSR J1105−6107 Since we find no evidence for γ-ray pulsations for PSR J1105−6107, we set an upper E that is converted into high-energy γ-rays beaming in our direction. limit on the fraction of The counterpart to 2EG J1103−6106 in the 3EG catalog, 3EG J1102−6103, has flux (32.5± 6.2)×10−8 photons cm−2 s−1 for E > 100 MeV, and photon index 2.47±0.21 (Hartman et al. 1999). Using these numbers, we infer with 99% confidence that PSR J1105−6107 converts less than 8 × 10−11 erg cm−2 s−1 and 3 × 10−11 erg cm−2 s−1 for duty cycles of 0.5 and 0.1, respectively. These limits correspond to efficiencies for conversion of spin-down luminosity of – 16 – less than 0.014 and 0.006 for 1 sr beaming and assuming a distance of 7 kpc, for duty cycles 0.5 and 0.1, respectively. These limits are not constraining on models of γ-ray emission. 5. Conclusions We have found evidence for γ-ray emission from the young, energetic radio pulsar PSR B1046−58, which, if correct, establishes a counterpart for the previously unidenti- fied high-energy γ-ray source 2EG J1049−5847 (GEV J1046−5840, 3EG J1048−5840). The evidence can be summarized as follows. (i) PSR B1046−58 has the ninth highest value of E/d2 of the known radio pulsars, with six of the eight higher slots occupied by known γ-ray pulsars. This suggests that PSR B1046−58 is likely to be detectable in γ-rays. (ii) The radio pulsar is spatially coincident with the unidentified EGRET source 2EG J1049−5847. (iii) The folded light curve for E > 400 MeV, obtained using contemporaneous radio ephemerides, is characterized by an H-test statistic 22.5; the chance probability of this value or higher having occurred is 1.2 × 10−4. This number should be multiplied by a factor of two for the two pulsars we searched. (iv) A spatial likelihood analysis of on- and off-pulse data reveals a point source coincident with the radio pulsar position whose significance is much greater on- pulse than off (10.2σ versus 5.8σ). (v) The only hard X-ray point source detected in the 95% EGRET error box is spatially coincident with the pulsar (Pivovaroff et al. 1999). (vi) The γ-ray light curve morphology resembles those of other the known γ-ray pulsars (the Crab and Vela pulsars, Geminga, and PSR B1951+32), namely, two narrow peaks separated by ∼0.4 in pulse phase. (vii) The inferred radio/γ-ray phase offset is consistent with trends seen for other pulsars. The implied efficiency for conversion of spin-down energy into E > 400 MeV γ-rays, 0.011±0.003 for d = 3 kpc and 1 sr beaming, is consistent with those seen for other pulsars. We also found evidence for spatially unresolved unpulsed γ-ray emission from the pulsar position, as has been reported for the Crab, Vela and Geminga pulsars. If the association between PSR B1046−58 and 2EG J1049−5847 emission exists, it should be confirmed by GLAST, the planned Gamma-ray Large Area Space Telescope, which will provide significantly higher angular resolution and more detector area than EGRET. It may be possible to confirm the γ-ray pulsations by detecting magnetospheric X-ray pulsa- tions from PSR B1046−58 at hard X-ray energies. The similarity of this pulsar's rotational parameters and high-energy γ-ray efficiency to those of PSR B1706−44 suggest TeV emission might be detectable. Thermal emission from the initial cooling of this young neutron star may also be detectable. If the association between PSR B1046−58 and 2EG J1049−5847 holds, that seven of the top nine pulsars ranked by E/d2 are γ-ray pulsars (with one of the two remainders not – 17 – having been studied at γ-ray energies – see Table 1) implies at the very least that the radio and γ-ray beams are aligned closely, or that γ-ray beams are wide. In fact, the morphology of the observed γ-ray pulse and the phase offset of the radio pulse provide support for the outer gap model of γ-ray emission (Romani & Yadigaroglu 1995 and references therein). This model predicts that an improved γ-ray profile should find bridge emission between the pulses whose outer edges should be sharp. In this model, most or all of the unidentified EGRET sources that have constant fluxes are young rotation-powered pulsars like PSR B1046−58, though most radio beams will not be detectable from Earth due to small radio beaming fractions. Our analysis of EGRET data from the direction of PSR J1105−6107 finds no evidence for γ-ray pulsations. This pulsar converts at most 0.014 (99% confidence) of its spin-down power to E > 100 MeV γ-rays beamed in our direction, assuming 1 sr beaming, a duty cycle of 0.5, and a distance to the source of 7 kpc. This limit is not constraining on models. Acknowledgements We are grateful to Dave Thompson for many helpful conversations, and Robert Hartman for communication of Third EGRET Catalog numbers prior to publication. We also thank Russell Pace for help with the Parkes observations. This research has made use of data obtained through the High Energy Astrophysics Science Archive Research Center Online Service, provided by the NASA/Goddard Space Flight Center. This work was supported by NASA grants NAG 5-3178 and NAG 5-3683 to VMK, and NASA LTSA grant NAG 5-3384 to JRM. We thank Joe Fierro for inspiration and assistance in the initial undertaking of this project. – 18 – REFERENCES Backer, D. C., Hama, S., Van Hook, S., & Foster, R. S. 1993, ApJ, 404, 636 Bailes, M. et al. 1994, ApJ, 425, L41 Bertsch, D. L., Dame, T. M., Fichtel, C. E., Hunter, S. D., Sreekumar, P., Stacy, J. G., & Thaddeus, P. 1993, ApJ, 416, 587 Buccheri, R. et al. 1983, A&A, 128, 245 Caraveo, P. A., Bignami, G. F., Mignani, R., & Taff, L. G. 1996, ApJ, 461, 91 Cheng, K. S., Ho, C., & Ruderman, M. 1986, ApJ, 300, 500 Cheng, K. S. & Zhang, L. 1998, ApJ, 498, 327 Chiang, J. & Romani, R. W. 1994, ApJ, 436, 754 Crawford, F., Kaspi, V. M., Manchester, R. N., & Lyne, A. G. 1999, in preparation Daugherty, J. K. & Harding, A. K. 1996, ApJ, 458, 278 de Jager, O. C. 1994, ApJ, 436, 239 Dermer, C. D. & Sturner, S. J. 1994, ApJ, 420, L75 Dougherty, J. K. & Harding, A. K. 1982, ApJ, 252, 337 Esposito, J. A., Hunter, S. D., Kanbach, G., & Sreekumar, P. 1996, ApJ, 461, 820 Fierro, J. M. 1995. PhD thesis, Stanford University Fierro, j. M., Michelson, P. F., Nolan, P. L., & Thompson, D. J. 1998, ApJ, 494, 734 Gotthelf, E. V. & Kaspi, V. M. 1998, ApJ, 497, L29 Harding, A. K. 1996, in Pulsars: Problems and Progress, IAU Colloquium 160, eds. S. Johnston, M. A. Walker, & M. Bailes, San Francisco, Astronomical Society of the Pacific, 315 Harding, A. K. & Muslimov, A. G. 1998, ApJ, 500, 862 Hartman, R. C. et al. 1999, ApJ. in press Hunter, S. D. et al. 1997, ApJ, 481, 205 – 19 – Johnston, S., Lyne, A. G., Manchester, R. N., Kniffen, D. A., D'Amico, N., Lim, J., & Ashworth, M. 1992, MNRAS, 255, 401 Johnston, S., Manchester, R. N., Lyne, A. G., Kaspi, V. M., & D'Amico, N. 1995, A&A, 293, 795 Kaaret, P. & Cottam, J. 1996, ApJ, 462, L35 Kanbach, G. et al. 1994, A&A, 289, 855 Kanbach, G. et al. 1996, A&AS, 120, 461 Kaspi, V. M., Bailes, M., Manchester, R. N., Stappers, B. W., & Bell, J. F. 1996, Nature, 381, 584 Kaspi, V. M., Bailes, M., Manchester, R. N., Stappers, B. W., Sandhu, J. S., Navarro, J., & D'Amico, N. 1997, ApJ, 485, 820 Kaspi, V. M., Crawford, F., Manchester, R. N., Lyne, A. G., Camilo, F., D'Amico, N., & Gaensler, B. M. 1998, ApJ, 503, L161 Kuiper, L., Hermsen, W., Verbunt, F., Belloni, T., & Lyne, A. 1998, in Proceedings of the 3rd INTEGRAL Workshop: The Extreme Universe, http://xxx.lanl.gov/abs/astro- ph/9812408 Lamb, R. C. & Macomb, D. J. 1997, ApJ, 488, 872 Leahy, D. A., Darbro, W., Elsner, R. F., Weisskopf, M. C., Sutherland, P. G., Kahn, S., & Grindlay, J. E. 1983, ApJ, 266, 160 Lyne, A. G. 1996, in Pulsars: Problems and Progress, IAU Colloquium 160, eds. S. Johnston, M. A. Walker, & M. Bailes, San Francisco, Astronomical Society of the Pacific, 73 Manchester, R. N. 1996, in Pulsars: Problems and Progress, IAU Colloquium 160, eds. S. Johnston, M. A. Walker, & M. Bailes, San Francisco, Astronomical Society of the Pacific, 193 Marshall, F. E., Gotthelf, E. V., Zhang, W., Middleditch, J., & Wang, Q. D. 1998, ApJ, 499, L179 Mattox, J. R. et al. 1996, ApJ, 461, 396 McLaughlin, M. A., Mattox, J. R., Cordes, J. M., & Thompson, D. J. 1996, ApJ, 473, 763 – 20 – Merck, M. et al. 1996, A&AS, 120, 465 Navarro, J. 1994. PhD thesis, California Institute of Technology Pivovaroff, M., Kaspi, V. M., & Gotthelf, E. V. 1999, ApJ, submitted Qiao, G. J., Manchester, R. N., Lyne, A. G., & Gould, D. M. 1995, MNRAS, 274, 572 Ramanamurthy, P. V. et al. 1995, ApJ, 447, L109 Ramanamurthy, P. V., Fichtel, C. E., Kniffen, D. A., Sreekumar, P., & Thompson, D. J. 1996, ApJ, 458, 755 Romani, R. W. & Yadigaroglu, I.-A. 1995, ApJ, 438, 314 Standish, E. M. 1982, A&A, 114, 297 Stappers, B. W., Gaensler, B. M., Johnston, S., & Frail, D. A. 1999. in preparation Steinberger, J., Kaspi, V. M., & Gotthelf, E. V. 1999, in Proceedings of the Elba Work- shop: Neutron Stars and Supernova Remnants, Memorie della Societa' Astronomica Italiana, in press Strickman, M. S. & Harding, A. K. 1998, BAAS, 192, 74.04 Sturner, S. J. & Dermer, C. D. 1994, ApJ, 420, L79 Sturner, S. J. & Dermer, C. D. 1995, A&A, 293, L17 Swanenburg, B. N. et al. 1981, ApJ, 243, L69 Taylor, J. H. & Cordes, J. M. 1993, ApJ, 411, 674 Taylor, J. H. & Weisberg, J. M. 1989, ApJ, 345, 434 Thompson, D. J. et al. 1993, ApJS, 86, 629 Thompson, D. J. et al. 1994, ApJ, 436, 229 Thompson, D. J. et al. 1995, ApJS, 101, 259 Thompson, D. J. 1996, in Pulsars: Problems and Progress, IAU Colloquium 160, eds. S. Johnston, M. A. Walker, & M. Bailes, San Francisco, Astronomical Society of the Pacific, 307 Thompson, D. J. et al. 1999, ApJ, in press – 21 – Torii, K. et al. 1998, ApJ, 494, L207 Ulmer, M. P. et al. 1993, ApJ, 417, 738 Yadigaroglu, I.-A. & Romani, R. W. 1995, ApJ, 449, 211 Yadigaroglu, I. A. & Romani, R. W. 1997, ApJ, 476, 347 Zhang, L. & Cheng, K. S. 1998, A&A, 335, 234 This preprint was prepared with the AAS LATEX macros v4.0. – 22 – Fig. 1.- Bottom panel: folded EGRET light curve for PSR B1046−58 for E > 400 MeV. Top panel: radio profile at 20 cm from the Parkes observatory. – 23 – Fig. 2.- Upper panel: number of events within θ67 for PSR B1046−58 as a function of lower energy cutoff in MeV. Lower panel: results of three statistical tests applied to folded profiles for PSR B1046−58 as a function of lower energy cutoff in MeV. The probabilities represent the chance of the measured value of the statistical test or higher having been obtained by chance. The three tests are the H test, the χ2 test for 10 bins (9 degrees of freedom), and the Z 2 4 test. – 24 – Fig. 3.- Folded EGRET light curve for PSR J1105−6107 for E > 400 MeV. – 25 – Table 1. Summary of Known High-Energy γ-ray Pulsars and PSRs B1046−58 and J1105−6107, ordered by τ . Pulsara Crab Vela B1706−44 B1046−58 J1105−6107 B1951+32 B0656+14 Geminga B1055−52 P ms 33 89 102 124 63 39 385 237 197 log τ b log Bc log Ed yr 3.1 4.1 4.2 4.3 4.8 5.0 5.0 5.5 5.7 G erg s−1 cm−2 12.6 12.5 12.5 12.5 12.0 11.7 12.7 12.2 12.0 38.7 36.8 36.5 36.3 36.4 36.6 34.6 34.5 34.5 de kpc 2.0 0.5 1.8 3.0 7.1 2.4 0.8 0.157+0.059 −0.034 1.5 E/d2 f rank g ηγ % 1 2 4 9 23 6 20 3 39 0.013±0.006 0.18±0.07 0.72±0.47 1.1±0.3 < 1.4 0.26±0.17 ∼0.1 1.61+0.22 −0.08 6–13 aPulsars in bold are the subject of this paper. bτ ≡ P/2 P cB ≡ 3.2 × 1019(P P )1/2 G d E ≡ 4π2I P /P 3 eDistances from Taylor & Cordes (1993) except for Geminga which is from Caraveo et al. (1996). fRanked 5th is PSR B1509−58, detected in low-energy γ-rays (Ulmer et al. 1993); 7th is the millisecond pulsar PSR J0437−4715; 8th is PSR J1617−5055, which was only recently discovered (Torii et al. 1998, Kaspi et al. 1998). gFor E > 100 MeV and 1 sr beaming, except for PSR B1046−58 which is for E > 400 MeV. Uncertainties include nominal distance uncertainties. Numbers are from Fierro (1995) except for Geminga, for which we have used the updated distance, PSR B0656+14, which is from Ramanamurthy et al. (1996), and PSR B1055−52, which is from Thompson et al. (1999). For PSR J1105−6107, the 3σ upper limit for duty cycle 0.5 is quoted. – 26 – Table 2. Summary of EGRET Observations for PSR B1046−58. VPa 0007 0060 0080 0120 0140 0170 0230 0320 2080 2150 2170 2300 2305 3010 3140 3150 3160 3385 4020 4025 4150 4240 5220 5310 6270 6300 MJD Range 48386–48392 48463–48476 48490–48504 48546–48560 48574–48588 48617–48631 48700–48714 48798–48805 49020–49027 49078–49083 49089–49097 49195–49198 49198–49202 49216–49223 49355–49368 49368–49375 49375–49384 49595–49615 49643–49650 49650–49657 49818–49832 49908–49923 50245–50248 50359–50371 50693–50700 50714–50728 ∆θb ◦ 21.1 31.2 25.2 31.3 2.7 32.4 34.8 22.5 28.1 32.4 32.4 11.1 8.7 24.1 16.8 16.8 28.7 24.1 23.5 19.8 27.1 30.9 2.7 3.7 1.6 1.6 N(θ67) Exposure 614 279 1032 235 3252 140 30 289 163 52 34 335 416 361 775 381 157 649 156 195 261 65 136 366 234 437 0.52 0.54 1.01 0.49 3.08 0.45 0.14 0.40 0.18 0.08 0.12 0.45 0.60 0.40 1.09 0.60 0.24 0.87 0.24 0.40 0.36 0.24 0.33 1.16 1.01 1.89 aEGRET Viewing Period. bAngular offset of pulsar from center of EGRET field of view. – 27 – Table 3. Summary of EGRET Observations of PSR J1105−6107. VPa 2300 2305 3010 3140 3150 3160 3385 4020 4025 4150 4240 5220 5310 6270 6300 MJD Range 49195–49198 49198–49202 49216–49223 49355–49368 49368–49375 49375–49384 49595–49615 49643–49650 49650–49657 49818–49832 49908–49923 50245–50248 50359–50371 50693–50700 50714–50728 ∆θb ◦ 14.0 11.8 27.0 13.7 13.7 26.9 27.0 20.4 16.7 28.0 28.9 5.0 6.7 3.6 3.6 N(θ67) Exposure 293 412 245 826 420 213 430 201 226 213 114 128 304 205 419 0.38 0.52 0.30 1.32 0.71 0.27 0.67 0.32 0.31 0.35 0.28 0.32 0.99 0.96 1.81 aEGRET Viewing Period. bAngular offset of pulsar from center of EGRET field of view.
astro-ph/0703630
2
0703
2007-06-09T06:57:31
Statefinder diagnostic for the modified polytropic Cardassian universe
[ "astro-ph" ]
We apply the Statefinder diagnostic to the Modified Polytropic Cardassian Universe in this work. We find that the Statefinder diagnostic is quite effective to distinguish Cardassian models from a series of other cosmological models. The $s-r$ plane is used to classify the Modified Polytropic Cardassian models into six cases. The evolutionary trajectories in the $s-r$ plane for the cases with different $n$ and $\beta$ reveal different evolutionary properties of the universe. In addition, we combine the observational $H(z)$ data, the Cosmic Microwave Background (CMB) data and the Baryonic Acoustic Oscillation (BAO) data to make a joint analysis. We find that \textbf{Case 2} can be excluded at the 68.3% confidence level and any case is consistent with the observations at the 95.4% confidence level.
astro-ph
astro-ph
Statefinder diagnostic for the modified polytropic Cardassian universe Department of Astronomy, Beijing Normal University, Beijing, 100875, P.R.China Ze-Long Yi and Tong-Jie Zhang∗ We apply the statefinder diagnostic to the modified polytropic Cardassian universe in this work. We find that the statefinder diagnostic is quite effective to distinguish Cardassian models from a series of other cosmological models. The s − r plane is used to classify the modified polytropic Cardassian models into six cases. The evolutionary trajectories in the s − r plane for the cases with different n and β reveal different evolutionary properties of the universe. In addition, we combine the observational H(z) data, the cosmic microwave background data and the baryonic acoustic oscillation data to make a joint analysis. We find that Case 2 can be excluded at the 68.3% confidence level and any case is consistent with the observations at the 95.4% confidence level. PACS numbers: 98.80.Es,95.35.+d,98.80.Jk I. INTRODUCTION Recent type Ia Supernova observations, along with other observations such as cosmic microwave background (CMB) and galaxy power spectra, support the fact that the present expansion of our universe is accelerating [1, 2]. In order to explain the accelerated expansion of the universe, cosmologists have tried to explore many cosmological models. A negative pressure term called dark energy is always taken into account, such as the cos- mological constant model with equation of state ωDE = pDE/ρDE = −1 where pDE and ρDE are pressure and den- sity of the dark energy, respectively [3], the quiessence whose equation of state ωQ is a constant between -1 and -1/3 [4], and the quintessence which is described in terms of a scalar field φ [5, 6]. Some other candidates are con- structed in different ways, such as the braneworld models which explain the acceleration through the fact that the general relativity is formulated in 5 dimensions instead of the usual 4 [7], and the Cardassian models which in- vestigate the acceleration of the universe by a modifica- tion to the Friedmann-Robertson-Walker (FRW) equa- tion [8]. In these cases, the dark energy component is not involved, but the accelerated expansion can still be obtained. The quantity ωeff is usually described as an effective equation of state for these models, and can be expressed by the Hubble parameter H and its derivatives with respect to redshift z [4]. As so many cosmological models have been developed, a discrimination between these contenders becomes nec- essary. In order to achieve this aim, Sahni et al. proposed a new geometrical diagnostic named the statefinder pair {r, s}, where r is generated from the scalar factor a and its derivatives with respect to the cosmic time t, just as the Hubble parameter H and the deceleration parame- ter q, and s is a simple combination of r and q [9]. The statefinder pair has been used to discriminate a series of cosmological models, including the LCDM universe with ∗Electronic address: [email protected] a cosmological constant Λ and a cold dark matter term (CDM), the Chaplygin gas, the holographic dark energy models, the quintessence, the braneworld models and so on. Clear differences for the evolutionary trajectories in the s − r plane have been found. In this paper, we apply the statefinder diagnostic to the modified polytropic Cardassian universe. The origi- nal Cardassian model based on two parameters Ωm0 and n was first suggested by Freese & Lewis [8]. It was gen- erated from a modification to the Friedmann equation. Such a universe is spatially flat and accelerating today. But it involves no dark energy term and is dominated by merely matter and radiation. These Cardassian mod- els predict the same distance-redshift relation as generic quintessence models, although they generate from com- pletely different physical principles. A generalized Car- dassian model -- the modified polytropic Cardassian uni- verse can be obtained by introducing an additional pa- rameter β into this model [10, 11, 12], which reduces to the original model if β = 1. The distance-redshift re- lation predictions of generalized Cardassian models can be very different from generic quintessence models, and can be differentiated with data from surveys of Type Ia Supernovae such as SuperNova/Acceleration Probe (SNAP). In all, the modified polytropic Cardassian uni- verse can predict more fresh physical information than the original Cardassian. It is worthy of more detailed discussions. In this work, we successfully classify the modified poly- tropic Cardassian models into six cases by the statefinder diagnostic. The cases with different n and β correspond to different evolutionary trajectories in the s − r plane. The fact that LCDM corresponds to a fixed point (0, 1) in the s − r plane plays a significant role for our classification. Also, it is very important to find where the evolutionary trajectories start and end. Another key standpoint is whether the evolutionary trajectory has a crossing with LCDM. We also study the relation between (n, β) and the crossing redshift zC, at which the modified polytropic Cardassian universe intersects with LCDM in the s − r plane. We find that the modified polytropic Cardassian models can be distinguished from other inde- pendent cosmological models by the statefinder diagnos- tic. In addition, we use the observational H(z) data de- rived from ages of the passively evolving galaxies [13], the newly measured value of the CMB shift parameter R [14] and the A-parameter which describes the baryonic acous- tic oscillation (BAO) peak [15] to make a combinational constraint. We assume a prior of H0 = 72 ± 8 suggested by the Hubble Space Telescope (HST) Key Project [16]. From the confidence regions, we find that Case 2 is not consistent with the observations at the 68.3% confidence level and all the six cases do not conflict with the obser- vations at the 95.4% confidence level. This paper is organized as follows: In Sec.2, we briefly review the modified polytropic Cardassian universe. In Sec.3, we introduce the statefinder pair {r, s}. In Sec.4, we apply the statefinder diagnostic to various Cardassian models. In Sec.5, we make a combinational constraint on the parameters of the modified polytropic Cardassian universe. In Sec.6, the discussions and the conclusions are given. II. THE MODIFIED POLYTROPIC CARDASSIAN UNIVERSE Measurements of CMB suggest a flat geometry for our If we consider a spatially flat FRW universe [17, 18]. universe, the basic equation can be written as H 2 = 8πG 3 ρ, (1) where G is Newton's universal gravitation constant and ρ is the density of summation of both matter and vacuum energy. Freese & Lewis [8] proposed a model called the Cardassian universe by adding a term on the right side of Eq.(1), H 2 = 8πG 3 ρm + Bρn m, (2) where n is assumed to satisfy n < 2/3 and ρm is always taken as a contribution of only the matter (in this paper we do not plan to consider radiation). If n = 0, it is identical to the cosmological constant universe. If B = 0, it reduces to the usual FRW equation, but with density of only matter. Thus, it is easy to get a new expression of H from Eq.(2), H 2 = H 2 0 [Ωm0(1 + z)3 + (1 − Ωm0)(1 + z)3n], (3) by using ρm = ρm0(1 + z)3 = Ωm0ρc(1 + z)3, (4) where ρm0 is current value of ρm and ρc = 3H 2 0 /8πG is the critical density of the universe. Clearly, this model predict the same distance-redshift relation as the quiessence with ωQ = n−1. But we notice that they have completely different essentials. The quiessence requires a dark energy component while the Cardassian does not. 2 In the earlier time, the universe is dominated by the first term in Eq.(3) and the ordinary FRW behavior works. The additional term which consists of only matter gradually becomes a dominant driver afterwards. This transition was found to occur at z ∼ O(1) [8]. Then the universe is caused to accelerate. The period of accelera- tion for this model is usually called the Cardassian era. The Cardassian model is attractive because the universe is flat and accelerating today but no vacuum energy is in- volved. And it has been demonstrated compatible with a series of observational tests, including the CMB data, the age of the universe, the structure formation and the cluster baryon fraction [8]. The modified polytropic Cardassian universe can be obtained by introducing an additional parameter β into the above model [10, 11, 12], H 2 = H 2 0 [Ωm0(1 + z)3 + (1 − Ωm0)fX(z)], (5) where fX(z) = Ωm0 Ω−β m0 − 1 1 − Ωm0 (1 + z)3[(1 + (1 + z)3(1−n)β )1/β − 1]. (6) The two parameters (n, β) are usually taken as n < 2/3 and β > 0. If β = 1, the model reduces to the original one characterized by Eq.(3) while if fX(z) = 1, the model just corresponds to LCDM. Similar to Wang et al., we also take Ωm0 = 0.3 as a prior in subsequent discussions. In the work of Wang et al., this model was compared with current supernova data and CMB data. It was proved that the existing data can be well fit for several chosen values of n and β. Also, the simulated data were constructed to make a discrimination between the modified polytropic Car- dassian universe and LCDM as well as the quintessence. Once Ωm0 is known with an accuracy of 10%, SNAP can determine the sign of the time dependence of dark energy density which provides a first discrimination be- tween various cosmological models [12]. In this work, we use a geometrical tool-the statefinder diagnostic to make a classification and a discrimination about the modified polytropic Cardassian models. III. A BRIEF OVERVIEW OF THE STATEFINDER DIAGNOSTIC The Hubble parameter H = a/a and the deceleration parameter q = −a/aH 2 are two traditional geometrical diagnostics. They only depend on the scalar factor a and its derivatives with respect to t, i.e., a and a. Through a = 1/(1 + z), the deceleration parameter q can be ex- pressed as q(z) = H ′ H (1 + z) − 1, (7) where H ′ is the derivative of H with respect to redshift z. The deceleration parameter q is a good choice to de- scribe the expansion state of our universe, but it is not 3 perfect enough to characterize the cosmological models uniquely. This shortage can be easily seen from the fact that many models may correspond to the same current value of q. And this difficulty can be overcome by an- other geometrical diagnostic -- the statefinder pair {r, s}. This approach has been used to distinguish a series of cosmological models successfully. For a spatially flat universe, the statefinder r is defined as follows [9] the statefinder diagnostic in several literatures [4, 9]. In the s − r plane, a vertical line with s = n and r changing monotonically from 1 to 1 + 9n(n − 1)/2 represents the evolutionary trajectory of the universe. For cases with β 6= 1, we first pay our attention to the epoches in the far past (a → 0) and the far future (a → +∞). It is clear that z → +∞ represents the former case and z → −1 stands for the latter. From Eq.(12), it is easy to find the limit condition r = ··· a aH 3 , (8) lim z→+∞ r(z) = 1, (14) ··· a is the third derivative of a with respect to t. s where is just a combination of r and q, s = r − 1 3(q − 1/2) . (9) The statefinder pair was first introduced to analyze a flat universe with a cold matter and a dark energy term. For these contenders, r is given by r = 1 + 9 2 ωDE(1 + ωDE)ΩDE − 3 2 ωDE H ΩDE. (10) where ΩDE = ρDE/ρc and ωDE is the derivative of ωDE with respect to t. And the other diagnostic, s = 1 + ωDE − 1 3 ωDE ωDEH . (11) From the two equations above it is easy to realize that LCDM corresponds to a fixed point (0, 1) in the s − r plane and the standard cold dark matter (SCDM) uni- verse with no dark energy term locates at (1, 1) forever. For this particularity, the current values of {r, s} pro- vide a considerable way to measure the distance from a specific model to LCDM. More generally, r and s can be given in terms of the Hubble parameter H and its first and second derivatives H ′ and H ′′ with respect to redshift z, which means that the value of r in the far past is inde- pendent on n and β. However, we can not derive the similar properties for the value of s from Eq.(13). It is related to both of n and β. For the limit of z → −1, from Eq.(12), we get lim z→−1 r(z) = 1 + 9 2 n(n − 1). (15) And from Eq.(13), we have lim z→−1 s(z) = n. (16) The values of both s and r are independent on β. Generally, some models have a crossing with LCDM in the s − r plane. By substituting the Hubble parameter, Eq.(5), in the expressions of r and s, Eq.(12) and Eq.(13), we find that the crossing with LCDM happens at the redshift zC = [(Ω−β m0 − 1) −n 1 − β + nβ 1 3(1−n)β − 1. ] (17) And we notice that the crossing can happen only if the following inequality is satisfied, (Ω−β m0 − 1) −n 1 − β + nβ ≥ 0. (18) r(z) = 1 − 2 H ′ H (1 + z) + H ′′ H + ( H ′ H )2, (12) Due to the prior Ωm0 = 0.3 and β > 0, Ω−β m0 − 1 > 0 is naturally satisfied. Thus we may describe the condition in Eq.(18) equivalently as s(z) = −2H ′(1 + z)/H + H ′′/H + (H ′/H)2 3(H ′(1 + z)/H − 3/2) . (13) Thus we can use the new tool to describe the evolutionary trajectories of the modified polytropic Cardassian uni- verse. IV. THE s − r PLANE FOR MODIFIED POLYTROPIC CARDASSIAN UNIVERSE For the original Cardassian model with β = 1, it cor- responds to a compatible expression with the quiessence with ωQ = n − 1. The quiessence has been studied using n 1 − β + nβ ≤ 0. (19) In order to understand the relation among n, β and zC clearly, we draw a β − zC plane for several fixed n in Fig.1. For n > 0, the crossing takes place at zC > 0 for any β, and zC changes little for larger values of β. For n < 0, whether zC > 0 or zC < 0 depends on the values of β. And zC is nearly equal to -1 if β is small enough. For the particular case which satisfies 1 − β + nβ = 0, the expected crossing happens at zC → ∞. Clearly, n = 0 and β = 1/(1 − n) are the two critical conditions. Now we use Eq.(19) to classify the modified polytropic Cardassian universe. We draw an n − β plane in Fig.2 n=−0.5 n=−1.0 n=−1.5 n=−2.0 0.1 0.2 0.3 β 0.4 0.5 0.6 r n=0.2 n=0.3 n=0.4 n=0.5 4.5 4 3.5 3 2.5 2 1.5 1 0.5 4 LCDM SCDM 1.5 2 2.5 β 3 3.5 0 −1 −0.5 0 s 0.5 1 C z 2 1 0 −1 0 C z 4 3 2 1 0 FIG. 1: β − zC planes for n=-0.5, -1.0, -1.5 and -2.0 (the top panel) as well as n=0.2, 0.3, 0.4 and 0.5 (the bottom panel). 3 2.5 2 β 1.5 1 0.5 0 −1 Case2 n<0 & β>1/(1−n) Case4 0<n<2/3 & β >1/(1−n) Case3 n=0 Case6 β=1/(1−n) Case1 n<0 & 0<β<1/(1−n) Case5 0<n<2/3 & 0<β<1/(1−n) −0.8 −0.6 −0.4 −0.2 n 0 0.2 0.4 0.6 FIG. 2: n − β plane based on n and β. The vertical solid line represents n = 0 (Case 3), and the declined solid curve stands for β = 1/(1 − n) (Case 6). and divide it into different regions by the critical curves n = 0 and β = 1/(1 − n): Case 1 n < 0 and 0 < β < 1/(1 − n); Case 2 n < 0 and β > 1/(1 − n); Case 3 n = 0; Case 4 0 < n < 2/3 and β > 1/(1 − n); Case 5 0 < n < 2/3 and 0 < β < 1/(1 − n); Case 6 β = 1/(1 − n). (20)   Case 1 (n < 0 and 0 < β < 1/(1 − n)): n=-0.5, β=0.1, 0.2, 0.3, 0.4 and 0.5 are taken for a qualitative analysis. We plot the s − r plane in Fig.3. Although the original Cardassian model with β = 1 is not involved in this case, we still draw this curve for a comparison. FIG. 3: s − r plane for Case 1. The star (0, 1) corresponds to LCDM, the star (1, 1) SCDM and other dots the current values of the Statefinder pair, i.e., (s0, r0) (also for subsequent cases). And the solid curves from bottom to top correspond to the evolutionary trajectories of n=-0.5, β=0.1, 0.2, 0.3, 0.4 and 0.5 respectively. The dot-dash curves represent the critical case β = 2/3. For a comparison, the vertical dashed line stands for the case of β = 1. All the curves start at r = 1 and 0 < s < 1 on the horizontal line, i.e., on the right of LCDM. After passing by LCDM, they arrive at their common end (−0.5, 4.38) in far future. And the crossing with LCDM happens at zC=-0.999, -0.84, -0.50, -0.13 and 0.25 for β=0.1, 0.2, 0.3, 0.4 and 0.5 respectively. Particularly, the fixed point (0, 1) for LCDM just is the beginning point for the critical case of β = 1/(1 − n) = 2/3. Case 2 (n < 0 and β > 1/(1 − n)): To be consistent with Case 1, we consider n = −0.5 again, but β=0.8, 1, 2, 3, 4 and 5 respectively. The s − r plane is plotted in Fig.4. The crossing with LCDM never occurs for this case. The vertical line s = n for the evolutionary trajec- tory of β = 1 divides the whole plane into two parts. The left part with n < s < 0 stands for 1/(1 − n) < β < 1 and the right part with s < n corresponds to β > 1. And all the evolutionary trajectories commence at one point with r = 1. The same as Case 1, they arrive at (−0.5, 4.38) in the end. Case 3 (n = 0): We plot the s − r plane for β=0.3, 0.7, 1, 2, 3 and 4 in Fig.5 . One point with r = 1 acts as the beginning point, and (0, 1) the common end. This diagram is divided into two segments by the LCDM fixed point, or equivalently the modified polytropic Cardassian model with n = 0 and β = 1. The segment with r > 1 corresponds to β > 1 while the other segment with r < 1 corresponds to β < 1. Case 4 (0 < n < 2/3 and β > 1/(1 − n)): The s − r plane for n = 0.2, β=2, 3, 4 and 5 is plotted in Fig.6. Clearly, all the curves start with r = 1 and s < 0, and then pass by LCDM after an arc route. All the lat- 5 LCDM SCDM 3.5 3 2.5 2 1.5 1 0.5 r −6 −5 −4 −3 s LCDMSCDM −2 −1 0 1 0 −3 −2.5 −2 −1.5 −1 s −0.5 0 0.5 1 r 12 10 8 6 4 2 0 FIG. 4: s − r plane for Case 2. The solid curves from right to left correspond to the evolutionary trajectories of n=-0.5, β=0.8, 2, 3, 4 and 5 respectively. And the vertical dashed line stands for the case of β = 1. Same as Fig.3, the dot-dash curve represents the critical case β = 2/3. FIG. 6: s − r plane for Case 4. The solid curves from right to left stand for the evolutionary trajectories of n = 0.2, β=2, 3, 4 and 5 respectively. The dot-dash curve represents the critical case β = 5/4. For a comparison, the vertical dashed line stands for the case of β = 1. 4.5 4 3.5 3 2.5 2 1.5 1 0.5 r LCDM SCDM 0 −3 −2.5 −2 −1.5 −1 s −0.5 0 0.5 1 LCDM 1 SCDM 0.8 0.6 r 0.4 0.2 0 0 0.2 0.4 0.6 0.8 1 s FIG. 5: s − r plane for Case 3. The solid curves from bottom to top stand for β=0.3, 0.7, 2, 3 and 4 respectively. The case of β = 1 just corresponds to the fixed point for LCDM, i.e., (0, 1). FIG. 7: s − r plane for Case 5. The solid curves from right to left correspond to the evolutionary trajectories of n=0.2, β=0.3, 0.6, 0.9 and 1.1 respectively. And the vertical dashed line stands for the case of β = 1. Same as Fig.6, the dot-dash curve represents the critical case β = 5/4. ter evolutionary parts of the trajectories nearly overlap with each other. They are insensitive to β. The point (0.2, 0.28) is the common end. The crossing redshifts are not too far from each other for different values of β, i.e., zC ≃ 0.3. It is interesting that the fixed point (0, 1) for LCDM is just the beginning point for the critical case of β = 1/(1 − n) = 5/4. And the Cardassian mod- els for this case can satisfy the weak energy condition ω = pX/(ρm + ρX) > −1 although the effective equation of state satisfies ωeff = pX/ρX < −1. weff < −1 is con- sistent with many observations such as CMB and large scale structure data [19, 20]. Case 5 (0 < n < 2/3 and 0 < β < 1/(1 − n)): Same as Case 4, n = 0.2 is considered, but β=0.3, 0.6, 0.9, 1.0 and 1.1. We plot the s − r plane in Fig.7. All the evolutionary trajectories start with r = 1 and s > 0. They arrive at (0.2, 0.28) in the end, and never pass by LCDM. The whole plane is divided into two parts by the vertical line s = n (corresponding to the case of β = 1). The left part satisfies 1 < β < 5/4 and β < 1 is satisfied for the right. 4.5 4 3.5 3 2.5 2 1.5 1 0.5 0 r LCDM SCDM −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8 1 s FIG. 8: s − r plane for Case 6. The point (0, 1) stands for n = 0. Then n increases by a interval of 0.1 until 0.6 for cases on the right of (0, 1) and n decreases by a interval of 0.1 until -0.5 for cases on the left of (0, 1). The other vertical dashed lines stand for the evolutionary trajectories of cases with β = 1 and different values of n. Case 6 (β = 1/(1 − n)): This is another critical case besides Case 3. n=-0.5, -0.4, -0.3, -0.2, -0.1, 0, 0.1, 0.2, 0.3, 0.4, 0.5, and 0.6 are considered. For this case, crossing with LCDM happens at zC → ∞. We plot the s − r plane in Fig.8. All the evolutionary trajectories start at LCDM in the past and end at (−1 + 3n/2, 1 + 9n(n − 1)/2) in the future. We have successfully given a qualitative analysis to the modified polytropic Cardassian universe from the statefinder viewpoint and it has been classified into six cases. A prior Ωm0 = 0.3 is adopted in this work. We also take some other values of Ωm0 for comparison and notice that Ωm0 is not a sensitive parameter for our analysis in this work. And it has been clear that the s − r plane is effective enough to discriminate the modified polytropic Cardassian models. The fact that LCDM corresponds to a fixed point (0, 1) in the s − r plane plays a signifi- cant role for our classification. Whether the evolutionary trajectory has a crossing with LCDM is our basic stand- point. Also, the beginning points and the ending points are both key factors for shapes of the evolutionary tra- jectories. V. DATA ANALYSIS FROM OBSERVATIONAL H(z) DATA, CMB AND BAO In order to understand the six cases above more clearly, we use the observational H(z) data derived from the passively evolving galaxies, the cosmic microwave back- ground (CMB) shift parameter R and the A-parameter which describes the baryonic acoustic oscillation (BAO) peak to make a joint analysis. 6 The Hubble parameter H(z) depends on the differen- tial age of the universe in this form H(z) = − 1 1 + z dz dt , (21) which provides a direct measurement for H(z) through a determination of dz/dt. By using the differential ages of passively evolving galaxies determined from the Gemini Deep Deep Survey [21] and archival data [22, 23, 24, 25], Simon et al. determined a set of observational H(z) data in the range 0 . z . 1.8 and used them to constrain the dark energy potential and its redshift dependence [13]. Yi & Zhang first first used them to analyze the holo- graphic dark energy models and got a consistent result with others [26]. The model-independent shift parameter R can be de- rived from CMB data. It is defined as [14] R = pΩm0Z zr 0 dz E(z) , (22) where E(z) = H(z)/H0 and zr = 1089 is the redshift of recombination. From the three-year result of WMAP [27], Wang & Mukherjee estimated R = 1.70 ± 0.03 [28]. Using a large spectroscopic sample of 46748 lumi- nous red galaxies covering 3816 square degrees out to z = 0.47 from the Sloan Digital Sky Survey, Eisen- stein et al. [15] successfully found the acoustic peaks in the CMB anisotropy power spectrum, described by the model-independent A-parameter, A = pΩm0[ 1 z1E1/2(z1) Z z1 0 dz E(z) ]2/3, (23) where z1 = 0.35 is the redshift at which the acoustic scale has been measured. Eisenstein et al. [15] suggested the measured value of the A-parameter as A = 0.469 ± 0.017. The best-fit parameters of the modified polytropic Car- dassian universe can be determined by minimizing [Hth(zi) − Hob(zi)]2 (R − 1.70)2 (A − 0.469)2 + + , χ2 = Xi σ2 i 0.032 0.0172 (24) where Hth(zi) is the theoretical Hubble parameter at zi, Hob(zi) is the observational Hubble parameter at zi and σi is the corresponding 1σ error. We get the best-fit val- ues n = −1.85 and β = 0.23, with χ2 min = 10.12. The best-fit results correspond to Case 1. And the cross- ing with LCDM in the s − r plane occurs at zC=0.32. The current values of the diagnostics are s0=-0.19 and r0=1.59. The confidence regions in the n − β plane are plotted in Fig.9, from which we find that Case 2 can be excluded at the 68.3% confidence level and all the six cases are consistent with the observational data at the 95.4% confidence level. VI. DISCUSSIONS AND CONCLUSIONS The statefinder diagnostic is powerful to discriminate various cosmological models. Differences of the evolu- 2.5 2 1.5 1 0.5 β 0 −1 −0.8 −0.6 −0.4 −0.2 n 0 0.2 0.4 0.6 FIG. 9: Confidence regions in the n − β plane for the joint analysis (the regions from inner to outer stand for confidence levels at 68.3%, 95.4% and 99.7% respectively). The vertical solid line represents n = 0 (Case 3), and the declined solid curve stands for β = 1/(1 − n) (Case 6). tionary trajectories in the s − r plane among a series of cosmological models have been found. For example, LCDM corresponds to a fixed point (0, 1) and the point (1, 1) represents SCDM. For the holographic dark energy model with c = 1 [29], the curve in the s − r plane com- mences at (2/3, 1) in the past and ends at LCDM in the future, with s monotonically decreasing from 2/3 to 0 and r first decreasing from 1 to a minimum value and then rising to 1 [30]. Both the quintessence tracker models (with tracker potentials V = V0/φα) and the Chaplygin gas models have arc evolutionary trajectories, but in dif- ferent regions [4, 31, 32]. The conditions −1 ≤ s ≤ 0 and r ≤ 1 are satisfied for the former cosmological model while 0 ≤ s ≤ 1 and r ≥ 1 for the latter. The evo- lutionary trajectories of the coupled quintessence models have more complicated evolutionary properties [33]. And quintessence models with other potentials were studied by Evans et al.[31], also along with a generalization of {r, s} to a non-flat universe. Also, the q − r plane has been widely used for discus- sion on the evolutionary property of the universe. For example, the point (0.5, 1) corresponds to SCDM in the q − r plane and the horizontal straight line from (0.5, 1) to (-1, 1) stands for LCDM. The braneworld models have been studied too, including Disappearing Dark En- ergy (DDE) as the simplest case [4]. LCDM separates the first braneworld model (named BRANE1 in [4] and the effective equation of state satisfies ωeff ≤ −1) from the second braneworld model (named BRANE2 in [4] and the effective equation of state satisfies ωeff ≥ −1) and DDE models. DDE both begins and ends at SCDM, forming a loop. However, although both BRANE1 and BRANE2 commence their evolutions at (0.5, 1) and end at (-1, 1), the diagnostic r satisfies r ≥ 1 for BRANE1 while r ≤ 1 7 for BRANE2. As the s − r plane has been found robust for our clas- sification for the modified polytropic Cardassian models, we do not intend to make use of the q − r plane. For the modified polytropic Cardassian universe, the evolu- tionary trajectories can be picked out easily with help of the s − r plane if β 6= 1. In fact, the original Car- dassian model with β=1 can not be discarded from the quiessence with ωQ = n − 1 because they correspond to identical expressions for both r and s. Such a consistence has also been mentioned by Freese & Lewis, as well as how to distinguish the two models [8]. Distinct differences in the s − r plane have been real- ized for the cases with different n and β for the modified polytropic Cardassian universe. For Case 1 and Case 4, the crossing with LCDM happens at some zC, while the same state never occurs for Case 2 and Case 5. And Case 3 and Case 6 just are two critical cases. Also, the beginning points and the ending points are tightly related to the shapes of the evolutionary trajectories. They se- riously depend on n and β, especially n. As n and β are found to be sensitive to the modified polytropic Car- dassian models, constraining the two parameters exactly becomes a valuable task. We use the observational H(z) data, the CMB data and the BAO data to make a com- binational constraint. We find that Case 2 can be ex- cluded at the 68.3% confidence level and all the six cases are consistent with the observational data at the 95.4% confidence level. Recent other constraints suggest some results far from consistent with each other. For example, in the work of Wang et al., the choices of n = 0.2, β = 1; n = 0.2, β = 2 and n = 0.2, β = 3 are all consistent with SN Ia and CMB observations [12]. The first case corresponds to the quiessence with the equation of state ωQ = −0.8, and this case is consistent with Case 5. The latter two cases with the crossing at zC=0.29 and 0.26 respectively are in agreement with Case 4. And the cur- rent values of the diagnostics are s0=0.13, r0=0.58 and s0=0.16, r0=0.45 for n = 0.2, β=2 and n = 0.2, β=3 respectively. Nesseris & Perivolaropoulos [34] suggested another result n = −23+8 −0.010 with a super- nova data set consisting of 194 SN Ia [35, 36]. Both Case 1 and Case 2 may be included within the given error- bar. The best fitting values lie in Case 1 and the cross- ing with LCDM happens at zC=0.37. Meanwhile, the present quantities are s0=-0.29 and r0=1.94. Evans et al. provided a best fitting result n = −0.94 and β = 0.06 with a simulated data set [31]. This result is in agreement with Case 1. The crossing happens in the far future, i.e., at zC=-0.999. And the current values s0=0.76 and r0=0.54 indicate a large distance from LCDM. As con- straints still remain weak, we expect for data with higher precision to provide more consistent fitting results in fu- ture. We also hope theses statefinder parameters can be determined more exactly and shed light on the nature of the cosmological models. −9, β = 0.025+0.008 VII. ACKNOWLEDGMENTS We are very grateful to the anonymous referee for his valuable comments that greatly improve this paper. And we are also grateful to Xin Zhang for his helpful suggestions. Z. L. Yi would like to thank Qiang Yuan and Jie Zhou for their kind help. This work was sup- ported by the National Science Foundation of China (Grants No.10473002 and 10273003), the Scientific Re- search Foundation for the Returned Overseas Chinese Scholars. 8 [1] A. G. Riess, A. V. Filippenko, P. Challis, A. Clocchiattia, A. Diercks, P. M. Garnavich, R. L. Gilliland, C. J. Hogan, S. Jha, K. R. Kirshner, B. Leibundgut, M. M. Phillips, D. Reiss, B. P. Schmidt, R. A. Schommer, R. C. Smith, J. Spyromilio, C. Stubbs, N. B. Suntzeff and J. Tonry, Astron. J. 116, 1009 (1998). [2] S. Perlmutter, G. Aldering, G. Goldhaber, R. A. Knop, P. Nugent, P. G. Castro, S. Deustua, S. Fabbro, A. Goo- bar, D. E. Groom, I. M. Hook, A. G. Kim, M. Y. Kim, J. C. Lee, N. J. Nunes, R. Pain, C. R. Pennypacker, R. Quimby, C. Lidman, R. S. Ellis, M. Irwin, R. G. McMa- hon, P. Ruiz-Lapuente, N. Walton, B. Schaefer, B. J. Boyle, A. V. Filippenko, T. Matheson, A. S. Fruchter, N. Panagia, H. J. M. Newberg and W. J. Couch, Astrophys. J. 517, 565 (1999). [3] S. M. Carroll, W. H. Press and E. L. Turner, Annu. Rev. Astron. Astrophys. 30, 499 (1992). [4] U. Alam, V. Sahni, T. D. Saini and A. A. Starobinsky, Mon. Not. R. Astron. Soc. 344, 1057 (2003). Stetson, Astrophys. J. 553, 47 (2001). [17] N. W. Halverson, E. M. Leitch, C. Pryke, J. Kovac, J. E. Carlstrom, W. L. Holzapfel, M. Dragovan, J. K. Cartwright, B. S. Mason, S. Padin, T. J. Pearson, M. C. Shepherd and A. C. S. Readhead, Astrophys. J. 568, 38 (2002). [18] C. B. Netterfield, P. A. R. Ade, J. J. Bock, J. R. Bond, J. Borrill, A. Boscaleri, K. Coble, C. R. Contaldi, B. P. Crill, P. de Bernardis, P. Farese, K. Ganga, M. Gia- cometti, E. Hivon, V. V. Hristov, A. Iacoangeli, A. H. Jaffe, W. C. Jones, A. E. Lange, L. Martinis, S. Masi, P. Mason, P. D. Mauskopf, A. Melchiorri, T. Montroy, E. Pascale, F. Piacentini, D. Pogosyan, F. Pongetti, S. Prunet, G. Romeo, J. E. Ruhl and F. Scaramuzzi, Astro- phys. J. 571, 604 (2002). [19] P. Schuecker, R. R. Caldwell, H. Bohringer, C. A. Collins, L. Guzzo, N. N. Weinberg, Astron. Astrophys. 402, 53 (2003). [20] A. Melchiorri, L. Mersini, C. J. Odman and M. Trodden, [5] R. R. Caldwell, R. Dave and R. J. Steinhardt, Phys. Rev. Phys. Rev. D 68, 043509 (2003). Lett. 80, 1582 (1998). [6] V. Sahni and L. M. Wang, Phys. Rev. D 62, 103517 (2000). [7] C. Csaki, M. Graesser, L. Randall and J. Terning, Phys. Rev. D 62, 045015 (2000). [8] K. Freese and M. Lewis, Phys. Lett. B 540, 1 (2002). [9] V. Sahni, T. D. Saini, A. A. Starobinsky and U. Alam, JETP Lett. 77, 201 (2003). [21] R. G. Abraham, K. Glazebrook, P. J. McCarthy, D. Crampton, R. Murowinski, I. Jorgensen, K. Roth, I. M. Hook, S. Savaglio, H. W. Chen, R. O. Marzke and R. G. Carlberg, Astron. J. 127, 2455 (2004). [22] T. Treu, P. Mφller, M. Stiavelli, S. Casertano and G. Bertin, Astrophys. J. Lett. 564, L13 (2002). [23] T. Treu, M. Stiavelli, P. Mφller, S. Casertano and G. Bertin, Mon. Not. Roy. Astron. Soc. 326, 221 (2001). [10] P. Gondolo and K. Freese, Phys. Rev. D 68, 063509 [24] L. A. Nolan, J. S. Dunlop, R. Jimenez and A. F. Heavens, (2003). [11] P. Gondolo and K. Freese, Phys. D hep-ph/0211397. [12] Y. Wang, K. Freese, P. Gondolo and M. Lewis, Astro- phys. J. 594, 25 (2003). Mon. Not. Roy. Astron. Soc. 341, 464 (2003). [25] P. L. Nolan, W. F. Tompkins, I. A. Grenier and P. F. Michelson, Astrophys. J. 597, 615 (2003). [26] Z. L. Yi and T. J. Zhang, Mod. Phys. Lett. A. 22, 41 [13] J. Simon, L. Verde and R. Jimenez, Phys. Rev. D 71, (2007). 123001 (2005). [14] C. J. Odman, A. Melchiorri, M. P. Hobson and A. N. Lasenby, Phys. Rev. D 67, 083511 (2003). [15] D. J. Eisenstein, I. Zehavi, D. W. Hogg, R. Scoccimarro, M. R. Blanton, R. C. Nichol, R. Scranton, H. Seo, M. Tegmark, Z. Zheng, S. Anderson, J. Annis, N. Bahcall, J. Brinkmann, S. Burles, F. J. Castander, A. Connolly, I. Csabai, M. Doi, M. Fukugita, J. A. Frieman, K. Glaze- brook, J. E. Gunn, J. S. Hendry, G. Hennessy, Z. Ivezic, S. Kent, G. R. Knapp, H. Lin, Y . Loh, R. H. Lupton, B. Margon, T. McKay, A. Meiksin, J. A. Munn, A. Pope, M. Richmond, D. Schlegel, D. Schneider, K. Shimasaku, C. Stoughton, M. Strauss, M. SubbaRao, A. S. Szalay, I. Szapudi, D. Tucker, B. Yanny and D. York, Astrophys. J. 633, 560 (2005). [16] W. L. Freedmann, B. F. Madore, B. K. Gibson, L. Fer- rarese, D. D. Kelson, S. Sakai, J. R. Mould, R. C. Ken- nicutt, Jr., H. C. Ford, J. A. Graham, J. P. Huchra, S. M. G. Hughes, G. D. Illingworth, L. M. Macri and P. B. [27] D. N. Spergel, R. Bean, O. Dore, M. R. Nolta, C. L. Bennett, G. Hinshaw, N. Jarosik, E. Komatsu, L. Page, H. V. Peiris, L. Verde, C. Barnes, M. Halpern, R. S. Hill, A. Kogut, M. Limon, S. S. Meyer, N. Odegard, G. S. Tucker, J. L. Weiland, E. Wollack and E. L. Wright. astro-ph/0603449. [28] Y. Wang and P. Mukherjee, Astrophys. J. 650, 1 (2006). [29] A. G. Cohen, D. B. Kaplan and A. E. Nelson, Phys. Rev. Lett. 82, 4971 (1999). [30] X. Zhang, Int. J. Mod. Phys. D 14, 1597 (2005). [31] A. K. D. Evans, I. K. Wehus, O. Gron and O. Elgaroy, Astron. Astrophys. 430, 399 (2005). [32] V. Gorini, A. Kamenshchik and U. Moschella, Phys. Rev. D 67, 063509 (2003). [33] X. Zhang, Phys. Lett. B 611, 1 (2005). [34] S. Nesseris and L. Perivolaropoulos, Phys. Rev. D 70, 043531 (2004). [35] J. L. Tonry, B. P. Schmidt, B. Barris, P. Candia, P. Chal- lis, A. Clocchiatti, A. L. Coil, A. V. Filippenko, P. Gar- navich, C. Hogan, S. T. Holland, S. Jha, R. P. Kirshner, K. Krisciunas, B. Leibundgut, W. D. Li, T. Matheson, M. M. Phillips, A. G. Riess, R. Schommer, R. C. Smith, J. Sollerman, J. Spyromilio, C. W. Stubbs and N. B. Suntzeff, Astrophys. J. 594, 1 (2003). [36] B. J. Barris, J. Tonry, S. Blondin, P. Challis, R. Chornock, A. Clocchiatti, A. Filippenko, P. Garnavich, S. Holland, S. Jha, R. Kirshner, K. Krisciunas, B. Lei- 9 bundgut, W. D. Li, T. Matheson, G. Miknaitis, A. Riess, B. Schmidt, R. C. Smith, J. Sollerman, J. Spyromilio, C. Stubbs, N. Suntzeff, H. Aussel, K. C. Chambers, M. S. Connelley, D. Donovan, J. P. Henry, N. Kaiser, M. C. Liu, E. L. Martin and R. J. Wainscoat, Astrophys. J. 602, 571 (2004).
astro-ph/9808019
1
9808
1998-08-03T21:47:23
Magnetic field estimation in Cyg X-3's jet
[ "astro-ph" ]
Multi-wavelength photometric observations of Cygnus X-3 were carried out at 18 cm through to 450 um, complemented by X-ray (2-10 keV) observations. The system was mildly active with cm fluxes at 150 -- 250 mJy. We find the spectrum to be flat with a spectral index of zero. Using a modified Wolf-Rayet wind model, and assuming emission is generated in synchrotron emitting jets from the source, we find an upper-limit to the magentic field of 20 G at a distance of $5 \times 10^{12}$ cm is required.
astro-ph
astro-ph
Magnetic field estimation in Cyg X-3's jet R.N. Ogleya, S.J. Bell Burnella, R.P. Fenderb, G.G. Pooleyc, E.B. Waltmand, M. van der Klisb aDepartment of Physics, The Open University, Walton Hall, Milton Keynes, MK7 6AA, UK bAstronomical Institute 'Anton Pannekoek', University of Amsterdam, 1098 SJ Amsterdam, The Netherlands cMRAO, Cavendish Laboratory, University of Cambridge, CB3 0HE, UK dRemote Sensing Division, Naval Research Laboratory, Code 7210, Washington, DC 20375-5351, USA Abstract Multi-wavelength photometric observations of Cygnus X-3 were carried out at 18 cm through to 450 µm, complemented by X-ray (2-10 keV) observations. The system was mildly active with cm fluxes at 150 -- 250 mJy. We find the spectrum to be flat with a spectral index of zero. Using a modified Wolf-Rayet wind model, and assuming emission is generated in synchrotron emitting jets from the source, we find an upper-limit to the magnetic field of 20 G at a distance 5 × 1012 cm is required. 1. Introduction / History Cygnus X-3 is an unusual binary: a neutron star (or black hole) and Wolf-Rayet-like secondary in a 4.8 hr period. This system flares at sporadic intervals, with various flux increases, but generally returns to a quiescent level of ∼100 mJy. The system has bipolar jets, and it is generally assumed these jets are composed of synchrotron-emitting electrons. Because of the wind, the synchrotron emission, which is by far the dominant emission at radio wavelengths, becomes optically thin at increasing distance with increasing wavelength. The sub-mm emission is relatively unexplored; Fender et al. [1995] first detected the emission, and found excessively high fluxes compared with an extrapolation from the radio. The cm wavelength data has been analysed by Waltman et al. [1996]. Behaviour at these wavelengths is due to synchrotron emission in a modified Wolf-Rayet wind. We use these previous observations as a starting point for our models. 2. Observations All data were taken during 1997 September 27, MJD 50717, and observations were arranged in order to observe the source as close in time as possible. Fig. 1 shows the long-term variation of Cyg X-3 from the Green Bank Interferometer (GBI) programme, from MJD 50400 to MJD 50860 (1996 November -- 1998 February), with an insert showing the variability of the source around the time of our SCUBA observations. One can see that, although a couple of major flares occurred (MJD 50485, 50610), these have no effect on our data. The system seems to have undergone a minor flare event around MJD 50700, and was experiencing a state of unrest during our observations, as it returned to quiescence. Fig. 2 shows the cm variability around the time of the SCUBA observations. The top panel is from the Ryle telescope at 2.0 cm and the bottom panel is from the GBI at 3.6 cm and 13 cm. The time of 1 Fig. 1. Flux history of Cyg X-3 from 1996 November to 1998 February at 13 cm. An insert from MJD 50700 to MJD 50770 is shown; our data was taken on MJD 50717. Although no major flares immediately precede the observations, the system appears to be relaxing after a minor flare -- leading to small-scale variations. the SCUBA observations is shown by the arrow in both plots. Fig. 3 shows the sub-mm variability from SCUBA. The top panel (a) is 850 µm, the bottom panel (b) is 2.0 mm. The 450 µm datum is not shown because poor weather prevented a reliable detection (our formal mean for the 120 minutes of integration at this wavelength is 80 ± 77 mJy). Because of the variability inherent in the source, the spectrum is not apparent. We use the mean flux densities over the whole of the observing times shown in figures 2 and 3 to produce an average spectrum, detailed in table 1 and Fig. 4. Errors are 1σ deviations from the mean. The datum at 450 µm is included for reference. Table 1 Time-averaged flux densities Telescope Wavelength Flux (mJy) Green Bank Interferometer 13 cm 120 ± 10 3.6 cm 200 ± 40 Ryle 2.0 cm 186 ± 32 JCMT SCUBA 2.0 mm 144 ± 18 850 µm 150 ± 46 450 µm 80 ± 77 2 Fig. 2. Cyg X-3 photometry at radio wavelengths. The top panel is Ryle telescope data at 2.0 cm. The bottom panel is from the GBI, square points are at 3.6 cm and triangles are at 13 cm. The observations by SCUBA are shown by the arrow. Fig. 3. Photometry at mm and sub-mm wavelengths. The top panel is at 850 µm, and the bottom panel is at 2.0 mm. The best fit including all the data is shown by a solid line and has a zero spectral index (within errors). We favour a model in which the electrons in the jet are emitting synchrotron radiation, and in which the wind becomes optically thin to the longer wavelengths at larger distances from the central source. Using a stellar wind model with the parameters given in Waltman et al. [1996], the radii at which emission becomes optically thin are 1.5 × 1014, 6 × 1013, 4 × 1013 and 5 × 1012 cm for 13, 3.6, 2.0 cm and 850 µm respectively. We assume material in the jet is observed at these, and increasing, distances from the core as the surrounding material becomes optically thin. Electrons emitting at 850µm would take 2000 s to travel a distance of 5 × 1012 cm (assuming a jet velocity of 0.3 c), and this places a lower-limit to their age. Since no spectral change occurs at this wavelength, we conclude that the high energy electrons have not aged in this time, placing an upper-limit on the magnetic field of ≤ 20 G at that distance. A higher magnetic field would create a spectral change at the shorter wavelengths. 3 Fig. 4. Spectrum from the time averaged data. The solid line is the best fit to all the wavelengths. References Waltman, E.B., Foster, R.S., Pooley, G.G., Fender, R.P., Ghigo, F.D., 1996, AJ, 112, 2690 Fender R.P., Bell Burnell, S.J., Garrington, S.J., Spencer, R.E., Pooley, G.G., 1995, MNRAS, 290, L65 4
astro-ph/0101137
1
0101
2001-01-09T18:14:37
Millimeter observations of radio-loud active galaxies
[ "astro-ph" ]
In order to study the nature of the far-infrared emission observed in radio-loud active galaxies, we have obtained 1.2 mm observations with the IRAM 30 m telescope for a sample of eight radio-loud active galaxies. In all objects we find that the 1.2 mm emission is dominated by non-thermal emission. An extrapolation of the non-thermal radio spectrum indicates that the contribution of synchrotron emission to the far-infrared is less than 10% in quasars, and negligible in the radio galaxies. The quasars in the sample show signs of relativistic beaming at millimeter wavelengths, and the quasar 3C334 shows evidence for strong variability.
astro-ph
astro-ph
A&A manuscript no. (will be inserted by hand later) Your thesaurus codes are: ASTRONOMY AND ASTROPHYSICS Millimeter observations of radio-loud active galaxies Ilse M. van Bemmel1,2 and Frank Bertoldi3 1 European Southern Observatory, Karl-Schwarzschildstr. 2, D -- 85748 Garching bei Munchen 2 Kapteyn Astronomical Institute, P.O. Box 800, NL -- 9700 AV Groningen 3 Max-Planck-Institut fur Radioastronomie, Auf dem Hugel 69, D-53121 Bonn 1 0 0 2 n a J 9 1 v 7 3 1 1 0 1 0 / h p - o r t s a : v i X r a Received date; accepted date Abstract. In order to study the nature of the far-infrared emission observed in radio-loud active galaxies, we have obtained 1.2 mm observations with the IRAM 30 m tele- scope for a sample of eight radio-loud active galaxies. In all objects we find that the 1.2 mm emission is dominated by non-thermal emission. An extrapolation of the non- thermal radio spectrum indicates that the contribution of synchrotron emission to the far-infrared is less than 10% in quasars, and negligible in the radio galaxies. The quasars in the sample show signs of relativistic beaming at millimeter wavelengths, and the quasar 3C 334 shows evidence for strong variability. Key words: galaxies: active -- galaxies:photometry -- quasars: general -- infrared: general -- infrared: galaxies -- radio continuum: galaxies 1. Introduction Photometry from radio to infrared wavelengths has shown that double-lobed radio-loud quasars show deep minima in their spectral energy distributions at millimeter wave- lengths (Antonucci et al. 1990), which indicates that the radio emission arises from components which cannot be related to the infrared emission. However, the nature of the far-infrared emission observed in double-lobed radio sources remains unclear. Most evidence (Haas et al. 1998, Polletta et al. 2000, and references therein) points at a thermal nature of the infrared emission, although a possi- ble contribution from relativistically beamed, non-thermal emission has not yet been carefully determined. Relativis- tic beaming can play an important role in the infrared emission of quasars, in which the relativistic jet is oriented closer to the line of sight (Barthel 1989). In radio-galaxies however, beaming is thought to have a negligible effect (Hoekstra et al. 1997). The significance of beaming may be estimated by ex- trapolating the radio core spectra to infrared wavelengths. offprint Send [email protected]) requests to: Ilse van Bemmel (bem- The non-thermal emission from the radio lobes can be safely neglected, as these have no bulk relativistic motions and thus show no beaming. However, the exact shape of the core spectrum is unknown at high frequencies. In order to improve the quality of the extrapolation of the non-thermal spectrum, we measured 1.2 mm con- tinuum fluxes for eight 3CR objects, of which seven have been observed with ISOPHOT on board ISO (Lemke et al. 1996; Kessler et al. 1996) and with the NRAO Very Large Array (VLA) (van Bemmel et al. 2000). We here compare the 1.2 mm fluxes with the integrated radio and infrared data, and for the quasars in the sample, with their radio core spectra. The general properties of our objects are listed in Ta- ble 1. The sample consists of three radio-loud quasars (QSR), two broad-line radio galaxies (BLRG), one narrow- line radio galaxy (NLRG) and the radio structure 3C 59. The latter was previously misidentified with the Seyfert 1 galaxy RBS 0281. The maps of Meurs & Unger (1991) show three components: a north-western source (the origi- nal 3C 59), a central source associated with RBS 0281, and a south-eastern source. 3C 59 has no optical counterpart, so it is either a lobe of RBS 0281, or a background core- dominated quasar. With ISOPHOT only RBS 0281 was observed. We shall refer to the Seyfert galaxy as RBS 0281, and to the north-eastern hotspot/background object as 3C 59. 2. Observations and data reduction 2.1. IRAM 30 m Continuum observations at 1.2 mm (250 GHz) were ob- tained between 12 and 15 December 1999 with the Max- Planck Millimeter Bolometer (MAMBO; Kreysa et al. 1998) at the IRAM 30 m telescope on Pico Veleta, Spain (Baars et al. 1987). MAMBO is a 37-element bolometer array, sensitive between 190 and 315 GHz. The half peak sensitivity range is 210 -- 290 GHz, with an effective band- pass center, somewhat dependent on the spectral slope of the observed emission, at 250 GHz. The beam for the feed horn of each bolometer is matched to the telescope beam of 10.6′′, and the bolometers are arranged in a hexagonal 2 van Bemmel & Bertoldi: Millimeter observations of radio-loud AGN Name 3C 33.1 3C 59 RBS 0281 3C 67 3C 277.1 3C 323.1 3C 334 3C 460 IAU B1950 0106+729 0204+293 0204+293 0221+276 1250+568 1545+210 1618+177 2318+235 z type L178 [W/Hz] 0.181 BLRG -- ? 0.110 Sey1 0.310 BLRG QSR 0.321 0.264 QSR 0.555 QSR 0.268 NLRG 26.89 24.01 -- 27.28 27.24 27.13 27.88 27.08 size [′′] 227 ∼ 4 ∼ 1 2.5 1.7 69 58 8 Table 1. Basic properties of the observed objects. Types are: NLRG = narrow-line radio galaxy, BLRG = broad line radio galaxy, QSR = radio quasar, Sey1 = Seyfert 1, ? = unknown. We adopt the most common classifica- tion in the literature. L178 assumes H0 = 75 km s−1 Mpc−1, ∝ ν−1. The last column lists the size of the q0 = 0.5 and Fν radio structure associated with the object, only the central component size is given for RBS 0281. pattern with a beam separation of 22′′. Observations were made in standard on-off mode, with 2 Hz chopping of the secondary reflector by 32′′. The pointing accuracy is typi- cally 2′′. The target was centered on the central bolometer of the array, and after each 10 seconds of integration, the telescope was nodded so that the previous "off" beam be- comes the "on" beam. Each scan of twelve 10 second sub- scans lasts 3 minutes, of which 1 minute integration falls on the sources, 1 minute off source, and 1 minute is used to move the telescope and start the integration. Typically 15 to 20 of such scans were performed for each object. Gain calibration was performed using observations of Mars, Uranus, and Ceres, resulting in a flux calibration factor of 12500 counts per Jansky, which we estimate to be accurate to 15%. A sky opacity correction factor was measured every 2 hours through total power sky dips. The data were analyzed using the MOPSI software package (Zylka 1998). For each bolometer the temporally correlated variation of the sky signal (sky-noise) was com- puted using the signals of neighbouring bolometers. The correlated noise is iteratively determined for each channel and subtracted. Due to unintentional mispointing, for 3C 59 and RBS 0281 the targeted positions differ from the radio po- sitions in the maps of Meurs & Unger (1991, see also Ta- ble 2). A signal is measured at the targeted position to- ward RBS 0281, which however is 10′′ off the radio peak, so that the millimeter flux from the peak radio position could well be higher than observed. Toward 3C 59, a 7 mJy signal was picked up by an off-center channel, closely cor- responding to the position of component D in the Meurs & Under radio map. Since the chances of observing a background source are very small, we assume that the off- center channel detects 3C 59, but also here the peak flux Object 3C 59 Position of RA (2000) DEC (2000) MAMBO target radio peak detection RBS 0281 MAMBO target radio peak 02 07 09.6 02 07 10.1 02 07 09.7 02 07 02.3 02 07 02.2 29 31 24 29 31 45.1 29 31 44 29 30 55.1 29 30 46.8 Table 2. Observed positions for 3C 59 and RBS 0281: cen- tral channel position and the peak flux positions in the Meurs & Unger (1991) maps are given. For 3C 59 we also list the position at which the signal is detected. could be higher than the observed flux. Since the MAMBO fluxes toward RBS 0281 and 3C 59 are uncertain, we will not make use of them in the analysis. They will be re- measured in a future observing campaign. 2.2. Radio and infrared data We collected radio fluxes for our objects from the NASA/IPAC Extragalactic Database (NED). Where available, we adopt those given at 178 MHz, 1.4 GHz, 2.7 GHz, and 4.9 GHz. For the three quasars additional fluxes from the unresolved core are available at 4.9, 15, 25 and 43 GHz from a previous programme (van Bemmel et al. 2000, hereafter BBG). For RBS 0281 only one radio point was found. The infrared data were obtained with ISOPHOT, using the P1, P2, C100 and C200 detectors in raster-mapping mode. A detailed description of the data reduction is given in BBG, including a list of the resulting flux densities. ISOPHOT data are available at 60, 90 and 160 µm for all objects except 3C 59. For 3C 33.1 and RBS 0281 we have additional ISOPHOT data at 12 and 25 µm. For 3C 323.1 an upper limit at 10 µm was obtained by Rieke & Low (1972). 3. Results Table 3 lists the radio and millimeter flux densities. The radio fluxes are integrated fluxes, which include the lobes and other extended structures. The millimeter fluxes only include the part of the object within the IRAM 30 m tele- scope beam. The resulting spectral energy distributions (SED) are plotted in Fig. 1, along with the radio flux den- sities of the unresolved core emission for the quasars. The 1.2 mm flux densities are always lower than the infrared and the integrated radio flux densities, consistent with observations by Antonucci et al. (1990). They found that quasars show a minimum in their SEDs at millimeter wavelengths. We obtained similar results for radio galax- ies. The millimeter flux densities are even lower than the fluxes expected from an extrapolation of the integrated ra- dio spectra, which we show as dashed lines in Fig. 1. For van Bemmel & Bertoldi: Millimeter observations of radio-loud AGN 3 Name 178 MHz 1.4 GHz 2.7 GHz 3C 33.1 3C 59 RBS 0281 3C 67 3C 277.1 3C 323.1 3C 334 3C 460 (Jy) 11.4±0.9 5.3±1.0 -- 9.0±1.1 8.9±0.9 10.0±1.3 10.9±0.8 8.1±1.0 (Jy) -- 1.60±0.03 0.026±0.002 3.10±0.09 2.50±0.09 2.45±0.12 2.20±0.11 1.60±0.08 (Jy) -- -- -- -- 1.56±0.06 1.50±0.08 1.00±0.05 0.90±0.05 4.9 GHz (mJy) 820±62 670±87 -- 860±113 880±90 840±112 500±76 440±59 250 GHz (mJy) 6±2 (7±1) (> 3) 10.3±0.7 20.6±1.5 13.5±0.9 <4 4.3±1.2 Table 3. NED archive radio fluxes Sν and newly obtained MAMBO 250 GHz fluxes for our sample. Additional data can be found in van Bemmel et al. (2000). Brackets indicate an off-center detection. this extrapolation we fitted a power law to the radio data with ν > 1 GHz. In cases where a flux was available at only one frequency, we use the average slope of the other objects, excluding 3C 59 and RBS 0281 though. Half of the objects in our sample have radio sizes larger than the 10.6′′ beam of the IRAM 30 m telescope at 1.2 mm. In these cases a direct comparison with the radio data may not be meaningful, unless we assume that the ra- dio lobes are not contributing to the millimeter emission (see section 4.3). For the quasars, core radio fluxes are available at even higher resolution (1 -- 2′′). The core radio SED shows no obvious relation to the millimeter fluxes, except maybe for 3C 323.1, where the core and millimeter fluxes show a typical self-absorbed synchrotron spectrum. To test whether the millimeter flux could arise from a cold, thermal component that also gives rise to the 160 µm emission, we fit an optically thin grey-body spectrum to the far-infrared flux, using a dust emissivity index β = 1.6. This value is an average of observed values in nearby active and normal galaxies. To be conservative, we adopted dust temperatures of 20 K, which is even colder than the coldest dust found in active galaxies with ISO (e.g. Siebenmorgen et al. 1999). The temperature is typical for dust in Galac- tic clouds and cirrus. The dust masses implied by such a cold component range from a few 107 to several 108 M⊙. From the extrapolated grey-body spectra it appears that the millimeter flux is dominated by non-thermal emission. In 3C 460, e.g., the 1.2 mm flux could be dominated by thermal dust emission if β < 1.4 or T < 16, but then the implied dust mass would be larger than 5 × 108 M⊙. Such a large dust mass is hard to reconcile with 3C 460 being an elliptical galaxy. There is only circumstantial evidence for a very cold dust component (T < 15 K) in galaxies. Starbursts and active galaxies do not show very cold dust emission at millimeter wavelengths. 3C 334 is probably a variable millimeter source. Previ- ous observations with the Owens Valley Radio Observa- tory found a 3 mm flux density of 37 ±2 mJy (van Bemmel ∝ ν−1, a 1.2 mm flux of 15 mJy et al. 1998). Assuming Fν would be expected, much higher than our upper limit of Name 3C 33.1 3C 59 RBS 0281 3C 67 3C 277.1 3C 323.1 3C 334 3C 460 QSR average RG average log L250 α178 1.4 α1.4 4.9 23.61 -- (>22.86) 24.33 24.67 24.30 <24.46 23.82 24.49 23.92 -- -- -0.61 -0.64 -- -0.52 -0.62 -0.68 -0.78 -0.79 -0.69 -0.65 -- -1.02 -0.83 -0.85 -1.18 -1.03 -0.96 -1.03 α4.9 250 -1.28 (-1.18) (>-0.42)∗ -1.15 -0.97 -1.07 <-1.25 -1.2 -1.02 -1.21 Table 4. Monochromatic 250 GHz luminosities (in W Hz−1) and radio spectral indices. Brackets indicate an off-center detection. (∗) for RBS 0281 α1.4 250 is given instead of α4.9 250. 3 mJy. 3C 334 is known to be variable at 4.85 GHz, with variations of order 10% over a few decades (van Bemmel et al. 1998). Variability tends to increase at higher fre- quencies, but one would not expect a dramatic change in flux such as implied by the OVRO and MAMBO measure- ments. 3.1. Luminosities and spectral indices Adopting H0 = 75, q0 = 0.5, and kΛ = 0, we computed lu- minosity densities in order to compare the millimeter lu- minosities of quasars and radio galaxies (Table 4). 3C 59 is excluded, because its redshift has not yet been deter- mined. Although we find that the infrared luminosities are comparable between the classes, there is evidence that the quasars are brighter at 1.2 mm than the radio galaxies. At 178 MHz the quasars are only marginally brighter. The sample is too small to draw firm conclusions, but it appears that the stronger millimeter emission from quasars may be due to a beamed component. When com- paring the integrated flux at 178 MHz and 1.4 GHz, the quasars are a factor 1.5 brighter than the radio galaxies. 4 van Bemmel & Bertoldi: Millimeter observations of radio-loud AGN Fig. 1. Rest frame spectral energy distributions of the objects observed with MAMBO. For 3C 59 we assume z = 0. Filled circles show the total integrated radio fluxes, open circles show the unresolved core radio fluxes, the filled diamond the MAMBO 1.2 mm fluxes, and open diamonds are ISOPHOT fluxes. The solid lines represent grey-body T = 20 K spectra, matching the 160 µm fluxes. The dashed lines are power law fits to > 1 GHz integrated radio data. van Bemmel & Bertoldi: Millimeter observations of radio-loud AGN 5 In case of isotropic synchrotron emission over the whole radio spectrum, this factor should be constant. However, it appears to increase toward higher frequencies; at 250 GHz the quasars are about three times brighter than the ra- dio galaxies. This indicates that we observe an additional component, which is unlikely to be due to dust emission, since that would be optically thin and therefore also vis- ible in the radio galaxies. For synchrotron emission, the excess can only be beamed emission, which causes a natu- ral anisotropy. It would be desirable to confirm this trend with a larger sample. There is clear evidence of spectral steepening of the radio spectrum towards millimeter wavelengths. The av- erage spectral index for the entire sample ranges from −0.7 at the lowest frequency to −1.2 at millimeter wavelengths (Table 4). Spectral steepening occurs in most objects in our sample, irrespective of their size or the resolution of the observations. 4. Discussion 4.1. Nature of the millimeter and far-infrared emission We find that the 1.2 mm flux densities are well below those expected from an extrapolation of the total radio flux den- sities. If cold dust gives rise to the millimeter emission, then the millimeter flux densities are expected to lie above the extrapolated radio spectra, and the grey-body extra- polation should fit the observed flux. In our data the oppo- site is true, and therefore the millimeter emission is proba- bly dominated by synchrotron emission, except maybe for 3C 460, where this case is not so clear. Since the far-infrared fluxes are more than an order of magnitude higher than the millimeter fluxes, the far- infrared emission is likely to arise from dust. The clear minimum of the SED at millimeter wavelengths excludes that the far-infrared emission is the continuation of the radio synchrotron spectrum of the radio lobes or the core. If the far-infrared emission were due to synchrotron radiation, it would have to arise from a very young elec- tron population. The younger the electrons, the closer they must be to the core, so that their emission would be even more strongly beamed than synchrotron emission at mil- limeter wavelengths. As a consequence, the quasars should be brighter than radio galaxies at infrared wavelengths, for which there is no evidence in our data. Instead we find that in the infrared the quasars and radio galaxies are comparably bright, while at 1.2 mm the quasars are clearly brighter than the radio galaxies. The possible contribution of relativistically beamed synchrotron emission to the far-infrared emission is es- timated by extrapolating the radio SED. This yields on average a less than 10% contribution in quasars, and a less than 1% contribution in radio galaxies. Although in 3C 323.1 the core radio spectrum would extrapolate well to the observed far-infrared fluxes, the 1.2 mm emission falls below this extrapolation, indicating a turnover in the core spectrum, which excludes a significant contribution of non-thermal emission to the far-infrared. An earlier study of the infrared emission from radio- loud active galaxies indicated that quasars are signifi- cantly brighter at 60 µm than radio galaxies (Heckman et al. 1992). This does not contradict what we find from our sample. The objects in the Heckman sample have an average redshift ∼ 0.5 − 0.9, thus in their rest frame the emission emerges at ∼ 30 − 40 µm. If the infrared emission arises from a circumnuclear torus, the dust is likely to be optically thick up to 60 µm (Pier & Krolik 1992, Granato & Danese 1994), and the emerging flux depends on the ori- entation of the torus. According to these models, the ob- served flux could vary by orders of magnitude, depending on the torus' optical depth. However, ISOPHOT studies (BBG) show no conclusive evidence that quasars are in- trinsically brighter in the far-infrared than radio galaxies, when they are matched in radio power and redshift. 4.2. Relativistic beaming in quasars Unified models for radio-loud AGN (Urry & Padovani 1995, Barthel 1989) suggest that quasars are oriented with their jets closer to the line of sight than radio galaxies. This implies that any beamed component is more evi- dent in quasars, but also that isotropic emission, such as optically thin dust emission, should not differ among the types. Our observations confirm this picture, showing that quasars are more luminous at millimeter wavelengths, where we expect relativistic beaming, whereas they do not differ in their far-infrared luminosity, which is due to isotropic cold dust emission, and therefore unaffected by beaming. The luminosity ratio QSR/RG increases with frequency, which can only naturally be explained by beam- ing. If this trend is confirmed in larger samples, millimeter observations in combination with radio observations could provide a direct measure of the orientation of a source. E.g. for a quasar and a radio galaxy of comparable 178 MHz power, the radio galaxy provides the unbeamed radio SED that can be subtracted from the quasar SED. The remain- ing emission is then due to beaming, and the strength of this beamed component depends directly on the viewing angle of the source. 4.3. Hotspot and lobe emission For three of our objects, the radio structures are larger than the beam size of the IRAM 30 m telescope, so that the hotspots fall outside the central bolometer channel. We made no attempt to observe the hotspots separately. The radio fluxes for all objects are integrated fluxes, including the lobe emission. The MAMBO fluxes are also integrated fluxes for the small objects, but core fluxes for the objects with larger radio sizes than the IRAM beam. If 6 van Bemmel & Bertoldi: Millimeter observations of radio-loud AGN the millimeter emission would be dominated by the lobe emission in all objects, there should be a clear turnover in the SEDs of the large objects, where the lobes are not observed with MAMBO. On the other hand, if the core dominates the millimeter emission in all objects, the SEDs should be comparable, irrespective of object radio size. We observe no clear difference between the radio -- millimeter SED of large and small objects, which seems to indicate that the core is dominating the millimeter emission. This does not imply that the lobes do not emit any millimeter emission, e.g. in Cygnus A the hotspots have been clearly detected with SCUBA at 850 µm (Robson et al. 1998). However, there might be a relation between the lumi- nosity density of the lobes at 1.2 mm and the size of the radio structure. Small radio sources are known to have much flatter spectra (Murgia et al. 1999) and thus can have much stronger lobes. In our MAMBO observations, all small sources are unresolved and thus we cannot tell which component is dominating. If the lobes dominate in small objects and the core dominates in larger ones, the expected difference in spectral indices will not be visible. We always observe the dominating regions, i.e. the lobes in the small radio sources and only the core in the larger ones. Judging from the radio core fluxes in 3C 277.1, the core is not the dominant 1.2 mm source here. The same can be true for 3C 67 and 3C 460, which are also small radio sources. The lobe emission will dilute the amount of beaming observed in small objects, thus for a proper es- timate of the amount of beaming the objects should have comparable radio sizes and be larger than the IRAM 30m telescope beam. 4.4. Variability We find evidence for variability in 3C 334. Since we have only one observation for each object, variability cannot be ruled out for the other objects. Previous SCUBA observa- tions of the other quasars (BBG) are consistent with the MAMBO detections, confirming the spectral steepening and the thermal nature of the infrared emission. For the radio galaxies there is remarkably little dispersion in the observed spectral slopes. Thus, variability must be small in all other objects. -- An extrapolation of the non-thermal radio spectrum to far-infrared wavelengths shows that the contribution of non-thermal emission at 160 µm is less than 10%. -- There is evidence that at 1.2 mm the quasar emission is stronger, and therefore possibly more beamed, than that from radio galaxies. -- The far-infrared luminosity on the other hand, does not differ between quasars and radio galaxies. Most likely because it arises from optically thin thermal emission, but optical thickness effects can play a role up to 100 µm in the rest frame of the objects. -- Radio and millimeter observations provide an inter- esting test for unification models of radio-loud active galaxies, in that they can provide a measure of the amount of beamed emission. Millimeter observations, in combination with radio observations, of a larger sample of radio-loud active galaxies are needed to ver- ify this. Acknowledgements. Thanks to Bob Fosbury and Peter Barthel for their motivation and help on the manuscript. Thanks to Alessandra Bertarini for assisting with the observations, and to the referee, Neal Jackson, for comments which greatly im- proved the manuscript. Special thanks to the MPIfR bolometer team for providing MAMBO and support, and to R. Zylka for writing the MOPSI data reduction package. NED is operated by the Jet Propulsion Laboratory, Cal- ifornia Institute of Technology, under contract with the Na- tional Aeronautics and Space Administration. ISO is an ESA project with instruments funded by ESA Member States (es- pecially the PI countries: France, Germany, the Netherlands and the United Kingdom) and with the participation of ISAS and NASA. References Antonucci R., Barvainis R., Alloin D., 1990, ApJ 353, 416 Baars J.W.M., Hooghoudt B.G., Mezger P.G., de Jonge M.J., 1987, A&A 175, 319 Barthel P.D., 1989, ApJ 336, 606 Granato G.L., Danese L., 1994, MNRAS 268, 235 Heckman T.M., Chambers K.C., Postman M., 1992, ApJ 391, 39 Hoekstra H., Barthel P.D., Hes R., 1997, A&A 319, 757 Kessler M.F., Steinz J.A., Anderegg M.E., et al., 1996, A&A 315, L27 Kreysa E., Gemuend H.-P., Gromke J., et al., 1998, SPIE 3357, 5. Conclusions 319 The main conclusions, drawn from our continuum obser- vations at 1.2 mm of eight radio-loud active galaxies, can be summarized as follows. -- The millimeter emission is dominated by non-thermal processes, whereas the far-infrared emission must be thermal. -- Any thermal contribution to the millimeter emission arising from cold dust is estimated to be less than 15%, except for 3C 460, where the case is not clear. Lemke D., Klaas U., Abolins J., et al., 1996, A&A 315, L64 Meurs E.J.A., Unger S.W., 1991, A&A 252, 63 Murgia M., Fanti C., Fanti R., et al., 1999, A&A 345, 769 Pier E.A., Krolik J.H., 1992, ApJ 401, 99 Polletta M., Courvoisier T.J.-L., Hooper E.J., Wilkes B.J., 2000, A&A in press, astro-ph 0006315 Rieke G.H., Low F.J., 1972, ApJ 176, L95 Robson E.I., Leeuw L.L., Stevens J.A., Holland W.S., 1998, MNRAS 301, 935 Siebenmorgen R., Krugel E., Chini R., 1999, A&A 351, 495 Urry C.M., Padovani P., 1995, PASP 107, 803 van Bemmel & Bertoldi: Millimeter observations of radio-loud AGN 7 van Bemmel I.M., Barthel P.D., Yun M.S., 1998, A&A 334, 799 van Bemmel I.M., Barthel P.D., de Graauw T., (BBG), 2000, A&A 359, 523 Zylka R., 1998, Pocket Cookbook for MOSPI Software, http://www.iram.es/Telescope/manuals/Datared/pockroo.ps
astro-ph/0606309
1
0606
2006-06-13T12:08:23
VLT/ISAAC Spectra of the H-beta Region in Intermediate-Redshift Quasars II. Black Hole Mass and Eddington Ratio
[ "astro-ph" ]
We derive black hole masses for a sample of about 300 AGNs in the redshift range 0 < z < 2.5. We use the same virial velocity measure (FWHM Hbeta broad component) for all sources which represents a significant improvement over previous studies. We review methods and caveats for determining AGN black hole masses via the virial assumption for motions in the gas producing low ionization broad emission lines. We derive a corrected FWHM measure for the broad component of H-beta that better estimates the virialized line emitting component by comparing our FWHM measures with a sample of reverberated sources with H-beta radial velocity dispersion measures. We also consider the FWHM of the FeII 4570 blend as a potential alternative velocity estimator. We find a range of black hole mass between log M ~ 6 - 10, where the black hole mass M is in solar masses. Estimates using corrected FWHM, as well as FWHM(Fe II) measures, reduce the number of sources with log M > 9.5 and suggest that extremely large M values (log M >~ 10) may not be realistic. Derived Eddington ratio values values show no evidence for a significant population of super-Eddington radiators especially after correction is made for sources with extreme orientation to our line of sight. Sources with FWHM(Hbeta broad component) <~ 4000 km/s show systematically higher Eddington ratio and lower M values than broader lined AGNs (including almost all radio-loud sources).
astro-ph
astro-ph
Astronomy & Astrophysics manuscript no. sulenticetalsubmission3rdprint (DOI: will be inserted by hand later) April 25, 2018 VLT/ISAAC Spectra of the Hβ Region in Intermediate-Redshift II. Black Hole Mass and Eddington Ratio⋆ Quasars J. W. Sulentic1, P. Repetto2,3, G. M. Stirpe4, P. Marziani2, D. Dultzin-Hacyan3, and M. Calvani2 1 Department of Physics and Astronomy, University of Alabama, Tuscaloosa, AL 35487, USA e-mail: [email protected] 2 Osservatorio Astronomico di Padova, INAF, Vicolo dell' Osservatorio 5, 35122 Padova, Italy e-mail: [email protected]; [email protected] 3 Instituto de Astronom´ıa, UNAM, Aptdo. Postal 70-264, M´exico, D. F. 04510, M´exico e-mail: [email protected],[email protected] 4 Osservatorio Astronomico di Bologna, INAF, Via Ranzani 1, 40127 Bologna, Italy e-mail: [email protected] Received / Accepted Abstract. We derive black hole masses for a sample of about 300 AGNs in the redshift range 0 < z < 2.5. We use the same virial velocity measure (FWHM HβBC) for all sources which represents a significant improvement over previous studies. We review methods and caveats for determining AGN black hole masses via the virial assumption for motions in the gas producing low ionization broad emission lines. We derive a corrected FWHM(HβBC) measure for the broad component of Hβ that better estimates the virialized line emitting component by comparing our FWHM measures with a sample of reverberated sources with Hβ radial velocity dispersion measures. We also consider the FWHM of the Feiiλ4570 blend as a potential alternative velocity estimator. We find a range of black hole mass between logMBH ∼ 6.0 -- 10.0, where MBH is in solar masses. Estimates using corrected FWHM(Hβ), as well as FWHM(Feii) measures, reduce the number of sources with log MBH > 9.5 and suggest that extremely large MBH values (log MBH>∼ 10) may not be realistic. Derived L/LEdd values show no evidence for a significant population of super-Eddington radiators especially after correction is made for sources with extreme orientation to our line of sight. Sources with FWHM(HβBC) <∼ 4000 km s−1 show systematically higher L/LEdd and lower MBH values than broader lined AGNs (including almost all radio-loud sources). Key words. quasars: emission lines -- quasars: general -- line: profiles -- black hole physics 6 0 0 2 n u J 3 1 1 v 9 0 3 6 0 6 0 / h p - o r t s a : v i X r a 1. Introduction Gravitational accretion onto supermassive black holes is generally accepted as the ultimate energy source of Active Galactic Nuclei (AGNs). The last decade has seen a major effort to derive reasonable estimates of black hole masses (MBH) by assuming virialized motions in the broad line emitting gas: MBH = v2rBLR G (1) the full width half maximum measured for a suit- able emission line. The factor f depends on the ge- ometry and details of the kinematics (Krolik, 2001; McLure & Dunlop, 2001; Onken et al., 2004). The use- fulness of the virial assumption for MBH determina- tion is best exploited using reverberation-mapping stud- ies, especially for the Balmer lines (Koratkar & Gaskell, 1991; Peterson & Wandel, 1999; Wandel et al., 1999; Kaspi et al., 2000; Peterson et al., 2004). Kaspi et al. (2000, 2005) derive a relation between Broad Line Region (BLR) distance rBLR and continuum luminosity where v is the velocity dispersion of the emit- ting gas at distance rBLR. The velocity dispersion can be written as v = f FWHM where FWHM is rBLR ∝ (λLλ)α (2) Send offprint requests to: J. W. Sulentic ⋆ Based on observations collected at the European Southern Observatory, Chile. Proposal ref.: ESO 072.B-0338(A) where the exponent α is constrained between 0.5 and 1.0, and is most likely ≈ 0.6 -- 0.7 for broad Hβ (HβBC) and optical continuum luminosity. 2 Sulentic et al.: Hβ in Intermediate z Quasars: MBH and L/LEdd There are caveats associated with this method. The virial assumption is not likely to be generally valid for the emission line gas in AGNs. It has been known for sev- eral decades that different emission lines in a source can show different width and profile shape (e.g., de Robertis, 1985; Sulentic, 1989). The virial assumption implies that the velocity dispersion will steadily decrease with dis- tance from the central black hole ∝ r− 1 2 . The time lag between continuum fluctuations and corresponding emis- sion line responses will therefore anti-correlate with line width. This trend has been confirmed in a few objects (Peterson & Wandel, 2000). Observations however show that profile width and shape depend on the ionization potential. The strongest high-ionization lines (HILs; e.g. Civλ1549) often display blueward asymmetric profiles, or even centroid blueshifts (up to several 1000 km s−1) with respect to the best esti- mates of the rest frame of the source (e.g., Gaskell, 1982; Marziani et al., 1996; Richards et al., 2002; Bachev et al., 2004). Blueshifts like those observed for Civλ1549 are in- dicative of obscuration and radial motions which invali- date the virial assumption. Low ionization lines (LILs) like Hβ are the best can- didates for emission arising from a virialized medium. All or much of the Balmer (and Feii) line emission is thought to arise from an accretion disk or a flattened cloud dis- tribution near the disk (e.g., Collin-Souffrin et al., 1988; Marziani et al., 1996). A major caveat, especially for the Balmer lines, involves the possibility that: (1) there are two or more emission components in a line, and (2) only one of them may arise in a region where the virial assump- tion holds. The Balmer lines do not usually show very large line shifts (i.e., shift ≤ FWHM) although profiles can be very asymmetric. Both red and blue asymmetries are observed for HβBC which is the most studied line because it is relatively unblended and is observable with optical spectrometers up to z ≈ 1 (Osterbrock & Shuder, 1982; Sulentic, 1989; Stirpe, 1990; Sulentic et al., 1990; Corbin, 1991; Corbin & Francis, 1994). The most ambiguous sources from the point of view of MBH determination show FWHM(HβBC) >∼ 4000 km s−1 (Population B, following Sulentic et al., 2000b) and red- shifted profiles and/or red asymmetries. Not all parts of the Hβ profile respond to continuum changes in the same way implying that some of the line emitting gas may be optically thin or, less likely, is not exposed to the variable Hi ionizing radiation (we will return to this issue in §5.2). The broad line profile in Pop. B sources may be due to two distinct emitting regions: (1) an optically thick classical BLR and (2) a broader and redshifted very broad com- ponent that may be optically thin or marginally optically thick to the Lyman continuum, originating in a distinct Very Broad Line Region (VBLR) (Marziani & Sulentic, 1993; Shields et al., 1995; Sulentic et al., 2000c). The red- shift of the VBLR component raises doubts that it arises from virialized gas. A strong BLR response to continuum fluctuations, coupled with a weak or absent response of the VBLR component, can lead to an overestimate of FWHM for the virialized BLR component resulting in an overes- timate of MBH (Wandel et al., 1999; Kaspi et al., 2000; Vestergaard, 2002). Sources with FWHM(HβBC) <∼ 4000 km s−1 (Population A, Sulentic et al., 2000b) should provide more reliable MBH estimates. The HβBC profile is usually well fit with a symmetric function (V´eron-Cetty et al. , 2001; Sulentic et al., 2002) and the BLR emission is thought to arise from a Keplerian disk. The most unreliable Pop. A sources, in a disk emission scenario, should be those observed near face-on where the rotational (i.e. virial) contribution to FWHM(HβBC) is minimal. At least some of the face-on sources may be identified as the so- called "blue outliers" which show a weak and significantly blueshifted [Oiii]λλ4959,5007 lines (Zamanov et al., 2002; Marziani et al., 2003b; Aoki et al., 2005; Boroson, 2005). The inferences in this, and the preceding paragraph have emerged from the Eigenvector 1 (E1) scenario that we have pursued for the past 5+ years (Sulentic et al., 2000a,b, following Boroson & Green (1992)). The results reported in this paper appear to confirm them. There are additional caveats connected with us- ing FWHM(HβBC) for MBH derivations. Line asym- metries can affect MBH estimates for both Pop. A and B sources. The vr displacement of the line cen- troid at fractional intensities (at half maximum, c( 1 2 )) can give useful information about the uncertainty of FWHM(HβBC) measurements. The average displacement value is a few hundred km s−1 with median c( 1 2 ) ≈ 400 km s−1 (Marziani et al., 2003a). This implies that deviations from a symmetric profile can affect MBH estimates by ≈ 40 %. Uncertainties are also intro- duced by: (1) contamination from overlapping/nearby lines such as Feii, Heiiλ4686, and [Oiii]λλ4959,5007 (Osterbrock & Shuder, 1982; de Robertis, 1985; Joly, 1988; Jackson et al., 1991), (2) FWHM measures based on single-epoch observations, (3) low S/N spectra and (4) spectra without Hβ narrow component (HβNC) subtrac- tion. The rBLR - Lλ relation is also not free from un- certainties. Reverberation mapping-based MBH determi- nations are certainly affected by the non-negligible radial extent of the optically thick BLR. The derived rBLR is not a very well defined quantity and α is also somewhat uncertain because of the intrinsic scatter in the correla- tion. Finally, reverberation data does not exist for high luminosity/redshift quasars requiring an extrapolation of the rBLR - Lλ relation in order to estimate MBH for these sources. Uncertainties for MBH derivations using single profile observation of HβBC are estimated to be a factor of 2 -- 3 (at a 1 σ confidence level), but may be as low as 30% if the velocity dispersion of the variable part of HβBC profile is employed as a virial estimator (Peterson et al., 2004). MBH estimates based on the virial relation retain a statis- tical validity considering that AGNs span a 5dex range in MBH (105 M⊙<∼ MBH <∼ 1010 M⊙) and that the relation has now been applied to large samples of objects (∼ 103; McLure & Dunlop, 2004). Sulentic et al.: Hβ in Intermediate z Quasars: MBH and L/LEdd 3 With these considerations in mind, and supported by previous results, we use the virial relation to compute MBH for ≈ 280 AGNs in our E1 sample of low red- shift (0 <∼ z <∼ 1) sources (Marziani et al., 2003b) sup- plemented with 25 intermediate redshift/high luminosity (1 <∼ z <∼ 2.5) quasars. We derive MBH and L/LEdd values in several ways. New VLT-ISAAC data are presented for 9 sources (§2) which supplement data already published for sixteen quasars (Sulentic et al., 2004). Line measures are presented in §4. We use HβBC and Feiiλ4570 line widths in a consistent way over the redshift range 0 <∼ z <∼ 2.5 (a range of 105 in luminosity). We compute black hole mass MBH, Eddington ratio L/LEdd (§5) and we discuss how mass determinations might be improved (§5.2) so that the evolution of MBH and L/LEdd with redshift can be con- sidered (§5.3 and §6). 2. Observations and Data Reduction New intermediate redshift data were obtained between 10/2003-03/2004 in service mode with the infrared spec- trometer ISAAC mounted on VLT1 (ANTU) at the European Southern Observatory. Each spectrum corre- sponds to a wavelength range (IR windows sZ, J, sH) that covers the region of redshifted HβBC and Feiiλ4570 or Feiiλ5130 at least in part. Reduction of quasar spectra and standard stars followed exactly the same procedures described in Sulentic et al. (2004). Wavelength calibration yielded rms residuals of 0.4, 0.6 and 0.9 A in the sZ, J and sH windows, respectively. Absolute flux scales of the spec- tra will be inaccurate because atmospheric seeing almost always exceeded the slit width (≈ 0".6) resulting in sig- nificant light loss. Table 1 summarizes the new observations and the ba- sic format is given below the table. All sources come from the Hamburg-ESO (HE) quasar survey, which is a flux limited (with limiting mB ≈ 17.5), color-selected survey (Wisotzki et al., 2000 ). Column 2 of Table 1 lists the blue apparent magnitudes from the HE survey papers (Reimers et al., 1996; Wisotzki et al., 2000 ) while Col. 3 lists the source redshift z computed as described in section 3. The brightest sources of the HE at intermediate red- shift were preferentially selected. Col. 4 indicates whether [Oiii]λ5007 was used to compute z as indicative of the source rest frame. The absolute magnitude MB reported in Col. 5 was computed by assuming H0 = 70 km s−1 Mpc−1, and relative energy density ΩM = 0.3 and ΩΛ = 0.7. The k correction was computed for a spectral index a = 0.6 (S ∝ ν−a). Col. 6 gives the ratio of log specific fluxes at 6 cm and 4400 A (log Rk). In most cases only NRAO VLA Sky Survey (NVSS) upper limits are available. Columns 7-12 give details of the observations explained in the foot- notes. The continuum S/N estimate given in Col. 12 was measured using a small region of the spectrum that was as flat and free of line emission as possible. 3. Data Analysis 3.1. Redshift Determination and Rest Frame Corrections Small offsets are present in the wavelength calibration, because the arc lamp frames were obtained in daytime, and therefore usually after grism movement. A correc- tion for these shifts was obtained by measuring the cen- troids of 2 -- 3 OH sky lines against the arc calibration and calculating the average difference, which reached at most 6.5 A or 2.5 pixels in either direction. Rest frame determination for the 9 new sources was usually esti- mated from the Hβ peak redshift (assumed rest frame λ = 4861.33 A). [Oiii]λ5007 (assumed rest frame λ = 5006.85 A) yielded a consistent measurement in only two sources (where the results were averaged. Two sources show no clear detection of [Oiii]λλ4959,5007 while the remaining five show a significant disagreement between [Oiii]λ5007 and HβNC. In these sources (we call the extreme exam- ples of them "blue outliers") it is not advisable to use [Oiii]λ5007 for redshift determination (Zamanov et al., 2002; Marziani et al., 2003b; Aoki et al., 2005; Boroson, 2005). The adopted estimate was used to deredshift the spectra while the dopcor IRAF task applied a (1+z)3 cor- rection to convert observed specific fluxes into rest frame values. Fig. 1 shows the deredshifted spectra. 3.2. Continuum and Feii subtraction Our spectral analysis made use of standard IRAF tasks with the first step involving continuum modelling and sub- traction. Using Chebyshev polynomials of 3rd or 2nd or- der, a reasonably smooth continuum subtraction was ob- tained for all sources. To estimate errors in the continuum assessment introduced by noise, we also defined a mini- mum and a maximum continuum. Continuum fluxes were chosen at about -- 3σ (minimum) and +3σ (maximum), where σ is the noise standard deviation from the most likely continuum choice. Errors in continuum placement defined by difference between the extreme continua and the most probable one were then propagated according to standard error theory. The results of this procedure are consistent with continuum fits employing very simple models (Malkan & Sargent, 1982; Shang et al., 2004). We assumed that the continuum underlying the Hβ spectral regions is due to two components: either a blackbody of temperature 25000◦ K or a power law of slope b = 0.7 (fν ∝ ν−b; assumed to be valid only locally around Hβ). In 5 sources the sole black body component produces a good fit; in 2 the blackbody component is dominant, and only in the remaining 2 cases the power-law alone can pro- vide a good fit. This method has some limitations due to the small spectral bandwidth covered by our spectra, to the relative strength of Feii and to internal reddening ef- fects. Since we did not attempt to change the blackbody temperature nor the slope of the power-law, we adopt the empirical continuum which is visually more accurate. 4 Sulentic et al.: Hβ in Intermediate z Quasars: MBH and L/LEdd The emission blends of Feii were subtracted using the template method based upon the spectrum of I Zw 1 (Boroson & Green, 1992; Marziani et al., 2003a). The strongest Feii blends fall in the wavelenghth ranges 4450 -- 4600 A (blue blend: Feiiλ4570) and 5200 -- 5600 A (red blend: Feiiλ5130). The method includes the following steps: (1) the template intensity is scaled to roughly match the observed spectrum, (2) a Gaussian broadening factor is estimated from FWHM HβBC (3) a careful estimate of minimum and maximum plausible broadening factor is made to set a reliable ±3σ uncertainty and (4) the template intensity is adjusted as necessary after broad- ening. The resultant template was subtracted from the continuum-subtracted spectra. The blue side of the spec- trum including Feiiλ4570 is missing, or not fully covered, in several sources (HE 0946 -- 0500, HE 1003+0149, HE 1017 -- 0009, HE 1249 -- 0648 and HE 1258 -- 0823). In these cases the best template fit was achieved for the red blend and a fixed ratio between the red and blue blends was as- sumed to reproduce Feiiλ4570. Fig. 2 shows the estimated Feii emission (green lines). Feii was subtracted before con- tinuum fitting in sources with strong Feii emission. The [Oiii]λλ4959,5007 lines were measured after Feii subtraction and taking into account the following con- siderations: (1) the flux ratio between [Oiii]λ5007 and [Oiii]λ4959 should be ≈3, (2) both lines should show identical profiles and (3) any HβBC emission underlying [Oiii]λλ4959,5007 is expected to have a smooth shape. Sources HE 1249 -- 0648 and HE 1258 -- 0823 see (Fig. 2) show a small bump at λ ≈ 5016 A which is either an Feii subtraction residual or emission from Heiλ5016 (weak red- shifted [Oiii]λ5007 would be almost unprecedented). After subtraction of HβNC (following Marziani et al., 2003a), the HβBC profile was fit with a high order spline function (IRAF task SFIT). This procedure does not yield a model fit but only an empirical fit that smooths the noise and reproduces the main features and inflections in HβBC. 4. Immediate Results 4.1. Line Measurements and Uncertainties Line Fluxes and Equivalent Widths Table 3 gives line mea- surements for the new VLT spectra with the basic for- mat given in the footnote. Cols. 3 and 4 give equivalent width (EW) measures for HβBC and Feiiλ4570 respec- tively. We evaluated uncertainties associated with the con- tinuum level (derived from the minimum and maximum reasonable continuum estimates) and line flux errors es- timated from the S/N . These estimates were combined quadratically to obtain uncertainties for EW measures. A similar procedure was applied to obtain uncertainty esti- mates for HβNC and [Oiii]λλ4959,5007. The EW uncer- tainty for HβNC was derived from the estimated maxi- mum and minimum possible HβNC component in the Hβ line. The relative error of the HβNC flux can be large (see Table 3) and in some cases an HβNC component may not be present. Fe iiopt width Feii emission is heavily blended so that widths of individual lines must be obtained from the best broadening parameter that was used for the tem- plate fit. This requires that we assume a constant width for all Feii lines which so far appears to be reason- able. FWHM(Feiiλ4570) values derived from the tem- plate broadening factor are reported in Col. 5 of Table 3. Uncertainty estimates for FWHM(Feiiλ4570) were ob- tained by increasing/decreasing the broadening factor un- til we could detect significant changes in the best fit. Simulated data reveal that it is possible to estimate the Feiiλ4570 width up to FWHM(Feiiλ4570) ≈ 6000 km s−1. Due to the very large uncertainty of Feiiλ4570 width de- termination for individual sources, MBH estimates based on FWHM(Feiiλ4570) measures are used mainly for con- firmatory purposes of statistical trends detected with Hβ. HβBC Line Profiles Measurements of FWHM(Hβ) to- gether with other important line parameters like asymme- try index, kurtosis and line centroid at various fractional intensities were derived using a FORTRAN program devel- oped for that purpose. These parameters are the same as defined in several previous papers (Marziani et al., 1996, 2003a; Sulentic et al., 2004) and are given in Table 3. Each line measure is followed in the next Col. by its appropri- ate uncertainty. Cols. 2 and 4 give the Full Width at Zero Intensity (FWZI) and FWHM. Col. 6 gives the asymme- try index (AI) as defined in Sulentic et al. (2004). Col. 8 lists kurtosis values while the remaining part of Table 3 lists measures of the HβBC centroid at various fractional intensities. (in km s−1). All uncertainties represent the 2σ confidence level. The dichotomy in HβBC profile shape (and many other properties) between Population A and B (Sulentic et al., 2002) is seen in the new source measures and in the rest of our higher redshift sample (Sulentic et al., 2004). Redward asymmetries (A.I. >∼ 0.2) are most often found in Pop. B sources. They also show HβBC profiles that are best fit with Gaussian functions, and some profile ap- pear composite. A few sources appear to deviate from the trend found in previous work that FWHM of HβBC and the Feii lines are very similar (Marziani et al., 2003a,c). HE 1249-0648 and HE 1258−0823 show FWHM(HβBC) ≫ FWHM(Feiiλ4570). While FWHM(Feiiλ4570) is sub- ject to large uncertainty, the difference is confirmed by careful reinspection of these spectra. This condition is seen in only 2/215 sources in the Marziani et al. (2003a) sample. Both (IRAS 07598+6508 and Mkn 235) are BAL QSOs which are also FIR ultra-luminous (Sulentic et al., 2006). In addition to a Civλ1549 BAL with high ter- minal velocity, these objects have a strong, blueshifted Civλ1549 emission line component. Blueshifted Balmer emission is probably associated with the high-ionization gas emitting Civλ1549. The absence of any detectable [Oiii]λλ4959,5007 emission, along with a possible Hei fea- ture at λ5016 A support the possibility that HE 1249-0648 and HE 1258−0823 could be BAL QSOs similar to IRAS Sulentic et al.: Hβ in Intermediate z Quasars: MBH and L/LEdd 5 07598+6508 and Mkn 235. Further support comes from an inspection of the HE survey spectra of both objects which indeed show blueshifted broad absorption in the Mgiiλ2800 line. reported in Table 2. All sources with z >∼ 0.8 are from the VLT sample. We do not show MBH values derived from MB as in Marziani et al. (2003b) for clarity, since they basically confirm the same trends obtained from the specific fluxes. 5. Black Hole Mass and Eddington Ratio 5.1. Basic Equations One can write the velocity dispersion v in Eq. 1 as v ≈ √3/2 FWHM(HβBC) in the case of randomly oriented ve- locities projected along the line-of-sight. The expression for black hole mass is then: MBH ≈ 3 4 rBLRFWHM(HβBC)2 G (3) In the absence of reverberation data one must rely on the correlation between reverberation radius and source luminosity rBLR ∝ (L5100)α, where L5100 is the specific luminosity at λ ≈ 5100 A (Kaspi et al., 2000, 2005). This formula relates BLR distance from the central continuum source and specific luminosity (ergs s−1A−1) near 5100 A. Following Kaspi et al. (2005) one can write: rBLR ≈ 0.697 · 1017 ·" λLλ(cid:0)5100A(cid:1) 1044 erg s−1 #0.67 cm (4) We can make two estimates of MBH using these rela- tionships assuming two different values for λLλ. The first method derives MB from tabulated values of V or B ap- parent magnitudes: λLλ(cid:0)5100A(cid:1) ≈ 3.14 · 1035−0.4(MB) ergs s−1 A second possibility involves estimating the specific luminosity near 5100 A directly from flux measures using the adopted continuum fits for our spectra. More explicitly λLλ is computed as follows: (5) λLλ(cid:0)5100A(cid:1) = 4πd2 Pλfλ(5100A) ergs s−1, (6) where dP is the redshift derived distance and fλ is the specific flux in the rest frame at 5100 A (after correction for Galactic extinction AB at the observed wavelength). The second choice has the advantage that the continuum fλ and FWHM(HβBC) are measured from the same spec- trum. Expected light losses are estimated to be ≈ 35 % of the quasar flux with average Paranal seeing (0."6). We apply this average correction to the observed flux before computing MBH. Substituting the expression for rBLR in the mass for- mula one obtains the following relation: MBH ≈ 5.48·106" λLλ(cid:0)5100A(cid:1) 1044 erg s−1 #0.67 1000km s−1 (cid:19)2 (cid:18) FWHM(HβBC) (7) where MBH is in solar mass units (M⊙). Fig. 3 shows the distribution of MBH as a function of z. MBH estimates are based on the FWHM(HβBC) measures 5.2. Improved MBH Estimators? As briefly summarized in §1, there is some evidence sug- gesting that the virial assumption is reasonable for LILs in a significant fraction of quasars (Population A; about 50-60% of low z quasars). Our VLT spectra confirm that any dependence of FWHM(HβBC) on source luminosity is weak (Sulentic et al., 2004). Table 3 suggests that low redshift trends for asymmetries and line shifts are pre- served in the intermediate z sample. It is therefore not certain that FWHM(HβBC) is a valid estimator of the virial velocity for all sources even after proper HβNC, [Oiii]λλ4959,5007, Feii subtraction. A more physical approach to MBH estimation uses the FWHM of the variable part of the HβBC profile (Peterson et al., 2004). Fig. 4 plots FWHMrms values from Peterson et al. (2004) versus our FWHM(HβBC) measures (Marziani et al., 2003a) for all sources in com- mon. Sources with FWHM <∼ 4000 km s−1 show a with FWHM>∼ 4000 km s−1. We recall correlation while the situation is less clear for sources this value indicates the nominal population A-B bound- ary (Sulentic et al., 2000b) that emerged in our E1 studies. A least-square best fit analysis yields a "cor- rected" FWHM(HβBC) estimate: FWHMcorr(HβBC)≈ −710(±800) + 1.13(±0.28) FWHM(HβBC) for FWHM <∼ 4000 km s−1 and FWHMcorr(HβBC)≈ 650(±1000) + 0.79(±0.14) FWHM(HβBC) for FWHM >∼ 4000 km s−1. that The slope depends somewhat on the fitting method (a ro- bust fit yields 0.94 and 0.72) but is always less steep for FWHM >∼ 4000 km s−1. An immediate implication is that optically thick BLR gas responding to continuum changes shows a velocity dispersion correlated with -- but slightly lower than that of the integrated profile at all FWHM. A break in the linear fit at 4000 km s−1 is consis- tent with several previous findings: (a) mean and pos- sibly systematic HβBC profile differences between Pop. A and B sources (Sulentic et al., 2002); (b) the lack of strong profile changes in Pop. A sources (Giveon et al., 1999); (c) profile asymmetries frequently observed in Pop. B sources. They might represent a distinct redshifted emis- sion component (Sulentic et al., 2002)which may arise in less optically thick gas than the rest of the HβBC pro- file (Sulentic et al., 2000c). If the redshifted component is emitted in an innermost VBLR, the FWHM of the whole HβBC profile is obviously increased over the value due to the line component that is actually responding, and that is most likely located farther away from the central con- tinuum source. Even if the virial assumption holds for the VBLR (but the frequent asymmetries warn us that this might not be the case), the use of the FWHM from the 6 Sulentic et al.: Hβ in Intermediate z Quasars: MBH and L/LEdd whole HβBC profile and of rBLR from the reverberating part, yields an MBH overestimate. This interpretation of the correlations in Fig. 4 is advanced with caution because of the small sample size and especially because of the poor statistics for sources with FWHM >∼ 4000 km s−1. It is also possible to produce a meaningful fit with a second- order polynomial. We apply a tentative correction to the FWHM measures FWHMcorr(HβBC) and therefore to re- sultant MBH estimates using the second-order fit shown in Fig. 4 which approximates very well the linear trends. If an optically thin/nonvirialized component is present in the HβBC profile of many sources then we might use as a virial estimator the FWHM of a line, or lines, arising in BLR gas but not likely to be present in the VBLR region. FWHM(Feiiλ4570) is an obvious alter- native because there is no evidence for a VBLR emis- sion component in the broad Feii blends. It is not strictly correct to use it because the rBLR- Lλ rela- tion was deduced for Hβ. The most serious difficulty lies in obtaining a reliable FWHM(Feiiλ4570) estimate from the heavily blended Feii emission. Considering the EW and FWHM limits for detection of Fe iiopt emis- sion (Marziani et al., 2003a) we conclude that a reason- able FWHM measurement is possible for ≈120 sources in our low z spectral atlas. Measurement uncertainties for FWHM(Feii) will be larger than for FWHM(HβBC) and are estimated to lie between 20 -- 50%. The best fit of FWHM(Feiiλ4570) vs. FWHM(HβBC) (km s−1) is consistent with FWHM(Feiiλ4570) ≈ FWHM(HβBC) for FWHM(HβBC) < is FWHM(Feiiλ4570) ≈ 0.67 · FWHM(HβBC) + 820 km s−1 if FWHM(HβBC) > The relationship between FWHM(HβBC) and FWHM (Feii) confirms that individual Feii lines show approxi- mately the same width as HβBC and as the rms HβBC component implying a common kinematic environment if ∼ 4000 km s−1, while it ∼ 4000 km s−1. FWHM(HβBC) <∼ 4000 km s−1. If FWHM(HβBC) >∼ 4000 km s−1, FWHM(Feii) follows a trend closer to that of the rms HβBC component. Therefore if the rms HβBC compo- nent arises from gas in virialized motion, the same can be reasonably assumed for the whole Feiiλ4570 emission. 5.3. MBH and L/LEdd Dependence on Redshift MBH,9⊙ = 3.85 · 106h−2(cid:18) L ·(cid:2)1.5(cid:0)1 − e− z LEdd(cid:19)−1 6.107(cid:1) +(cid:0)1 − e− z · 10−0.4mB · 1.266(cid:1)(cid:3)2 · (1 + z)(1−a). Appendix B details how Eqn. 8 was derived. If we adopt a limiting magnitude mB ≈ 17.5 (appropriate for the Hamburg-ESO survey) and consider a maximum Eddington ratio L/LEdd≈ 1 we obtain the minimum MBH detectable as a function of z. Limiting MBH curves are shown in Fig. 3, Fig. 5, and Fig. 6. All sources show MBH >∼ 106 M⊙. In the range 1 <∼ z <∼ 2.5 we begin to find a significant black hole popu- lation in the range ∼ 109.5 M⊙ <∼ MBH <∼ 1010 M⊙. If we consider MBH values derived from Feiiλ4570 as well as from FWHMcorr(HβBC) (Fig. 5, open and filled symbols respectively) we get a somewhat different picture. Almost all sources lie below ≈ 5· 109 M⊙ (all sources with z <∼ 1); only three sources whose MBH has been computed with from FWHMcorr(HβBC) lie significantly above this limit. Our sample of intermediate z quasars with compatible quality data is still small so caution is needed. It should corr is self-consistent, be pointed out that the use of HβBC since no Hβ surrogate line was used. The upper panel of Fig. 6 identifies sources on the ba- sis of radio-loudness where open and solid symbols denote radio-loud (RL; see definition in Sulentic et al., 2003) and radio-quiet (RQ) sources respectively. At low z RQ sources are distributed across the full range of MBH up to MBH>∼ 109 M⊙, although RL sources tend to have larger MBH (Metcalf & Magliocchetti, 2006; Marziani et al., 2003b, and references therein). The same consideration apply to Pop. A and Pop. B sources (lower panel of Fig. 6), and this is not surprising since most Pop. A sources are RQ. At intermediate z, selection effects limit the detectable black holes to MBH>∼ 109 M⊙, so that only the high end of the quasar MBH distribution can be traced. It is too early to decide whether systematic differences in the MBH distribution of RQ and RL sources may still exist at in- termediate z, at least from the present data. 5.3.1. MBH 5.3.2. L/LEdd Fig. 3 shows the distribution with z of MBH esti- mates derived using uncorrected FWHM(HβBC) mea- sures. This can be compared with Fig. 5 where we show corresponding distributions of MBH derived us- ing corrected FWHMcorr(HβBC) measures and using FWHM(Feiiλ4570). FWHMcorr(HβBC) values were de- rived from the second-order relation in Fig. 4. The distribution of data points in the log MBH vs. z plane reflects selection effects intrinsic to any flux limited sample. If we consider AGNs that are radiating at a given L/LEdd we easily obtain MBH for a given apparent mag- nitude: Fig. 7 shows the distribution of L/LEdd estimates as a function of z. A preliminary analysis in terms of L/LEdd, and including discussion of bolometric luminosity estima- tion, was presented in Marziani et al. (2003b). The existence of apparently super-Eddington radia- tors in the low-z part of Fig. 7 involves the most extreme NLSy1 sources. One caveat about interpreting any of them as super-Eddington involves the significant uncertainties associated with these estimates. If we assume a typical un- certainty of ±50% for the bolometric correction (neglect- ing beaming or lensing) and ±10% for the virial estimator (basically the uncertainty in the FWHM measurement) we Sulentic et al.: Hβ in Intermediate z Quasars: MBH and L/LEdd 7 find a ∆ logL/LEdd≈ 0.13 (neglecting the scatter in Eqn. 2). A serious source of uncertainty ignored until now in- volves source orientation with respect to the line-of-sight. The virial velocity dispersion that HβBC is assumed to measure is now widely assumed to involve Keplerian ro- tation in an accretion disk or in a flattened distribution near the disk. If our candidate super-Eddington sources are mostly viewed with (face-on disk and/or pole-on jet) then MBH can be significantly underestimated. An orien- tation correction applied to extreme NLSy1 sources (we call them Pop. A "blue outliers") interpreted as face-on, will move them below L/LEdd ≈ 1 (Marziani et al., 2003b, see arrows in their Figure 12). At high z we so far find no strong evidence for any super-Eddington sources. The thick solid line in Fig. 7 shows the expected L/LEdd de- tection limit deduced from Eqn. 8 for a flux-limited survey (mB ≈ 17.5) and a quasar with MBH ≈ 4 · 109 M⊙. All quasars in our sample fall above the minimum detectable L/LEdd, suggesting that masses larger than ≈ 4 · 109M⊙ are not strictly necessary from our data (see §6.1 for fur- ther discussion). A physical basis for the dichotomy between Pop. A and Pop. B (Sulentic et al., 2000b; Marziani et al., 2001, 2003b) is supported by this analysis in the sense that Fig. 4 shows evidence for a change at about FWHM(HβBC) ≈ 4000 km s−1. Fig. 7 identifies Pop. A and B sources as filled and open symbols respectively. Here we apply the luminosity-dependent definition of the Pop. A-B bound- ary as defined in Fig. 6 of Sulentic et al. (2004). It shows that the Eddington ratio of Pop. A sources is systemat- ically larger than that of Pop. B, and that the apparent boundary between the two populations may increase with redshift. If we focus on the intermediate z quasars then Pop. A sources show an average L/LEdd ≈ 0.78 compared to 0.27 for 10 Pop. B sources. Even these small samples of Pop. A and B sources show significantly different L/LEdd distributions according to a Kolmogorov-Smirnov (K-S) test. RL AGNs are systematically low L/LEdd radiators since they are almost entirely Pop. B sources. It is interesting to note that with: (1) a limiting magni- < tude mB ≈ 16.5, (2) L/LEdd<∼ 1 and (3) MBH ∼ 4·109 M⊙ we should detect fewer sources beyond z ≈ 2 (no source below the dot-dashed curve of Fig. 7). If our assumptions about the quasar bolometric correction are valid up to that redshift, selection effects may influence the relative frequency of Pop. A and B sources (the two Populations have different L/LEdd distributions) rather than the in- trinsic properties of LILs. In this case selection effects on L/LEdd should strongly influence the so-called "Baldwin effect" involving Civλ1549 and other HILs (Bachev et al., 2004, and references therein) because Civλ1549 is more prominent in Pop. B, and Pop. B sources are more easily lost at high z. 6. Discussion The present paper provides MBH estimates that have three advantages: (1) a consistent data analysis procedure is em- ployed over the entire redshift range 0.0 <∼ z <∼ 2.5 by using the same MBH tracer, HβBC; (2) S/N and resolu- tion of the spectroscopic data sample are high enough to permit a careful study of the HβBC profile, and (3) the data quality allows reasonable estimates of FWHM Feii in many of the sources. 6.1. What are the largest black hole masses? Netzer (2003) discussed several problems with MBH >∼ 1010 M⊙. If the black hole mass vs. bulge mass (Mbulge) relation (Ferrarese & Merritt, 2000) is valid at high z then MBH ∼ 1010 M⊙ would imply stellar velocity dispersion σ⋆ ≈ 700 km s−1 (following Gebhardt et al., 2000) and resultant bulge masses Mbulge >∼ 1013M⊙ which are not observed at low-z (McLure & Jarvis, 2004; Netzer, 2003; Wang, 2003). Recent results for the fundamental plane of elliptical galaxies, and the most massive spheroids at z < ∼ 0.3 from Sloan Digital Sky Survey measures, confirm ∼ 500 km s−1 (SDSS, Bernardi et al., 2003, 2005, that σ⋆ ∼ 500 km s−1 are likely due to chance all galaxies with σ > superposition). < There are several proposed interpretations of this prob- lem: (1) the MBH -- Mbulge relationship may not strictly hold for all hosts, (2) the virial assumption is not appli- cable, (3) results are plagued by such large uncertainties -- including the one of the luminosity index α -- that very large mass estimates are not real (Vestergaard, 2002), and (4) some systematic effects may not have been considered. 6.1.1. The MBH -- Mbulge Relationship Nuclei with MBH >∼ 5 · 109 M⊙ are not observed in galax- ies of the local Universe if a direct black hole mass de- termination is possible from circum-nuclear kinematics (Marconi & Hunt, 2003). They are expected to be rare and difficult to find, considering also that they should be in a dormant or nearly dormant stage at the present epoch. Integrating the quasar luminosity function at z ≈ 1.5 (Boyle et al., 2000), we find a comoving density of quasars above the HE limiting magnitude (which corresponds to MB≈ −27.1 for a k−correction a = 0.6) ∼ 3 · 10−8 Mpc−3. This indicates that the present-day density of the most massive black holes that were once luminous quasars should be very low, ∼ 2 · 10−9 Mpc−3. Even so, they should be much more frequent than the very mas- sive spheroids that would host them if the MBH -- Mbulge relationship is valid. To estimate the density of spheroids with σ⋆ bution function provided by Sheth et al. (2003) to unob- served domains in σ. Integrating the Sheth et al. (2003) ∼ 500 km s−1, one obtains that the co- function for σ⋆ moving density of all local spheroids is three orders of magnitude lower (∼ 6 · 10−13 Mpc−3) than that of the most massive black holes. A possible implication is that the MBH -- Mbulge relation is not linear or of universal validity, i.e. some galaxies host larger black hole masses ∼ 500 km s−1, one must extrapolate the σ distri- > > 8 Sulentic et al.: Hβ in Intermediate z Quasars: MBH and L/LEdd than expected. This conclusion should remain valid also at intermediate redshifts. A constant MBH/Mbulge ratio cannot hold forever if bulge mass grows by secular evolu- tionary processes. Evolution goes in the sense that the spheroids may increase their masses at a later cosmic age (z <∼ 2; Treu et al., 2004; Peng et al., 2005) signifi- cantly more than their central black hole: the most mas- sive black hole should be already "brightly shining" at z ≈ 2 (McLure et al., 2005, and references therein). A nonlinear MBH -- Mbulge relation has been proposed (Laor, 2001) for local hosts, with MBH accounting for only 0.05 % of the bulge mass in low-luminosity galaxies and 0.5 % in giant ellipticals. In this case, MBH∼ 1010 M⊙ would imply Mbulge∼ 1012 M⊙, which is at the upper end of the spheroid masses measured locally (Marconi & Hunt, 2003). The problem can be made to disappear if MBH<∼ 3 · 109 M⊙. Bernardi et al. (2005) find a maximum σ⋆ ≈ 500 km s−1. These local galaxies (with expected MBH ≈ 3·109 M⊙) may have been the hosts of the luminous quasars ∼ z < at 1 < ∼ 2 since the number density of spheroids with ∼ 350 km s−1 is ∼ 4·10−7 Mpc−3, still somewhat larger σ⋆ than the present-day density of supermassive black holes > that were radiating at MB <∼ −25 (∼ 10−7 Mpc−3). The curves in Fig. 7, where MBH∼ 4 · 109 M⊙ has been assumed, suggest that there is no need for MBH >∼ 1010 M⊙ at z < ∼ 2.5. Similarly the brightest HE sources in the redshift range 1 < ∼ 2 (mB ≈ 15.0 at z ≈ 1.7) are consistent with L/LEdd≈ 1 if MBH∼ 4 · 109 M⊙. In other words, if MBH∼ 1010 M⊙, then a source at L/LEdd≈ 1 would be brighter than the brightest quasars observed in the range 1 < ∼ 2 (if H0 is close to the value assumed in this paper). ∼ z < ∼ z < 6.1.3. Statistical Errors It is interesting to consider in more detail the sources that show MBH ∼ 1010 M⊙ in our sample. HE 0248 -- 3628 with the largest estimated MBH is also one of the highest L/LEdd radiators with L/LEdd∼ 1. It shows an anoma- lously large (by a factor ∼10) flux with respect to our other VLT sources which makes it an extremely luminous quasar (MB<∼ −30; see Sulentic et al., 2004). It falls near the Pop. A -- B boundary in Eigenvector 1 space and is a borderline radio-loud source by our definition (Sulentic et al., 2003). It disappears as an outlier if the MBH estimated from the tabulated MB is adopted. HE 2355 -- 4621 is a second source with MBH∼ 1010 M⊙. It behaves like a normal radio-quiet Pop. B source. The HβBC profile shows a prominent red- ward asymmetry, which is strongly affecting the width at half maximum. It is also interesting to consider HE 1104 -- 1805, which is the most luminous source in our new sam- ple with (MB≈ -29.5). It behaves like an ordinary Pop. A source and shows MBH≈ 5 · 109 M⊙ which may be over- estimated if the continuum is lens brightened (Appendix A). ∼ z < The high-mass wing of the MBH distribution in the redshift range 1 < ∼ 2 is consistent with the wing of a Gaussian peaked at ≈ 3 ·109 M⊙ and dispersion ∆ log MBH ≈ 0.3 if MBH is computed from FWHMcorr(HβBC) (4 ·109 M⊙ if no correction to HβBC is applied). The Gaussian dispersion is consistent with the estimated er- rors of individual measurements. K − S tests do not favor significantly different peak masses or a much different dis- persion. This suggests that sources with MBH >∼ 3 ·109 M⊙ of our data might be mostly due to random errors associated to the uncertainty in individual MBH measure- ments. 6.1.2. Could A Non-virial Component Yield a Huge 6.1.4. Systematic Effects MBH? One might envision a non-virial component that increases with source luminosity and systematically broadens the HβBC profile. In our so-called Pop. A sources, the high ionization wind that can dominate Civλ1549 emission might reasonably be expected to produce Balmer line emission as well. Such an additional (blueshifted) com- ponent on the HβBC profile would increase FWHM mea- sures (Sulentic et al., 2006, HE 1249 -- 0648 and HE 1258 -- 0823 may show this effect). One intriguing possibility is that high-luminosity sources are extreme Pop. A sources which is supported by models invoking a radiation pres- sure driven wind. Such an effect would make the correc- tion (∼ 500 km s−1) deduced for low-z Pop. A sources inadequate. At the other extreme, strong redward asym- metries observed in Pop. B sources indicate that the in- tegrated profile may be affected by gravitational redshift and non-virial motions (Marziani et al., 2003b). A quanti- tative analysis of these suggestions needs careful analysis beyond the scope of this paper, but in both cases any cor- rection would lower MBH. A first systematic effect considers the uncertainty in the index α. Our MBH estimates show the onset of MBH>∼ 3·109 M⊙ at z ∼0.8; at z >∼0.8 we observe some of the most luminous HE quasars. The high luminosity range of the Kaspi et al. (2005) relationship remains poorly sampled: there are just 2-3 sources in the range 45 <∼ log λLλ <∼ 46 (to which most of our intermediate z sources belong), creating a sample bias and making the correlation analysis intrinsically unstable. In addition, a recent reanalysis of the rBLR -- λLλ correlation suggests a value of α as low as 0.5 (Vestergaard & Peterson, 2006). If α is overestimated by ∆α ≈ 0.17 over a luminosity range of ∼ 10, MBH may be overestimated by ∆ log MBH≈ 0.2 in our intermediate z sources. Pop. A sources show good evidence that LILs are emit- ted in a strongly flattened system, probably an accretion disk or gas co-planar with the disk. Civλ1549 in these sources seems to be dominated by a wind component (Bachev et al., 2004). If these considerations apply also to Pop. B sources (but it is by no means clear, given the large FWHM(HβBC) of sources believed to be observed Sulentic et al.: Hβ in Intermediate z Quasars: MBH and L/LEdd 9 7. Conclusions Nine intermediate z VLT/ISAAC spectra with high res- olution and S/N supplement an earlier sample of 17 sources. Emission line measurements on HβBC and Feii presented in this paper strengthen the conclusion of Sulentic et al. (2004) that luminosity effects are weak or absent in the low-ionization lines of AGNs. Results on the HβBC profile are consistent with the population A-B hy- pothesis and Eigenvector 1 parameter space concept de- veloped for low z AGNs. We computed virial masses and Eddington ratios for the 25 intermediate-z objects plus about 280 lower z sources, using the same emission line over the entire redshift range for the first time. We also have how the distributions of MBH and L/LEdd vs. z are shaped by selection effects intrinsic to any flux-limited survey at the low MBH end. At the high MBH end, masses exceeding a few 109 M⊙ may be rare if corrections for non-virial broadening and statistical er- rors are taken into account. This suggestion is based on just 25 objects distributed over the entire redshift range 0.9 <∼ z <∼ 2.5. Confirmation from a larger sample of intermediate-z observations is needed. < pole-on) and if the maximum angle between disk axis and line-of-sight is ≈ 45◦, a correction could imply a factor ∼ 2 systematic increase in MBH. However, these consider- ation of orientation effects may not even reopen the prob- lem of very large MBH values: if our suggestion of a maxi- mum MBH≈ 3·109 M⊙ is appropriate, taking into account the systematic orientation effects would yield a maximum MBH≈ 6 · 109 M⊙ which is still plausible. The result that L/LEdd<∼ 1 at z >∼ 0.8 would be reinforced by systematic orientation effects. Summing up, our data suggest that very large masses ∼ 1010 M⊙ may be not be real, and may be predominantly due to statistical errors and emission line profile broaden- ing that is in part non-virial. The data presented in this paper are consistent with MBH not exceeding 3·109 M⊙ for our sources if the correction to FWHM(HβBC) described earlier is applied. 6.2. The Best MBH Estimators LILs like Mgiiλ2800 and Feii may yield more reliable re- sults than HβBC. Hi Balmer line emission can be substan- tial from gas in a variety of physical conditions. In Pop. B sources, a very broad component may increase the FWHM of the integrated HβBC profile (also mimicked by low S/N data; McIntosh et al., 1999; Shemmer et al., 2004). This very broad component may be optically thin to the ion- izing continuum, and therefore non-responsive to contin- uum changes. The HβBC profile of Pop. A sources may be affected by a high-ionization component mentioned earlier. Feii is thought to be emitted in a region very optically thick to Lyman continuum which is probably photoionized (Vestergaard & Peterson, 2005; Wang et al., 2005), as the part of HβBC responding to continuum changes should be. It is not surprising that the rever- berating part of HβBC and Feii provides width estimates which are consistent, since they are expected to mea- sure the width of a similar sub-region within the BLR. Similar considerations apply to Mgiiλ2800 (Wills et al., 1985) since Mgiiλ2800 should be mainly emitted in the same zone as Feii. McLure & Dunlop (2004) present virial MBH estimates for ≈13000 quasars in the redshift inter- val 0.1 <∼ z <∼ 2.1 based on spectra from the SDSS first data release. The mean MBH increases with increasing red- shift basically as shown in Fig. 3. The mass values found by them are also consistent with a limiting MBH around 3 · 109 M⊙, with large scatter. They use FWHM(HβBC) or FWHM(Mgiiλ2800) and find a consistency between the most massive at z ≈ 2 and those at z ≈ 0. Given measure- ment difficulties, and doubts about the virial assumption for most other lines, corrected measures for HβBC, Feii, and Mgiiλ2800 may offer the best hope for reliable MBH and L/LEdd estimates out to z ≈ 2.5. 10 Sulentic et al.: Hβ in Intermediate z Quasars: MBH and L/LEdd Acknowledgements. We thank a referee for many useful sug- gestions and the tenacity to ensure that they were taken into account. We also thank Lutz Wisotzki for providing us with the HE optical spectra. D. D-H, and JS acknowledge financial support from grant IN100703 from PAPIIT, DGAPA, UNAM. Appendix A: Notes on Individual Sources A.1. HE0512-3329 HE 0512 -- 3329 was discovered as a probable gravitation- ally lensed quasar with the Space Telescope Imaging Spectrograph (STIS). It is a doubly imaged QSO with a source redshift of z = 1.58 and an image separation of 0".644. The flux ratio A B of the lensed images shows a strong dependence on wavelength. In the R and I bands, A is brighter than B by about 0.45 mag while the two are almost equal in the band B. For smaller wavelengths, especially close to the limit near 2000 A, B becomes much brighter than A by about 1.3 mag. A natural ex- planation for this effect is differential reddening caused by different extinction effects in the two lines of sight (Gregg et al., 2000). Microlensing by stars and other com- pact object in the lensing galaxy also plays a role in this source (Wucknitz et al., 2003). Unfortunately, the small separation does not allow to us to distinguish component A and B on the ISAAC spectrum. The HE 0512 -- 3329 ac- quisition image is compatible with an unresolved source. Considering the flux ratios in the R and I band, it is rea- sonable to assume that our spectrum is dominated by the A component. A.2. HE1104-1805 HE 1104-1805 is a double-image lensed quasar discov- ered by Wisotzki et al. (1993). The image separation is ∆θ = 3′′.19, the source redshift is zs = 2.319, and the lens redshift is zl = 0.729. Wisotzki et al. (1995 ) reported that the continuum flux in both images is highly variable but that the line fluxes do not change, as expected if mi- crolensing is operating. On the acquisition image of HE 1104 there is a second source at 3.5" but it is completely off-slit. Sulentic et al.: Hβ in Intermediate z Quasars: MBH and L/LEdd 11 Table 1. Basic Properties of Sources and Log of Observations Object name (1) HE0507−3236 HE0512−3329 HE0926−0201 HE0946−0500 HE1003+0149 HE1017−0009 HE1104−1805 HE1249−0648 HE1258−0823 ma B (2) 17.36 17.03 16.23 16.24 16.45 16.69 16.45 16.72 16.26 zb (3) Linec M d B (5) (4) log Re K (6) Datef (7) Bandg DITh Ni exp Airmassj (8) (9) (10) (11) 1.5770 (7) 1.5873 (7) 1.6824 (7) 1.1013 (7) 1.0809 (7) 1.1295 (7) 2.3192 (7) 1.1940 (7) 1.1632 (7) 1 1 1 1 1,2 1 1,2 1 1 −27.6 <∼ 0.51 −28.0 <∼ 0.38 −29.0 <∼ -0.33 −28.0 <∼ 0.03 −27.7 −27.6 −29.6 <∼ 0.19 −27.7 <∼ 0.23 −28.1 <∼ 0.04 0.20 0.87 2003-10-22 2003-10-22 2004-02-11 2004-02-11 2004-02-13 2004-02-28 2004-02-11 2004-02-05 2004-03-29 J J J sZ sZ sZ sH sZ sZ 180 180 180 180 180 180 180 180 180 18 16 12 12 12 12 20 16 17 1.23-1.09 1.07-1.02 1.11-1.17 1.17-1.27 1.19-1.14 1.97-1.62 1.14-1.35 1.04-1.06 1.06-1.16 S/Nk (12) 25 -- 20 20 -- 15 30 -- 10 30 20 20 -- 10 20 30 -- 15 30 a Apparent B magnitude corrected because of Galactic absorption. b Redshift, with uncertainty in parenthesis. c Lines used for redshift calculations: 1: Hβ, 2: [Oiii]λ5007. d Absolute B magnitude, computed for H0=70 km s−1Mpc−1, ΩM = 0.3, ΩΛ = 0.7, and k-correction spectral index a=0.6. e Decimal logarithm of the specific flux ratio at 6cm and 4400 A (effective wavelength of the B band). Upper limits are from the NVSS (≈ 2.5 mJy), and would place all undetected sources in the RQ domain. f Date refers to time at start of exposure. g Photometric band. h Detector Integration Time (DIT) of ISAAC, in seconds. i Number of exposures with single exposure time equal to DIT. j Airmass at start and end of exposure. k S/N at continuum level in the proximity of Hβ. Two values are reported in case of different S/N on the blue and red side of Hβ (blue side first). The S/N value is with N estimated at a 2σ confidence level i.e., 2 times the rms. Table 2. Measurements of Fluxes, Equivalent Widths and FWHM of Strongest Lines Object name F(HβBC)a W(HβBC)b W(Feiiλ4570)c FWHM(Feiiλ4570)d (1) (2) HE0507−3236 HE0512−3329 HE0926−0201 HE0946−0500 HE1003+0149 HE1017−0009 HE1104−1805 HE1249−0648 HE1258−0823 12.5 ± 3.0 20.0 +5.0 −3.0 26.0 ± 4.0 17.0 +3.0 −2.0 7.0 ± 1.0 10.5 ± 1.0 20.0 ± 4.0 15.5 ± 2.0 24.0 ± 3.0 (3) 62 ±20 75 +30 −10 72 ± 15 59 +20 −10 48 ± 10 60 +15 −10 72 ± 15 63 ± 15 44 ± 10 (4) 17 ± 3 47 ± 10 20 ± 2 19 +5 −3 67 ± 12 11 +5 −3 39 +7 −5 31 +5 −2 20 ± 4 (5) 4100 +1600 −900 2400 +1300 −500 3400 ± 1200 2900 ± 1000 2100 ± 1400 3600 +2100 −1300 4100 ± 1700 1400 +1000 1400 +1000 −400 −0 Appendix B: Minimum MBH and L/LEdd as a Function of z The absolute B magnitude MB can be related to the spe- cific luminosity at 5100 A assuming an average spectral shape between 4400 (effective wavelength of B band) and 5100 A. If the spectral shape is described by a power-law (fν ∝ ν−b) with b = 0.3 (Marziani et al., 2003b): log[λLλ(λ = 5100A)] = −0.4MB + 35.497. (B.1) If we further assume that L = 10λ · Lλ, with λ=5100 A i.e., a bolometric correction factor 10, we can write: log[λLλ(λ = 5100A)] = L 10 · LEdd · 1.3 · 1047MBH,9⊙ (B.2) where the black hole mass MBH has been written in units of 109 M⊙. The absolute magnitude MB is: MB = mB + 5 − 5 log dL − 2.5(a − 1) log(1 + z) (B.3) where the luminosity distance dL is in parsec, and the last term is the K(z) correction. the co-moving distance: Assuming ΩM 6= 0, ΩΛ 6= 0, and Ωk = 0, we have for H0 Z z (B.4) dC = dz ′ c 0 pΩM(1 + z)3 + ΩΛ This integral can be approximated with residuals less c than 3% at all z (maximum error for z → 0). 6.107(cid:1) + 0.996(cid:0)1 − e− dC ≈ H0 (cid:2)(1.500)(cid:0)1 − e− z z 1.266(cid:1)(cid:3) (B.5) 12 Sulentic et al.: Hβ in Intermediate z Quasars: MBH and L/LEdd Object name F(HβNC)e W(HβNC)f W([Oiii]λ4959)f W([Oiii]λ5007)f (6) HE0507−3236 HE0512−3329 HE0926−0201 HE0946−0500 HE1003+0149 HE1017−0009 HE1104−1805 HE1249−0648 HE1258−0823 (7) 7: 3: 28: 14 +9 −5 15 +18 −5 15 +10 −5 60 ± 18 18 ± 5 9: (8) 0.3: 0.1: 0.8: 0.9 1.0 0.8 2.0 ± 0.5 0.7 ± 0.3 0.8 ± 0.3 (9) 2.7 ± 0.5 2.5 ± 0.5 5.6 ± 0.5 1.1 ± 0.5 0.6 ± 0.5 1.4 ± 0.5 3.9 ± 0.5 ... ... (10) 8.4 ± 0.6 6.5 ± 1.0 15.2 ± 1.0 3.4 ± 0.6 1.7 ± 0.5 3.8 ± 0.6 11.4 ± 0.5 ... ... a Rest frame flux of HβBC in units of 10−13 ergs s−1 cm−2 and ±2σ confidence level uncertainty. b Rest frame equivalent width of HβBC in A ±2σ confidence level uncertainty. c Rest frame equivalent width of the Feiiλ4570 blend in A ±2σ confidence level uncertainty. d FWHM of lines in the Feiiλ4570 blend and uncertainty at 2σ, in km s−1. See text for details. e Rest frame flux of HβNC in units of 10−16 ergs s−1 cm−2 and ±2σ confidence level uncertainty. Colons indicate highly uncertain values. f Rest frame equivalent width of HβNC, [Oiii] l4959, and [Oiii]λ5007 in A, with uncertainty at 2σ. Colons indicate highly uncertain values. Table 3. HβBC Line Profile Measurements Source (1) FWZIa ∆a,b (3) (2) FWHMa ∆a,b (5) (4) A.I.c (6) HE0507−3236 HE0512−3329 HE0926−0201 HE0946−0500 HE1003+0149 HE1017−0009 HE1104−1805 HE1249−0648 HE1258−0823 16000 16000 17000 20000 9000 16000 15000 21000 17000 2100 3700 1400 2800 2100 3800 1500 1000 2600 3500 3100 5100 3600 2900 6200 4300 4900 4400 300 −0.15 300 −0.16 500 0.21 250 −0.06 280 −0.06 290 −0.04 260 0.11 500 0.13 500 −0.08 ∆b (7) 0.10 0.07 +0.07 −0.19 0.09 +0.07 −0.11 +0.08 −0.05 0.09 +0.11 −0.07 +0.20 −0.08 Kurt.d ∆b (9) (8) c(0/4) ∆a,b (11) (10) 0.31 0.31 0.27 0.37 0.33 0.41 0.38 0.31 0.23 0.06 0.04 0.06 0.05 0.05 0.04 0.05 0.05 0.06 −500 −100 1200 −600 −400 200 900 1400 −700 1100 1900 700 1400 1000 1900 800 500 1300 The luminosity distance is: References dL = dC(1 + z) (B.6) and we can use the new relationship to derive a working relationship for MBH: MBH,9⊙ = 3.85 · 106h−2(cid:18) L z LEdd(cid:19)−1 6.107(cid:1) +(cid:0)1 − e− · 10−0.4mB · 1.266(cid:1)(cid:3)2 z ·(cid:2)1.5(cid:0)1 − e− assume a = 0.6. Differences for 0.3 < where h = H0/75. Curves shown in Figs. 3, 5, 6, 7 ∼ 0.6 are minor. ∼ a < Aoki, K., Kawaguchi, T., & Ohta, K. 2005, ApJ, 618, 601 Bachev, R., Marziani, P., Sulentic J. W., Dultzin-Hacyan D., Calvani M., ApJ, in press Baldwin, J. A., et al. 1996, ApJ, 461, 664 Bernardi, M., et al. 2003, AJ, 125, 1866 Bernardi, M., et al. 2005, ArXiv Astrophysics e-prints, arXiv:astro-ph/0510696 (1 + z)(1−a), Boroson, T. A. 2003, ApJ, 585, 647 Boroson, T. 2005, ArXiv Astrophysics e-prints, arXiv:astro-ph/0505127 Boroson, T. A., & Green, R. F. 1992, ApJS, 80, 109 Boyle, B. J., Shanks, T., Croom, S. M., Smith, R. J., Miller, L., Loaring, N., & Heymans, C. 2000, MNRAS, 317, 1014 Brotherton, M. S., Wills, B. J., Steidel, C. C., & Sargent, W. L. W. 1994, ApJ, 423, 131 Brotherton, M. S. 1996, ApJS, 102, 1 Collin-Souffrin, S., Dyson, J. E., McDowell, J. C., & Perry, J. J. 1988, MNRAS, 232, 539 Sulentic et al.: Hβ in Intermediate z Quasars: MBH and L/LEdd 13 Source (12) c(1/4) ∆a,b (13) (14) c(1/2)a ∆a,b (16) (15) c(3/4)a ∆a,b (18) (17) c(0.9)a ∆a,b (20) (19) HE0507−3236 HE0512−3329 HE0926−0201 HE0946−0500 HE1003+0149 HE1017−0009 HE1104−1805 HE1249−0648 HE1258−0823 −300 −500 800 −300 −100 −900 200 400 −400 300 200 +400 −1100 200 200 300 300 400 +700 −300 50 −180 −150 −320 −50 −610 −180 −70 −330 150 150 250 120 130 130 130 240 220 100 −80 −160 −160 −20 −350 −190 −60 −30 90 80 130 100 70 140 120 130 110 90 −30 −110 −70 10 −360 −190 −110 −40 50 50 80 70 50 120 80 80 50 a In units of km s−1. b 2σ confidence level uncertainty. c Asymmetry index defined as in Marziani et al. (1996). d Kurtosis parameter as in Marziani et al. (1996). Corbin, M. R. 1991, ApJ, 375, 503 Corbin, M. R., & Francis, P. J. 1994, AJ, 108, 2016 de Robertis, M. 1985, ApJ, 289, 67 Ferrarese, L., & Merritt, D. 2000, ApJL, 539, L9 Gaskell, C. M. 1982, ApJ, 263, 79 Gebhardt, K., et al. 2000, ApJL, 539, L13 Giveon, U. et al. 1999, MNRAS, 306, 637 Gregg, M. D., Wisotzki, L., Becker, R. H., Maza, J., Schechter, P. L., White, R. L., Brotherton, M. S., & Winn, J. N. 2000, AJ, 119, 2535 Grupe, D., Beuermann, K., Mannheim, K., & Thomas, H.-C. 1999, A&A, 350, 805 Marziani P., Zamanov R., Sulentic J. W., Calvani M., 2003c, MNRAS, 345, 1133 S. 2000, MNRAS, 314, L17 Mathur, S., Kuraszkiewicz, J., & Czerny, B. 2001, New Astronomy, 6, 321 McIntosh, D. H., Rieke, M. J., Rix, H.-W., Foltz, C. B., & Weymann, R. J. 1999, ApJ, 514, 40 McLure, R. J., & Dunlop, J. S. 2001, MNRAS, 327, 199 McLure, R. J., & Dunlop, J. S. 2004, MNRAS, 352, 1390 McLure, R. J., & Jarvis, M. J. 2004, MNRAS, 353, L45 McLure, R. J., Jarvis, M. J., Targett, T. A., Dunlop, J. S., & Best, P. N. 2005, ArXiv Astrophysics e-prints, arXiv:astro-ph/0510121 Jackson, N., Perez, E.,& Penston, M. V. 1991, MNRAS, Metcalf, R. B., & Magliocchetti, M. 2006, MNRAS, 365, 249, 577 101 Joly, M. 1988, AAp, 192, 87 Kaspi S., Smith P.S., Netzer H., et al. 2000, ApJ, 533, 631 Kaspi, S., Maoz, D., Netzer, H., Peterson, B. M., Vestergaard, M., & Jannuzi, B. T. 2005, ArXiv Astrophysics e-prints, arXiv:astro-ph/0504484 Koratkar, A. P., & Gaskell, C. M. 1991, ApJL, 370, L61 Krolik, J. H. 2001, ApJ, 551, 72 Kuraszkiewicz, J. K., Green, P. J., Crenshaw, D. M., Dunn, J., Forster, K., Vestergaard, M., & Aldcroft, T. L. 2004, ApJS, 150, 165 Laor, A. 2001, ApJ, 553, 677 Malkan, M. A., & Sargent, W. L. W. 1982, ApJ, 254, 22 Marconi, A., & Hunt, L. K. 2003, ApJ, 589, L21 Marziani, P., & Sulentic, J. W. 1993, ApJ, 409, 612 Marziani P., Sulentic J.W., Dultzin-Hacyan D., Calvani M., Moles M., 1996, ApJS 104, 37 Netzer, H. 2003, ApJ, 583, L5 Onken, C. A., Ferrarese, L., Merritt, D., Peterson, B. M., Pogge, R. W., Vestergaard, M., & Wandel, A. 2004, ApJ, 615, 645 Osterbrock, D. E., & Shuder, J. M. 1982, ApJS, 49, 149 Peng, C. Y., Impey, C. D., Ho, L. C., Barton, E. J., & Rix, H.-W. 2005, ArXiv Astrophysics e-prints, arXiv:astro-ph/0509155 Peterson, B. M., & Wandel, A. 1999, ApJL, 521, L95 Peterson, B. M., & Wandel, A. 2000, ApJ, 540, L13 Peterson, B. M., et al. 2004, ApJ, 613, 682 Reimers, D., Koehler, T., & Wisotzki, L. 1996, AApS, 115, 235 Richards, G. T., Vanden Berk, D. E., Reichard, T. A., Hall, P. B., Schneider, D. P., SubbaRao, M., Thakar, A. R., & York, D. G. 2002, AJ, 124, 1 Marziani P., Sulentic J.W., Zwitter T., Dultzin-Hacyan Rokaki, E., Lawrence, A., Economou, F., & Mastichiadis, D., Calvani M., 2001, ApJ, 558, 553 (M01) A. 2003, MNRAS, 340, 1298 Marziani P., Sulentic J. W., Zamanov R., Calvani M., Dultzin-Hacyan D., Bachev R., Zwitter T., 2003a, ApJS, 145, 199 (M03) Marziani, P., Zamanov, R., Sulentic, J. W., Dultzin- Hacyan, D., Bongardo, C., & Calvani, M. 2003b, ASP Conf. Ser. 290: Active Galactic Nuclei: From Central Engine to Host Galaxy, 229 Shang, Z., et al. 2004, ArXiv Astrophysics e-prints, astro-ph/0409697 Sheth, R. K., et al. 2003, ApJ, 594, 225 Shemmer, O., Netzer, H., Maiolino, R., Oliva, E., Croom, S., Corbett, E., & di Fabrizio, L. 2004, ApJ, 614, 547 Shields, J. C., Ferland, G. J., & Peterson, B. M. 1995, ApJ, 441, 507 14 Sulentic et al.: Hβ in Intermediate z Quasars: MBH and L/LEdd Stirpe, G. M. 1990, A&AS, 85, 1049 Sulentic, J. W. 1989, ApJ, 343, 54 Sulentic, J. W., Zheng, W., Calvani, M., & Marziani, P. 1990, ApJ, 355, L15 Sulentic, J. W., Marziani, P., & Dultzin-Hacyan, D. 2000a, ARA&A, 38, 521 Sulentic, J. W., Marziani, P., Zwitter, T., Dultzin-Hacyan, D., & Calvani, M. 2000b, ApJ, 536, L5 Sulentic, J. W., Marziani, P., Zwitter, T., Dultzin-Hacyan, D., & Calvani, M. 2000c, ApJL, 545, L15 Sulentic, J. W., Marziani, P., Zamanov, R., Bachev, R., Calvani, M., & Dultzin-Hacyan, D. 2002, ApJ, 566, L71 Sulentic J. W., Zamfir S., Marziani P., Bachev R., Calvani M., Dultzin-Hacyan D., 2003, ApJ, 597, L17 Sulentic, J. W., Stirpe, G. M., Marziani, P., Zamanov, R., Calvani, M., & Braito, V. 2004, AAp, 423, 121 (Paper I) Sulentic, J. W., Dultzin-Hacyan, D.., Marziani, P., Bongardo, C., Braito V., Zamanov, R., Calvani, M. 2006, RevMexAAp, accepted Treu, T., Malkan, M. A., & Blandford, R. D. 2004, ApJ, 615, L97 V´eron-Cetty, M.-P., V´eron, P., & Gon¸calves, A. C. 2001, A&A, 372, 730 Vestergaard, M. 2002, ApJ, 571, 733 Vestergaard, M., & Peterson, B. M. 2005, ApJ, 625, 688 Vestergaard, M., & Peterson, 2006, B. M. astro-ph/0601303 Wandel, A., Peterson, B. M., & Malkan, M. A. 1999, ApJ, 526, 579 Wanders, I., et al. 1995, ApJL, 453, L87 Wang, J. 2003, AJ, 125, 2859 Wang, J., Wei, J. Y., & He, X. T. 2005, A&A, 436, 417 Wills, B. J., Netzer, H., & Wills, D. 1985, ApJ, 288, 94 Wisotzki, L., Koehler, T., Ikonomou, M., & Reimers, D. 1995, A&A, 297, L59 Wisotzki, L., Koehler, T., Kayser, R., & Reimers, D. 1993, A&A, 278, L15 Wisotzki, L., Christlieb, N., Bade, N., Beckmann, V., Kohler, T., Vanelle, C., & Reimers, D. 2000, A&A, 358, 77 Wu, X.-B., Wang, R., Kong, M. Z., Liu, F. K., & Han, J. L. 2004, AAp, 424, 793 Wucknitz, O., Wisotzki, L., Lopez, S., & Gregg, M. D. 2003, A&A, 405, 445 Zamanov, R., Marziani, P., Sulentic, J. W., Calvani, M., Dultzin-Hacyan, D., & Bachev, R. 2002, ApJ, 576, L9 Sulentic et al.: Hβ in Intermediate z Quasars: MBH and L/LEdd 15 Fig. 1. Calibrated VLT-ISAAC spectra for 9 new intermediate-redshift quasars. Abscissae are rest-frame wavelength in A, ordinates the rest-frame specific flux in units of 10−15 ergs s−1 cm−1 A−1. 16 Sulentic et al.: Hβ in Intermediate z Quasars: MBH and L/LEdd Fig. 2. Spectral Atlas for the new intermediate-redshift quasars. Left-hand panels show the continuum-subtracted Hβ spectral region. Axes are the same as Fig. 1. The best-fit Feii emission model is traced as a thin (green) line. Right-hand panels show the enlarged Hβ profiles after continuum and Feii subtraction. The (blue and red) thick line shows a spline fit of the short and long wavelength sides of the HβBC profile, respectively. Sulentic et al.: Hβ in Intermediate z Quasars: MBH and L/LEdd 17 Fig. 3. Distribution of MBH estimates as a function of z using uncorrected FWHM(HβBC) values. Intermediate z sources observed with ISAAC are at z >∼ 0.8. The solid curve estimates the minimum detectable MBH for our sample -- see text. 18 Sulentic et al.: Hβ in Intermediate z Quasars: MBH and L/LEdd Fig. 4. FWHM measured from the variable part of the HβBC profile for low-z sources with reverberation data (Peterson et al., 2004) versus FWHM(HβBC) (km s−1) measures from Marziani et al. (2003c). Solid straight lines show best fits for FWHM(HβBC)≤ 4000 km s−1(Pop. A) and for sources with broader lines(Pop. B). The solid curve shows a weighted least-square fit of all data with a second order polynomial. The diagonal dot-dashed line indicates the location of equal ordinate and abscissa values. Sulentic et al.: Hβ in Intermediate z Quasars: MBH and L/LEdd 19 Fig. 5. MBH computed from corrected FWHMcorr(HβBC) as a function of z (filled circles). MBH estimates computed from FWHM(Feiiλ4570) measures are shown as open symbols. Symbol types and solid curve are as described for Fig. 3. 20 Sulentic et al.: Hβ in Intermediate z Quasars: MBH and L/LEdd Fig. 6. MBH computations using same samples as Fig. 3, where RQ and RL sources are indicated by filled and open symbols respectively (upper panel). Triangles indicate sources for which radio data are insufficient to compute a meaningful RK. The lower panel distinguishes MBH for Pop. A (filled symbols) and Pop. B (open symbols). The solid curve is as described for Fig. 3. Fig. 7. Distribution of L/LEdd with z following Fig. 3 with λLλ values derived from the 5100 A rest-frame flux and with α = 0.67. E1 Populations A and B are indicated by filled and open symbols respectively. The solid line traces the minimum L/LEdd for a fixed mass value of 4· 109 M⊙ to observe a quasar above the limiting magnitude of the HE quasar survey mB ≈ 17.5. The dot-dashed lines indicate the same limit but assuming limiting magnitude mB ≈ 16.5.
astro-ph/9710033
1
9710
1997-10-02T21:58:44
Long-Term Flux Monitoring of LSI +61 303 at 2.25 and 8.3 GHz
[ "astro-ph" ]
LSI +61 303 is an exotic binary system consisting of a ~10 Msun B star and a compact object which is probably a neutron star. The system is associated with the interesting radio source GT0236+610 that exhibits bright radio outbursts with a period of 26.5 days. We report the results of continuous daily radio interferometric observations of GT0236+610 at 2.25 and 8.3 GHz from 1994 January to 1996 February. The observations cover 25 complete (and 3 partial) cycles with multiple observations each day. We detect substantial cycle-to-cycle variability of the radio emission characterized by a rapid onset of the radio flares followed by a more gradual decrease of the emission. We detect a systematic change of the radio spectral index alpha which typically becomes larger than zero at the onset of the radio outbursts. This behavior is suggestive of expansion of material initially optically thick to radio frequencies, indicating either that synchrotron or inverse Compton cooling are important or that the free-free optical depth to the source is rapidly changing. After two years of observations, we see only weak evidence for the proposed 4-year periodic modulation in the peak flux of the outbursts. We observe a secular trend in the outburst phases according the the best published ephemeris. This trend indicates either orbital period evolution, or a drift in outburst orbital phase in response to some other change in the system.
astro-ph
astro-ph
Long-Term Flux Monitoring of LSI +61◦ 303 at 2.25 and 8.3 GHz 7 9 9 1 t c O 2 1 v 3 3 0 0 1 7 9 / h p - o r t s a : v i X r a P. S. Ray1, R. S. Foster2, E. B. Waltman3 Code 7210 Remote Sensing Division Naval Research Laboratory Washington, DC 20375 M. Tavani4 Columbia Astrophysics Laboratory Columbia University New York, NY 10027 F. D. Ghigo5 National Radio Astronomy Observatory P.O. Box 2 Green Bank, WV 24944 Received ; accepted 1E-mail: [email protected], NRC Research Associate 2E-mail: [email protected] 3E-mail: [email protected] 4E-mail: [email protected] 5E-mail: [email protected] -- 2 -- ABSTRACT LSI +61◦ 303 is an exotic binary system consisting of a ∼ 10 M⊙ B star and a compact object which is probably a neutron star. The system is associated with the interesting radio source GT 0236+610 that exhibits bright radio outbursts with a period of 26.5 days. We report the results of continuous daily radio interferometric observations of GT 0236+610 at 2.25 and 8.3 GHz from 1994 January to 1996 February. The observations cover 25 complete (and 3 partial) cycles with multiple observations each day. We detect substantial cycle-to-cycle variability of the radio emission characterized by a rapid onset of the radio flares followed by a more gradual decrease of the emission. We detect a systematic change of the radio spectral index α (defined as Sν ∝ να) which typically becomes larger than zero at the onset of the radio outbursts. This behavior is suggestive of expansion of material initially optically thick to radio frequencies, indicating either that synchrotron or inverse Compton cooling are important or that the free-free optical depth to the source is rapidly changing. After two years of observations, we see only weak evidence for the proposed 4-year periodic modulation in the peak flux of the outbursts. We observe a secular trend in the outburst phases according the the best published ephemeris. This trend indicates either orbital period evolution, or a drift in outburst orbital phase in response to some other change in the system. Subject headings: radio continuum: stars -- stars: individual (LSI +61◦ 303) -- stars: binaries -- X-rays: binaries -- 3 -- 1. Introduction The exotic binary system LSI +61◦ 303 (= V615 Cas = GT 0236+610) is one of a remarkable class of X-ray and radio emitting binaries which includes the well-known sources Cir X-1, Cyg X-3 and SS433 (e.g. Hjellming & Johnston 1988). LSI +61◦ 303 is a ∼ 10 M⊙ B0 star at a distance of ∼ 2 kpc (Frail & Hjellming 1991) whose spectrum shows evidence for rapid rotation and a high-velocity equatorial wind (Hutchings & Crampton 1981). The B star is in a binary system with a compact companion, most likely a neutron star. This source is particularly interesting among high-mass binaries because of its strongly variable emission at wavelengths from radio to X-rays and probably γ-rays. Although this source has been the target of a great many observations, the energy source and emission mechanisms responsible for its peculiar behavior remain a mystery. GT 0236+610 is a highly variable radio source which exhibits periodic radio outbursts with a period of 26.496 ± 0.008 d (Taylor & Gregory 1984). However, these outbursts are not stable in phase. Outburst maxima have been seen from phase 0.45 to 0.95, but bright maxima seem to occur near 0.6 (Paredes, Estalella, & Rius 1990). Phase zero is arbitrarily defined to be at JD 2443366.775 (Taylor & Gregory 1982). A ∼ 4 year periodic modulation has been proposed in the amplitude of the radio maxima (Gregory et al. 1989), possibly due to precession of a relativistic jet, or variable accretion from the B star. The modulation has been fit with a sinusoidal function with a period of 1476 ± 7 d (Paredes, Estalella, & Rius 1990) or 1605 ± 24 d (Mart´ı 1993). VLBI observations near maxima show a 2 mas double source implying source size of about 5.5 × 1013 cm (3.7 AU) and an equipartition value of 0.7 G for the magnetic field (Massi et al. 1993). The source shows evidence for rapid expansion ( >∼ 400 km/s), but not for relativistic bulk motions such as those seen in GRS 1915+105, GRO J1655−40, SS433 and Cyg X-3 (see Mirabel & Rodriguez (1994) for a review). -- 4 -- In addition, LSI +61◦ 303 is associated with a weak, variable X-ray source which has been observed with Einstein, ROSAT, and ASCA (Bignami et al. 1981, Goldoni & Mereghetti 1995, Taylor et al. 1996, Leahy, Harrison, & Yoshida 1997). The X-ray source (in the one well-sampled cycle) is brightest (∼ 1.5 × 1033 erg/s, 0.07-2.48 keV, assuming a distance of 2.0 kpc and without correcting for absorption) during the radio quiet phase, although the relationship is poorly determined because of very sparsely sampled X-ray data. The X-ray spectrum is an absorbed power-law with photon index ∼ −1.7 with no evidence for an Iron line (Leahy, Harrison, & Yoshida 1997). LSI +61◦ 303 is exceptional among X-ray binaries in that it is the most probable counterpart to the COS-B and EGRET γ-ray source 2CG 135+01 (Hermsen et al. 1977, Fichtel et al. 1994, Tavani et al. 1996) with an implied luminosity of ∼ 8 × 1034 erg/s above 100 MeV (for a distance of 2.0 kpc). The radio position is between the 68% and 95% confidence level contours of the best γ-ray position determination (Thompson et al. 1995, Kniffen et al. 1997) by EGRET. Both COMPTEL and OSSE have observed emission from this region (van Dijk et al. 1996, ) but have not been able to exclude the nearby QSO0241+622 as the source. In 1994 January, we began dual-frequency radio monitoring observations of LSI +61◦ 303 to follow the radio flux density and spectral index of the source during ASCA X-ray observations (Leahy, Harrison, & Yoshida 1997, Ray et al. 1996). These observations were continued through 1996 February and supported Compton Gamma-Ray Observatory observations (Tavani et al. 1996). Multi-frequency observations are a good way to probe the physical conditions of the source. Because of the large cycle-to-cycle variations, it is imperative to track the state of the radio emission during high-energy observations. This paper summarizes the radio observations (§2) and discusses implications for models (§3) of this unique system. -- 5 -- 2. Observations The observations reported here were made several times per day (with a few gaps) from 1994 January 27 to 1996 February 23. The monitoring was performed with the National Radio Astronomy Observatory Green Bank Interferometer (GBI) in Green Bank, West Virginia. The GBI consists of two 26 m antennas on a 2.4 km baseline each of which has a pair of cooled receivers to simultaneously receive signals at 2.25 and 8.3 GHz. The correlators process 35 MHz of bandwidth at each frequency. After being shut down due to lack of funding from 1996 April through November, the GBI is currently being operated by NRAO and NASA primarily for radio monitoring of X-ray sources. All data taken is made available to the public immediately via the WWW (http://www.gb.nrao.edu/gbint/GBINT.html). Ten minute scans on LSI +61◦ 303 were performed 1 -- 10 times per day within ±5 hours of source transit. Each scan consists of a vector sum of the correlator amplitude and phase from 30 s integrations. Measured correlator amplitudes are converted to flux densities by comparison to standard, regularly observed calibrators. The resulting data set consists of 6628 flux density measurements of the source at two frequencies over 2.07 yr. Twenty-five outbursts have nearly complete coverage and three more have partial coverage (due to hardware or scheduling problems). The receivers have a system temperature of ∼ 40 K, and individual flux density measurements have errors of: σ2.25GHz = h(0.006)2 + (0.014S2.25GHz)2i1/2 σ8.3GHz = h(0.009)2 + (0.049S8.3GHz)2i1/2 Jy, Jy. (1) (2) Note that the constant term has been significantly lowered since 1989 (Fiedler et al. 1987) by the installation of new cooled HEMT receivers. A fraction of this error estimate is due to a difficult to calibrate systematic gain variation in the system as a function of hour angle. The gain variations are likely due to a combination of changing aperture efficiency due to -- 6 -- deformations of the antenna surface, and weather-dependent phase coherence problems. We have attempted to correct for the gain variations by fitting the hour-angle dependent effect observed in a constant source with a second order polynomial. This was used to correct the observed fluxes for LSI +61◦ 303. The magnitude of the effect at ±5h is about 7% and 20% for 2.25 and 8.3 GHz, respectively. The daily average flux density of LSI +61◦ 303 at 2.25 and 8.3 GHz as well as the calculated spectral index are shown in Figure 1 and 2. The error bars on this and other figures are ±1σ. There is considerable variation from outburst to outburst, but inspection of the light curves reveals that the higher frequency tends to peak earlier and fade more quickly than the lower frequency. By cross correlating the light curves at each frequency, we measure a mean time delay between the 8.3 GHz outburst and the 2.25 GHz outburst of 0.4 days with a large RMS scatter of 0.5 days. In addition, many of the outbursts show two or more peaks during the decline of the main outburst. 2.1. Spectral evolution An important feature of this monitoring program is the simultaneous dual-frequency observations. This allows accurate measurements of the spectral index as a function of time (orbital phase) in this highly variable source. From two frequency data, we cannot tell anything about the shape of the spectrum, but we can infer a spectral index (α) assuming that Sν ∝ να. During typical outbursts we observe the spectral index to rise to a peak with α between 0 and 0.5 then decay back to the quiescent value of α ∼ −0.4. Our mean light curves with derived two-frequency spectral index between 2.25 and 8.3 GHz are shown in Figure 3. Note, however, that the actual radio light curve and spectral index behavior may substantially change from cycle to cycle. The changing spectral index is a result of the typical outburst profile decaying more slowly at 2.25 GHz than at 8.3 GHz. By fitting -- 7 -- Gaussian profiles to the outbursts, we find that the characteristic widths of the 2.25 GHz profiles are 28% larger than that of the 8.3 GHz profiles. The only previous measurements of the spectral index evolution during an outburst are the statistical results of Taylor & Gregory (1982) and one outburst observed by Taylor & Gregory (1984). Both of these results reported two-frequency spectral indices between 5 and 10 GHz. Realistic models for the radio spectrum of this source are not simple power laws (Mart´ı 1993), and free-free absorption will be much more important at 2.3 GHz than at 5 GHz, so direct comparison of our results to previous ones is somewhat difficult. However, in the one outburst with two-frequency measurements, the spectral index is slightly negative at the peak and the decay phase of the outburst is consistent with a constant spectral index. Also, their statistical results show a strong peak in spectral index of nearly α = 2 at around phase 0.4. We see no evidence for positive spectral indices except at the peak of the outbursts. There are several energy loss mechanisms which may be important in an expanding cloud of relativistic electrons including adiabatic expansion, synchrotron emission, and inverse Compton scattering (Pacholczyk 1970). Each of these affects the shape of the electron energy distribution function and thus the observed spectral index in different ways. Inverse Compton scattering and synchrotron losses both have dE/dt ∝ E 2 which result in a spectral index changing toward more negative values during the decay phase of the outburst. Adiabatic expansion implies dE/dt ∝ E which results in a constant spectral index, which is not seen in the decay phase of the outbursts. The multiple peaks in the light curves of the outbursts indicate continuing particle injection. It is clear that the production of energetic particles in this source cannot be modeled as a simple delta function of time. The observed spectral index evolution may also be (at least partially) the result of the source being embedded in a high-optical-depth envelope such as the ionized stellar wind -- 8 -- of the B star. In this case, the emitted radio emission will be modified by the free-free opacity of the wind. This envelope may be completely opaque near periastron and the optical depth may remain greater than unity for 6 -- 22 days (Massi et al. 1993). Thus, the observed outburst onset may occur well after the initial acceleration event. Since the wind opacity is a strong function of frequency (the optical depth τ = T −1.5 e ν −2.1EM, where the emission measure EM = R n2 to brighten first if this is an important effect, as is observed. eds; Pacholczyk 1970), one would expect the higher frequency 2.2. Long-term periodicity The proposed long-term periodicity has previously been fitted with rather sparse data: 14 peak flux measurements covering only 2 -- 3 periods of the modulation (Mart´ı 1993). Possible explanations include: variable beaming from a precessing relativistic jet (Gregory et al. 1989), variable wind velocity from the B star resulting in a variable accretion rate onto the neutron star (Mart´ı & Paredes 1995), or precession of the B star in the neutron star magnetic field (Lipunov & Nazin 1994). The peak fluxes (at 8.3 GHz) of the 18 outbursts observed by the GBI are shown in Figure 4 along with the previous observed outburst maxima (observed at 5, 8, or 10 GHz) used to fit the sinusoidal model by Paredes, Estalella, & Rius (1990) and Mart´ı (1993). An expanded view of only the new data is shown in Figure 5. Both figures include the previously published models for the modulation. Our data rule out the four-year modulation model of Paredes, Estalella, & Rius (1990), but are marginally consistent with the model of Mart´ı (1993) in which Smax = A cos [ 2π P (t − t0)] + B, (3) where A = 114 ± 8 mJy, t0 = JD 2443464 ± 100, A = 114 ± 9 mJy, and P = 1605 ± 24 d. -- 9 -- However, no evidence for any secular trend in the peak flux data at 2.25 GHz (shown in Figure 6) is observed. 2.3. Ephemeris Figure 7 shows the phase of the outburst maxima as a function of date. These phases were determined by cross correlating the observed fluxes with a template. Due to incomplete sampling of the fluxes, and occasional high points which may be due to radio interference, fitting outburst phases by cross correlation is more robust than simply taking the time of the measurement with the highest flux. The offset between the two frequencies may be partially due to a different template with a different fiducial point being used for each frequency. A secular trend in the outburst phases is evident. To characterize this effect, we have fit the outburst times of the 8.3 GHz data to a linear model tn = t0 + n × P, (4) where tn is the peak time of the nth outburst. Fitting for t0 and P using 26 measured outburst times yields t0 = JD2449393.5 ± 0.4, and P = 26.69 ± 0.02 days. With these parameters, equation 4 provides a good ephemeris for recent outburst maxima. The RMS difference between the model and the actual outburst times is 1.1 days. Since the differences of the measured outburst times from a linear model could have contributions from many sources including intrinsic variability, cross correlation errors due to shape variations from outburst to outburst, and normal measurement errors, the standard deviations used in the fitting process were calculated from the RMS residuals after fitting a linear model to the data. We note that our best estimate of the period is significantly different (> 9σ) from the -- 10 -- best previously published value of 26.496 ± 0.008 days (Taylor & Gregory 1984). There are several possible explanations for this result. Either there has been real orbital period evolution in the 10 -- 12 years since the previous result was published, or the outburst peaks tend to move around in orbital phase in response to some other parameter in the system, such as the B star mass-loss rate, or a precession period. This other parameter may also be related to the possible modulation in the outburst maxima discussed above. The required orbital period derivative to explain currently observed period is (3.5 ± 0.03) × 10−5, consistent with the upper limit of 1.6 × 10−4 from Taylor & Gregory (1984). The timescale implied by P/ P is 2 × 103 yr, two orders of magnitude less than that observed in the two HMXB systems (Cen X-3 and SMC X-1) with measured orbital period derivatives. Thus the interpretation as true orbital period change seems unlikely. These two models predict rather different behavior over the next several years. Real orbital period evolution will result in a continued gradual lengthening of the period while a drift in the outburst orbital phase predicts that the phase should recover back to 0.6, possibly in response to the 4-year modulation. A complete analysis of all published data over the 20 years since the source was discovered is in progress (M. Peracaula, personal communication) which may resolve this question. 3. Discussion Most models of this source share the feature that the radio emission is due to synchrotron emission from relativistic electrons moving in a magnetic field. However, there is no consensus as to the means of production of these particles, or their relation to the high-energy emission. Several mechanisms for creating the required relativistic electron population which have been proposed for LSI +61◦ 303 are summarized below to demonstrate the wide variety of possible mechanisms. -- 11 -- Maraschi & Treves (1981) (see also Tavani 1995) proposed that the relativistic electrons are from a young pulsar wind, and that the X-ray and γ-ray emission comes from Compton scattering of optical photons from the primary B star at a shock boundary between the pulsar wind and the stellar wind. Taylor & Gregory (1984) proposed that super-Eddington accretion onto the neutron star near periastron results in luminosity-driven shocks which may account for the particle production. Paredes et al. (1991) proposed an adiabatically expanding synchrotron source with prolonged injection of relativistic particles. Lipunov & Nazin (1994) proposed that relativistic electrons from a pulsar wind are captured by the magnetosphere of the B star and cooled by synchrotron losses. Recent work (e.g. Campana et al. 1995) has focused on the accretion regimes of rapidly rotating magnetized neutron stars in eccentric binaries. Similarly, Zamanov (1995) proposed that the outbursts occur at the transition of a magnetized neutron star from a propeller phase to an ejector phase as the mass transfer rate changes in an eccentric orbit. It is likely that LSI +61◦ 303 is a close relative of the two recently discovered radio pulsars in B-star binaries (PSR J0045−7319 and PSR B1259−63) and the 69-ms X-ray pulsar A 0535−668. All three systems contain a rapidly rotating, highly magnetized neutron star orbiting a massive main-sequence companion. The primary differences between these systems are the mass-loss rate of the OB star, the geometry of the orbit, and the energy flux of the pulsar wind. These three factors determine which accretion regime the system is in. For J0045−7319, M < 10−10 M⊙ yr−1 (Kaspi, Tauris, & Manchester 1996) and for B1259−63, 100 < (v/10 km s−1)( M /10−8 M⊙ yr−1) < 104 (Tavani & Arons 1997). Thus, these two systems have fairly weak stellar winds, active pulsars producing strong outflows, and B1259−63 does not get very close to the companion at periastron, allowing both to remain in the radio pulsar regime since matter is unable to accrete onto the pulsar magnetosphere. On the other hand, A 0535−668, may have a weaker pulsar wind, and might encounter the thick equatorial disk of the Be star near periastron. This allows -- 12 -- accretion at about the Eddington rate to occur and produces a powerful X-ray pulsar. For LSI +61◦ 303, a reasonably good estimate for the mass loss rate from the companion is M ∼ 10−7 M⊙ yr−1 (Waters et al. 1988). Campana et al. (1995) suggest that while in LSI +61◦ 303 the stellar wind ram pressure may be sufficient to overcome the pulsar wind pressure and accrete onto the magnetosphere, the angular velocity might be super-Keplerian resulting in the matter being ejected by the propeller mechanism. Alternately, a relatively young pulsar may power the high energy emission from LSI +61◦ 303 (Maraschi & Treves 1981, Tavani 1995) by a shock mechanism which can be shown to be in agreement with the intensity and spectral ranges observed for LSI +61◦ 303 in the X-ray range and for 2CG 135+01 in the gamma-ray range (Tavani et al. 1996). We note that the long-timescale monitoring at 2.25 GHz adds significant new information compared to previous observations of LSI +61◦ 303 that were made a much higher frequency. This can be immediately realized by considering the relevance of Figure 3 giving the mean spectral index. The flattening of the spectrum at phases coincident with the peak radio fluxes is evident. Free-free absorption along the line-of-sight or synchrotron self-absorption may be responsible for the spectral flattening near phase 0.6. If we assume a critical absorption frequency at, say, 5 GHz, and a gas temperature of 104 K, we obtain a free-free optical depth larger than unity for an emission measure EM ≥ 2 · 1025 (cgs units). We can also estimate the source size L if the spectral flattening is caused by synchrotron self-absorption. For an estimated radio luminosity of ∼ 2 × 1030 erg s−1, a local magnetic field of ∼ 1 G, we deduce L ∼ 7 × 1012 cm. We notice that the estimated value of L is not dissimilar from the source size adopted in the adiabatic/synchrotron cooling model by Paredes et al. 1991. The integration time of 10 minutes prevents us from placing a strong limit on the source size from rapid variability, but we do find that the source varies by more than a factor of two within 15 minutes. This is fully consistent with the timescale L/c ∼ 4 min. -- 13 -- 4. Summary After 20 months of continuous radio observations of GT 0236+610 we find only weak evidence for long-term modulation of peak flare amplitudes on a four year time scale. We observe spectral index variability during the outbursts which implies that either adiabatic expansion is not the dominant energy loss mechanism in the expanding plasmon or that the optical depth due to free-free opacity is changing during the outburst. We also observe an apparent lengthening of the period between outbursts which is fit with a period of 26.69 ± 0.02 days. This may be due to an orbital period derivative of (3.5 ± 0.03) × 10−5 or due to the outburt phases being modulated in response to some other parameter of the system. A single self-consistent model for LSI +61◦ 303 remains elusive. Critical questions include: Are the X-rays and γ-rays produced by the same population of relativistic electrons? Is the compact object a young pulsar? What is the orbital eccentricity? A critical diagnostic for distinguishing between the various models is the time dependence and correlation between the radio, X-ray and γ-ray emission. Multiwavelength observations with GRO, XTE, and SAX will be important. A collection of resources to facilitate multifrequency observations of this source is available on the WWW at http://www.srl.caltech.edu/personnel/paulr/lsi.html. Radio observations at more than two frequencies will also be useful to separate the effects of free-free absorption, synchrotron and inverse Compton losses in the evolution of the source spectrum. Finally, a more accurate determination of the Keplerian orbital parameters is badly needed. During this work, the Green Bank Interferometer was operated by the National Radio -- 14 -- Astronomy Observatory for the Naval Research Laboratory and was supported by the Office of Naval Research. Partial support for the observations and analysis of this source were supported by a National Aeronautics and Space Administration interagency fund transfer. Basic research in precision pulsar astrophysics at the Naval Research Laboratory is supported by the Office of Naval Research. A portion of this work was performed while one of the authors (PSR) held a National Research Council-NRL Research Associateship. This research made use of the Simbad database, operated at CDS, Strasbourg, France. -- 15 -- REFERENCES Bignami, G. F., Caraveo, P. A., Lamb, R. C., Markert, T. H., & Paul, J. A. 1981, ApJ, 247, L85 Campana, S., Stella, L., Mereghetti, S., & Colpi, M. 1995, A&A, 297, 385 Fichtel, C. E. et al. 1994, ApJS, 94(2), 551 Fiedler, R. L. et al. 1987, ApJS, 65, 319 Frail, D. A. & Hjellming, R. M. 1991, AJ, 101, 2126 Goldoni, P. & Mereghetti, S. 1995, A&A, 299, 751 Gregory, P. C., Xu, H.-J., Backhouse, C. J., & Reid, A. 1989, ApJ, 339, 1054 Hermsen, W. et al. 1977, Nature, 269, 494 Hjellming, R. M. & Johnston, K. J. 1988, ApJ, 328, 600 Hutchings, J. B. & Crampton, D. 1981, PASP, 93, 486 Kaspi, V. M., Tauris, T. M., & Manchester, R. N. 1996, ApJ, 459, L717 Kniffen, D. A. et al. 1997, ApJ, 486, 126 Leahy, D. A., Harrison, F. A., & Yoshida, A. 1997, ApJ, 475, 823 Lipunov, V. M. & Nazin, S. N. 1994, A&A, 289, 822 Maraschi, L. & Treves, A. 1981, MNRAS, 194, 1P Mart´ı, J. 1993. PhD thesis, Universitat de Barcelona Mart´ı, J. & Paredes, J. M. 1995, A&A, 298, 151 -- 16 -- Massi, M., Paredes, J. M., Estalella, R., & Felli, M. 1993, A&A, 269, 249 Mirabel, I. F. & Rodriguez, L. F. 1994, in Proceedings of the 17th Texas Symposium on Relativistic Astrophysics, Annals of the New York Academy of Sciences Pacholczyk, A. G. 1970, Radio Astrophysics, (San Francisco: Freeman) Paredes, J. M., Estalella, R., & Rius, A. 1990, A&A, 232, 377 Paredes, J. M., Mart´ı, J., Estalella, R., & Sarrate, J. 1991, A&A, 248, 124 Ray, P. S., Foster, R. S., Waltman, E. B., Ghigo, F. D., & Johnston, K. J. 1996, in Radio Emission from Stars and the Sun, ed. A. R. Taylor & J. M. Parades, volume 93 of ASP Conference Series, 249 Strickman, M. S., Tavani, M., Coe, M. J., Steele, I. A., Fabregat, J., Mart´i, J., Paredes, J. M., & Ray, P. S. 1998, ApJ. submitted Tavani, M. 1995, in The Gamma-Ray Sky with GRO and SIGMA, ed. M. Signore, P. Salati, & G. Vedrenne, (Dordrecht: Kluwer), 181 Tavani, M. & Arons, J. 1997, ApJ, 477, 439 Tavani, M. et al. 1996, A&AS, 120, 243 Taylor, A. R. & Gregory, P. C. 1982, ApJ, 255, 210 Taylor, A. R. & Gregory, P. C. 1984, ApJ, 283, 273 Taylor, A. R., Young, G., Peracaula, M., Kenny, H. T., & Gregory, P. C. 1996, A&A, 305, 817 Thompson, D. J. et al. 1995, ApJS, 101, 259 van Dijk, R. et al. 1996, A&A, 315, 485 -- 17 -- Waters, L. B. F. M., Taylor, A. R., van den Heuvel, E. P. J., Habets, G. M. H. J., & Persi, P. 1988, A&A, 198, 200 Zamanov, R. K. 1995, MNRAS, 272, 308 This manuscript was prepared with the AAS LATEX macros v4.0. -- 18 -- Fig. 1. -- Flux history of LSI +61◦ 303 at 2.25 and 8.3 GHz from 27 January 1994 to 8 March 1995. The arrows mark phase 0.6 of the published radio ephemeris (Taylor & Gregory 1984). The triangles mark the outburst phase as determined by cross-correlating with a template (see Figure 7). The triangles are missing where there was insufficient data to get a good peak. -- 19 -- Fig. 2. -- Continuation of Figure1 from 8 March 1995 to 26 February 1996. -- 20 -- Fig. 3. -- Phase averaged fluxes and spectral index of LSI +61◦ 303. -- 21 -- Fig. 4. -- Peak outburst flux of LSI +61◦ 303. The dotted line is the best-fit model of Paredes, Estalella, & Rius (1990). The dashed line is the more recent model of Mart´ı (1993). Stars and upward pointing arrows (lower limits) are the previously published data on which the model was based. All data beyond day 9000 are from this work. -- 22 -- Fig. 5. -- Expanded version. The dotted line is the best-fit model of Paredes, Estalella, & Rius (1990). The dashed line is the more recent model of Mart´ı (1993). -- 23 -- Fig. 6. -- Same as Figure 5 except using the 2.25 GHz data. -- 24 -- Fig. 7. -- Outburst phase determined by cross-correlating with a template.
0804.0410
1
0804
2008-04-02T18:28:24
The Galactic gamma-ray club
[ "astro-ph" ]
The exclusive Galactic gamma-ray club has opened up to new members. Supernova remnants, pulsar wind nebulae, and massive binary systems hosting a compact object have recently joined the young pulsars as firmly established sources of gamma rays in the Milky Way. Massive young stellar clusters are on the waiting list to join the club. Only the fine imaging recently obtained at TeV energies could resolve specific sources. The samples are sparse, but raise exciting questions. The jet or pulsar-wind origin of the emission in binaries has been hotly debated, but it seems that both types of systems have been recently detected. The nature of the radiation in shock accelerators is still questioned: do nuclei contribute a lot, a little, or not to the gamma rays and what energy do they carry away from the shock budget? The acceleration process and the structural evolution of the pulsar winds are still uncertain. The magnetic field distribution in all these systems is a key, but poorly constrained, ingredient to model the multi-wavelength data, particle transport and electron ageing. It must, however, be determined in order to efficiently probe particle distributions and the acceleration mechanisms. The source samples soon to be expected from GLAST and the Cherenkov telescopes should bring new valuable test cases and they will, for the first time, shed statistical light on the collective behaviour of these different types of accelerators.
astro-ph
astro-ph
30TH INTERNATIONAL COSMIC RAY CONFERENCE The Galactic Gamma-Ray Club ISABELLE GRENIER Laboratoire AIM, CEA/DSM-CNRS-Université Paris Diderot, Service d’Astrophysique, CEA Saclay, 91191 Gif/ Yvette, France [email protected] Abstract: The exclusive Galactic γ-ray club has opened up to new members. Supernova remnants, pulsar wind nebulae, and massive binary systems hosting a compact object have recently joined the young pulsars as firmly established sources of γ rays in the Milky Way. Massive young stellar clusters are on the waiting list to join the club. Only the fine imaging recently obtained at TeV energies could resolve specific sources. The samples are sparse, but raise exciting questions. The jet or pulsar-wind origin of the emission in binaries has been hotly debated, but it seems that both types of systems have been recently detected. The nature of the radiation in shock accelerators is still questioned: do nuclei contribute a lot, a little, or not to the γ rays and what energy do they carry away from the shock budget? The acceleration process and the structural evolution of the pulsar winds are still uncertain. The magnetic field distribution in all these systems is a key, but poorly constrained, ingredient to model the multi-wavelength data, particle transport and electron ageing. It must, however, be deter- mined in order to efficiently probe particle distributions and the acceleration mechanisms. The source samples soon to be expected from GLAST and the Cherenkov telescopes should bring new valuable test cases and they will, for the first time, shed statistical light on the collective behaviour of these different types of accelerators. GeV and TeV gamma-ray sources Major advances in our understanding of the universe have often come from improving angular resolution at all wavelengths. The prowess of achieving several arc- minute resolution at TeV energies with the Cherenkov telescopes, and soon at GeV energies with GLAST, indeed opens a new era in γ-ray astronomy. The very secluded club of identified γ-ray sources, which has only accepted young pulsars and blazars for decades, has recently expanded to supernova remnants, pulsar wind nebulae, and γ-ray binaries. The high-energy facets of these objects of course raise new and exciting questions that I will try to briefly review. Yet, such a resolution does not compare with the imaging capabili- ties at lower wavelengths and deciding between true source identification and mere spatial coincidence in the crowded environments along the Milky Way will still be a key issue for years to come. Source detection at GeV energies, unlike at TeV ener- gies, has to fight against the intense and highly struc- tured interstellar background that results from cosmic- ray interactions in the interstellar gas and soft radiation field. This background is of particular interest to the ICRC scientists because it probes the cosmic-ray den- sity through the Galaxy, yet, with respect to point sources, its imperfect modelling induces systematic errors in source significance, flux, location, and spatial extension. Figure 1a illustrates the difference between the 3rd EGRET (3EG) catalogue [1] and the revised THE GALACTIC GAMMA-RAY CLUB Figure 1: Distribution in Galactic coordinates of (a) the revised EGRET point sources detected at energies above 100 MeV on top of an improved interstellar emission model and using 9 years of EGRET data. Blue dots, open circles, and black dots respectively mark the confirmed 3EG sources, the unconfirmed ones and new sources; (b) the sources currently de- tected at TeV energies by the Cherenkov telescopes (courtesy of Robert Wagner). one presented at the conference [2]. The use of nine years of EGRET data instead of four at the time of the 3EG catalogue has allowed detection of 31 new sources, but the fact that as many as 107 former sources have not been confirmed in the new analysis is primar- ily due to the improved background that includes new CO clouds at medium latitudes, new HI maps corrected for stray radiation, and new maps for the dark gas phase in the nearby clouds. The latter is dark in the sense that it is not properly traced by HI and CO emis- sion. This gas phase has been found in the nearby Gould Belt clouds [3]. It appears as an excess of both dust emission and diffuse γ radiation over what the HI and CO data should yield. Both excesses are strongly correlated and form spatially coherent structures. So, both dust and cosmic rays trace some 'dark' gas that forms an extended layer at the transition between the dense CO cores and the outer HI envelope of the clouds. The dark gas column-densities compare with that measured in HI and CO, so it provides both γ-ray intensity and structure that were not accounted for in the earlier background models. Because source detec- tion methods in γ rays search for significant and point- like photon excesses above the predicted background, an ensemble of point sources with the wide EGRET PSF would compensate for the missing cloud structures and yield an excellent fit to the data. These 3EG sources have not been confirmed in the new analysis; they consist of only six sources with a likely AGN counterpart and 101 unidentified sources, in particular the faint and persistent ones that have long been spa- tially associated with the Gould Belt [4, 5, 6]. The im- proved GLAST resolution [7] will soon provide much more accurate maps of the diffuse interstellar emission, but the same uncertainties in gas tracers and in the cosmic-ray flux pervading the distant clouds will still impact source detection and localization near the detec- tion threshold. At TeV energies, the interstellar inten- sity falls off more rapidly than most source spectra. 30TH INTERNATIONAL COSMIC RAY CONFERENCE 2007 This is why we presently detect more TeV than GeV sources at low latitude, as illustrated in Figure 1b. The concentration of sources at l < 60° is due to the cover- age of the HESS survey. Irradiated clouds It is, however, striking that the vast majority of TeV sources have radial extensions of several arc minutes that correspond to diameters of 20 or 30 parsecs in the inner spiral arms and molecular ring of the Milky Way. Pulsar wind nebulae have been seen as extended sources of TeV radiation and will be reviewed below, yet an exciting alternative is that some of these sources trace where the cosmic-ray sources dwell, in other words that they are due to pockets of enhanced cosmic- ray flux diffusing away from their source(s) and irradi- ating the gas on their way. The HESS J1800-240 A and B sources may serve as useful examples [8]. They cor- relate well with the integrated CO intensity map of massive (0.5-1.5 105 M⊙) clouds at a distance of 2 and 4 kpc in the Sagitttarius and Scutum-Crux arms. The TeV flux requires 10 to 30 times the local cosmic-ray density if it is due to neutral pion decay. This is much more than the modest increase in cosmic-ray density expected in the inner Galaxy from large-scale studies of the interstellar GeV emission [9]. The increase with respect to the flux at the solar circle does not exceed a factor of 1.5 or 2. Another source in this direction, HESS J1801-233, correlates with CO intensity. It coin- cides with the rim of the W 28 supernova remnant, but may not be related to it. The shock wave from W 28 is known to run into 2000 solar masses of molecular gas which could be heavily irradiated if acceleration at the shock manages to remain active in the dense medium or if particles accelerated earlier and trapped down- stream still leak out of the remnant. Yet, the HESS source spatially overlaps with only half of the shocked gas. The other half exhibits almost as much mass, but is not detected. It is not obvious that the small mass defi- cit compared to the TeV-bright side can explain the lack of emission. The TeV source correlates with a fore- ground or background CO complex of about 2 104 M⊙ which may or may not be spatially related to the shocked clouds. It is difficult to tell in this crowded direction. Again, 10 to 30 times the local cosmic-ray density is needed to explain the flux in terms of pion decay in this cloud. It could be provided by W 28 or by other sources embedded in the cloud. The three HESS sources associated with CO exhibit equivalent spectra, with soft photon indices of 2.5-2.7 typical of interstellar emission. When protons of energy Ep diffuse for a time τ away from their source, with an interstellar diffusion coefficient D ≈ 1024 (Ep/10 GeV)0.6 m2 s-1, the pion-decay spectrum breaks at an energy Eγ ≈ 0.17 Ep that shifts with distance L from the source as L2 ≈ 6 D(Ep).τ [10]. Lower energy particles cannot reach that far in the same amount of time, so the spectrum drops at lower energies. The spectrum above the broad peak falls as E-2.6 for an E-2 source injection spectrum. The flux also peaks around this characteristic timescale τ(L, E) since spherical expansion dilutes the flux on longer time scales (in the same energy band). The clouds of 104 to 105 solar masses that we are dis- cussing have typical sizes of 20 to 60 pc, respectively, near virial equilibrium. The soft E-2.7-2.5 spectra seen by HESS require a peak proton energy of 600 GeV or less, that can be seen out to 50 pc after a thousand years of diffusion. The flux estimates derived by Aharonian & Atoyan [10] agree reasonably well with the observed ones. More distant clouds would exhibit too hard spec- tra and undetectable fluxes at TeV energies. The same 50 pc cloud complex would yield a flux detectable by GLAST after 50 kyr, with a flat energy spectrum near the peak that would allow separation from the bulk of the diffuse interstellar emission. It should be noted, however, that the diffusion properties in the interstellar medium are not well characterized, especially in dense environments where lower diffusion coefficients may prevail. For a 10 times lower coefficient, one would have to wait for 10 kyr for the higher energy particles to diffuse out to get a soft enough TeV spectrum in the 50 pc large cloud, but not too long because the flux drops by an order of magnitude after 50 kyr [11]. Es- tablishing several examples of ‘over-irradiated’ clouds at GeV and TeV energies would therefore prove very useful to explore the diffusion properties as well as to locate where the cosmic-ray sources live. Another example of cloud irradiation has been pro- posed toward the Galactic centre [12] where the corre- lation between the molecular gas distribution (as traced by CS) and the TeV intensity profile may suggest an overabundance of freshly accelerated cosmic rays. Given the turmoil in the inner 150 pc of the Milky Way, it is unfortunately one of the most difficult places to interpret spatial distributions. Given the tortured struc- ture of the magnetic field and its possible strength of 100 nT to account for the bright radio arcs, one cannot use the local value of the diffusion coefficient. The latter should also largely differ in and out of the plane. One may ask how the particles can efficiently cross the radio arc region and numerous other magnetic threads that stretch perpendicular to the plane between the central TeV source assumed by the authors and the peaks detected at l = ± 0.6°. An estimate of the fraction of particles that escape out vertically would be useful. THE GALACTIC GAMMA-RAY CLUB A second problem is that the secondary electron- positron flux produced in the hadronic interactions fails by nearly a factor 100 to mach the radio emission found in Sgr B2. It does not explain either the unusu- ally high ionization level measured in the cloud (∼ 4 10-16 s-1) [13]. A pure inverse Compton origin of the TeV data is difficult to reconcile with the X-ray and radio fluxes in the cloud and the required magnetic field is too low [13]. On the other hand, the distribution of the TeV emission in the whole region follows that of the fluorescent 6.4 keV line from neutral iron which is probably excited in the clouds by very low energy cosmic rays [14]. So these clouds, unlike others at l > 1°, appear to harbour an excess of long-lived, low- energy cosmic rays as well as fresh high-energy ones. A passing wave of particles produced by Sgr A East or another central supernova remnant of the past millennia is not the end of the story. Gamma-ray champagne bubbles? Alternatively, the TeV intensity around Sgr A* corre- lates with the position of the very young and incredibly active star-forming regions of the Arches, Quintuplet, and Sgr B2 clusters, as well as the less prominent, but still very active, Sgr B1 and C regions. TeV radiation has also been detected toward two other young and massive stellar clus- ters: by HESS to- ward Westerlund 2 in the giant RCW49 HII region [15], and by Milagro and MAGIC on the edge of Cygnus OB2 [16, 17]. These sources, J1023-575 HESS a n d T e V J2032+4130, have not been identified with other likely γ- ray emitters such as supernova remnants, pulsars or their wind nebulae. They do not coincide with EGRET sources. These spatial coincidences raise the interesting possibility of efficient particle accelera- tion in these young environments. Coincidences between COS-B sub-GeV sources and OB associations, together with anomalies in the cosmic-ray composition, had prompted the idea of SNOBs [18] in which ions would be first injected and accelerated by the supersonic winds of massive stars and then by the shock wave of a nearby supernova remnant. The recent measurements by ACE of isotopic ratios in the local cosmic rays point to an acceleration site inside OB associations to explain the presence of about 20 % Wolf-Rayet material in the composition [19]. The lack of 59Ni also suggests a period of at least 105 years between nucleosynthesis and acceleration [20], supporting the two-step scenario of SNOBs. The cloud complex that has given birth to the cluster pro- vides targets for hadronic interactions to shine in γ rays. The power released by OB associations in the combined form of stellar winds and supernova shocks is large enough to sustain the required γ-ray luminosi- ties for standard mechanical-to-cosmic-ray energy con- version efficiencies of a few percent. EGRET, however, has seen only few sources toward OB associations (in the Cygnus region, Car OB 1b and 2, Sco OB2d, and Gem OB1) and they may only be chance co- incidences in these crowded directions. associa- Potential tions with TeV sources appear to be more promising, but different because the related OB clus- ters rather are unique in stellar content and youth [21, 22, 23, 24]. Figure 2: significance contours of the HESS J1023-575 source (7 and 9 σ as dashed and solid curves, respec- tively) overlaid on the Spizter and Chandra images of Westerlund 2 and its HII region RCW 49. The PAH infrared emission (in grey) outlines the cloud highly perturbed response to the ionization fronts and winds from the very massive stars. The X rays (in colour) reveal the compact and massive stellar cluster that powers the intense activity in this highly obscured re- gion. 30TH INTERNATIONAL COSMIC RAY CONFERENCE 2007 Figure 3: location and extent of the TeV J2032+4130 source (red dot, as seen by MAGIC [17]) with respect to the massive Cyg OB2 cluster (white circle, [23]) breaking out in a CO shell (green ellipse). The parent cloud and ionized gas are shown as CO contours (red) overlaid on the 1420 MHz intensity map [25]). The EGRET source EGR J2033+41 is centred within the yellow error circle, but any extended source within the dashed yellow circle would not be resolved by EGRET. They are referred to as ‘super star clusters’. They result from extraordinary bursts of star formation that have converted 103.7 to 104.6 solar masses of gas in stars heavier than the Sun. They are extremely rich in mas- sive stars (e.g. 150 O stars in the Arches, 120 ± 20 O stars in Cyg OB2, down to > 12 0 stars in the more modest Westerlund 2 cluster). The huge ionizing pho- ton fluxes of 1050.8 to 1051.6 s-1 have carved out wide HII regions around them. These photons can efficiently produce pairs on the TeV γ rays, but one can easily check that, given the cluster sizes, the photon densities are not lethal to γ rays except within light minutes of one of the massive members [26]. The inverse Comp- ton scattering of all the starlight can double the amount of interstellar GeV γ rays in the cluster vicinity. In the case of Cyg OB2, it would form a 1° wide source of γ rays with a flux near the threshold of the GLAST tele- scope [27]. All these clusters have manufactured giant stars of 85 M⊙ in Westerlund 2 and over 100 M⊙ in the Arches and Quintuplet. The large number of LBV and WR members that are still alive attests of their youth (5% of all known Galactic WR in the Arches, > 17 in the Quintuplet, 3 in Cyg OB2, and 2 in Wester- lund 2). Ages of 2.5 ± 0.5, 2, and 2-3 Myr have been derived for the Arches, Cyg OB2, and Westerlund 2, respectively [22, 28, 29]. Little is known from Sgr B2 since the stars still hide in their parent cloud, but the presence of 50 compact HII regions strongly suggests it is another Arches cluster in the making. The Quintuplet is slightly more evolved, with an age of 4 ± 1 Myr. If these systems turn out to be responsible for the TeV emission, their extreme youth calls for another scenario than the more evolved case of SNOBs because of the lack of supernovae. Even if remnants could easily hide in the intense free-free radio emission, the clusters are too young to have produced several supernovae. Me- chanically, these young systems are almost as powerful as SNOBs. Cesarsky & Montmerle [30] have estimated that stellar winds from WR and O stars dominate the cluster power during the first few million years, espe- cially in the case of massive associations with stars heavier than 30 M⊙ as observed in the super-star clus- ters. The power exceeds 1031 W if the cluster hosts 40 stars with mass-losses > 10-5 M⊙ yr-1 and terminal velocities of order 2500 km/s. Supernovae hardly take over the overall budget when the cluster is 5 or 6 mil- lion years old. Whereas Fermi acceleration by multiple random shocks from supernova remnants in more evolved OB associations has been studied to account for cosmic-ray energies and composition near or be- yond the knee [31], the shock sizes, velocities, and separation in space and time do not apply to the denser, more turbulent case of the colliding winds from tens of massive stars in the compact configuration of the super clusters. Losses in the dense bubble shells also limit the acceleration. Figure 2 shows how intensively the stars have ionized and sculpted the gas around them in Westerlund 2. A ridge of radio emission follows the rim of the bubble centred on the main cluster and along its interface with the bubble blown by WR 20b, but it is dominated by free-free emission. Particle acceleration has been observed as synchrotron radio emission in the colliding winds of massive stars within binary systems. It should yield detectable fluxes of γ rays, if not at TeV energies because of the heavy toll of pair absorption on the intense stellar radiation field, at least to tens of GeV energies [32]. However, the large spatial extent of THE GALACTIC GAMMA-RAY CLUB HESS J1023-575 toward Westerlund 2 (25 pc in radius at 8 kpc) is not compatible with an origin of the emis- sion in the colliding winds of the WR 20a binary. The whole Westerlund 2 cluster is also too compact to ex- plain the large TeV source. A more diffuse cause, per- haps in the collective action of separate winds, is needed. This is also true for the more compact case of the TeV J2032+4130 source (4.4 pc in size at 1.5 kpc) which is seen toward a sub-group of 10 or more O stars of the Cyg OB2 cluster. The position of the EGRET EGR J2033+41 source (consistent with 3EG J2033+4118) is well centred on the dense core of the cluster and it could encompass emission from the entire system because of the wide EGRET point-spread- function. The massive eclipsing binary V729 Cyg in Cyg OB2 cannot account for the observed luminosity [33]. Interestingly, the γ-ray sources toward Westerlund 2 and Cyg OB2 are found where the hot bubble of the main HII region breaks free into the low density me- dium, blowing away the ambient gas in a champagne flow that sweeps the magnetic field lines away with it in a mushroom-like configuration. Adiabatic losses in the expanding flow must be severe, but particles can stream along the field lines and flow back to the edge of the cloud, or they can diffuse along the edge of the champagne flow to produce γ rays [34]. The problem is to accelerate cosmic rays from the bubbling inside the champagne bottle, within the HII bubble, to begin with. This γ-ray champagne bubble scenario, as well as the possible connection between young massive clusters and γ-ray sources has yet to be verified observationally in other examples, taking advantage of the upcoming surveys of the Galactic plane by GLAST and the Cher- enkov telescopes and combining their data to constrain the high-energy spectrum. Having missed until recently the importance of Cyg OB2 at a mere distance of 1.5 kpc reminds us that super star clusters may have es- caped radio searches in the Galaxy. Cosmic-ray acceleration in supernova remnants Collisionless shocks in supernova remnants are thought to produce the bulk of the cosmic rays up to the knee. They are numerous and powerful enough to sustain the total cosmic-ray power in the Galaxy. Synchrotron X rays also indicate that the forward shock effectively accelerates electrons to tens of TeV (see below). Yet, the accelerated ions are very elusive and they have not been firmly observed so far. Strong, but indirect, evi- dence of their presence has come from several impor- tant clues in the X-ray images. The thinness of the syn- chrotron filaments observed behind the forward shock of young remnants (such as Cas A, Kepler, Tycho, and SN 1006) imply very severe losses from large magnetic field strengths, well in excess of the compressed inter- stellar field ([35] and references therein). The electrons rapidly cool down and their radiation shifts to UV en- ergies, below the spectral window of the X-ray tele- scope. It has been proposed that the magnetic amplifi- cation is due to cosmic-ray streaming upstream [36, 37]. Another clue comes from the X-ray morphology of these remnants. All models of diffusive shock accelera- tion predict that ions receive far more energy than elec- trons. When this energy drain becomes significant, the gas hydrodynamics and shock profile are modified. The shocked gas becomes more compressible and it piles up in a much thinner shell outside the contact discontinu- ity. Instead of the classical compression ratio r = 4, average values of 6 ≤ r ≤ 8 are predicted after several centuries of acceleration activity. The post-shock tem- perature thus drops by an order of magnitude or more compared to the classical case [38]. Several observa- tions support this possibility. A modified shock can reconcile the unusually low electron temperature meas- ured in the young 1E0102.2-7219 remnant with its large shock velocity [39]. Detailed spectro-imaging of Tycho and Kepler also shows that the forward shock is twice closer to the contact discontinuity than one would expect when the shock is not disturbed by accelerated ions [40]. These are strong indications of efficient ac- celeration. The case of Cas A is less conclusive. A wide shell is seen between the shock and the ejecta. It may be due to inefficient acceleration, despite the large magnetic amplification implied by thin X-ray filaments, or to the remnant expansion in the wind of its progeni- tor [40]. Three remnants, G347.3-0.5, Vela Jr, and RCW 86, have now been imaged at TeV energies [41, 42, 43]. G347.3-0.5 and RCW 86 may be the remnants of the historical supernovae SN 393 and SN 185. The three objects share many common traits. Their large size (11 pc at 1.3 kpc, 17 pc at 1 kpc, and 18 pc at 2.8 kpc, re- spectively), their large ratio of synchrotron to thermal X-ray flux, together with their weak radio emission suggest that they have expanded at high speed in a low- density medium, possibly a stellar wind bubble or an OB cavity. So, they would be near the end of the free expansion phase or in early Sedov stage. Part of their shells has reached a denser environment a few centu- ries ago (to the west and southwest of G347.3-0.5 [44] and southwest of RCW 86 [45]). Their non-thermal X- ray filaments are quite broader than those of the younger remnants quoted above, thus implying a lower degree of magnetic amplification. The TeV flux appears 30TH INTERNATIONAL COSMIC RAY CONFERENCE 2007 well correlated with the X-ray synchrotron one. Syn- chrotron losses dominate over inverse Compton ones in their spectral energy distributions (SEDs), even if the TeV emission is entirely electronic. It is my purpose here to grossly constrain the range of shock velocities (vsh), post-shock magnetic fields (Bd), and diffusion coefficients at high energy that are consistent with the observed energies and flux ratios of the SED peaks, and with the observed width of the synchrotron and γ-ray profiles. It is also my purpose to convey that both pos- sible origins for the γ rays (pion-decay from in-situ ions or electrons up-scattering the cosmological mi- crowave background (CMB) radiation and dust infrared emission) still face serious difficulties. Following the formalism of Parizot et al. [35], I will assume that both the upstream and downstream diffu- sion coefficients D(E) scale as k0 times the Bohm limit, so D(E) = k0E/3eB for an electron or proton of energy E in a B field. The most efficient acceleration takes place when k0 = 1. I also assume a mean compression ratio r = 7 and a ratio of downstream to upstream mag- netic field strength Bd/Bu = 0.83 × r in the case of iso- tropic magnetic turbulence [46]. The synchrotron peak energy and filament width provides two constraints. Equating the 1st order Fermi acceleration timescale and the synchrotron loss timescale upstream and down- stream yields an estimate of the maximum electron energy Eemax ∝ Bd-1/2 k0-1/2 vsh. The synchrotron cut-off energy therefore scales as hνsyn-cut ∝ vsh2 k0-1 and pro- vides a first constraint between the shock velocity and k0. It is displayed in the left plots of Figure 4 for the three remnants. Because of diffusion, advection, and synchrotron losses, the high-energy electron distribu- tion falls off exponentially downstream and the syn- chrotron profile at a given X-ray frequency, hνX, has a projected full width at half maximum Δθsyn that fol- lows equation 24 in [35]. The ratio Bd3/2/vsh is thus constrained by the three observables hνsyn-cut, hνX, and Δθsyn. This relation is plotted as the blue line in the right plots of Figure 4. The cut-off energies being poorly constrained in the current SEDs, an uncertainty within a decade is quite plausible. It implies an equiva- lent uncertainty in k0. Its impact is shown in Figure 4 as the blue shaded area. For the input parameters of the three remnants, one finds that the downstream field roughly scales with hνsyn-cut-1/3.Δθsyn-2/3. This is why the constraint on Bd, displayed again as the blue shaded area, is more robust. The γ rays provide other constraints. The peak inverse Compton (IC) flux cannot exceed the peak γ-ray flux observed in the SED, so the magnetic energy density has a lower limit set by this flux ratio and the energy density of the ambient soft radiation field. This lower limit is met if one assumes the TeV data to arise from pure IC emission. The lower limit appears as the lower horizontal red line in Figure 4. I have used 0.26 MeV m-3 for the cosmological background and 0.27 MeV m-3 for the local density of of thermal emission from cold (24 K) dust for Vela Jr and G347.3-0.5 [47]. RCW 86 resides in the much brighter inner Galaxy, but in the quieter interarm region between the Carina and Scutum-Crux arms, so the ambient inrared density may be twice larger than the local value. If IC emission dominates over pion decay, the peak γ-ray energy, hνγcut, provides another relation between the down- stream field, hνsyn-cut, hνγcut, and the soft photon ener- gies (Esoft = 0.66 and 5.8 meV for the CMB and cold dust photons, respectively). It appears as the grey shaded area in Figure 4. In the same IC framework, the width of the TeV profile should not exceed the ob- served one. This provides another lower limit in the (Bd, vsh) phase space that depends on Esoft, hνsyn-cut and the intrinsic width ΔRγ of the γ-ray shell as inferred from the observations at the energy hνγ. This lower limit appears as the two broken red lines in Figure 4 (solid and dotted for the CMB and IR target photons, distance (kpc) angular radius projected FWHM of the X-ray filaments intrinsic fractional width dR/R of the shell at 1 TeV synchrotron peak energy (keV) synchrotron peak energy flux (eV cm-2 s-1) inverse Compton peak energy (TeV) inverse Compton peak energy flux (eV cm-2 s-1) G 347.3-0.5 1.3 30’ 40” at 2 keV 45 % ∼ 1 ∼ 190 ∼ 5 ∼ 19 Vela Jr 0.2 or 1.0 60’ 50” at 5 keV 22.5 % ∼ 0.9 ∼ 100 ∼ 4 ∼ 20 RCW 86 2.8 22’ 100” at 2 keV 50 % ∼ 0.15 ∼ 8 ∼ 1 ∼ 4 Table 2: supernova remnant characteristics used to constrain the downstream magnetic field and shock velocity. THE GALACTIC GAMMA-RAY CLUB respectively). The blue lines and broken red lines change by only 25% when using the classical r = 4 compression ratio instead of r = 7. The observational constraints summarized in Table 1 were taken from the data in [35, 41, 42, 43, 45, and references therein]. In the absence of velocity measurements, one expects the shock speed to vary with the remnant radius R and age τage as vsh = 2R/3τage during the free expansion phase. The velocity can be derived for G347.3-0.5 and RCW 86 if they are related to the historical supernovae. It is plotted as the vertical dashed line in Figure 4. On the other hand, we can set a maximum velocity by rea- sonably requiring the age of the three objects to be older than 1000 years. The upper limit is plotted as the solid vertical line in Figure 4. The results for G347.3-0.5, Vela Jr at 1 kpc, and RCW 86 show that their current shock speed should exceed 3800, 3600, and 1500 km/s, respectively, to accelerate electrons to the observed cut-off energies in X rays with the maximum efficiency (k0 = 1, Bohm limit). Large field strengths, of order 10 nT, are also required to explain the thinness of the X-ray synchrotron fila- ments. These values are less subject to the uncertainty in the synchrotron cut-off energy. As expected, they are slightly lower than the fields strengths of 20 to 30 nT derived for the younger historical remnants [35]. Let us first examine the case of dominant IC emission at high energy. The large magnetic fields required by the X-ray filaments are consistent with the peak energy observed by HESS. The energetic electrons preferably up-scatter soft, CMB-like, photons which correspond to the lower end of the grey shaded region. The large magnetic fields are also consistent with the thickness of the TeV shell. IC emission also provides a natural ex- planation for the tight spatial correlation between the X-ray and TeV fluxes. Yet, the necessary large mag- netic fields are not consistent with a dominant IC origin of the γ-ray flux. A ‘pure’ electronic origin requires the low fields of 1.5, 1.0, and 0.8 nT inferred from the peak-flux ratio between the X-ray and γ-ray compo- nents for the 3 remnants. Similar values have been found for G347.3-0.5 and Vela Jr by modelling the observed SEDs with a power-law distribution of elec- trons [41, 42]. These fields are, however, too low to explain the sharp synchrotron filaments and are only marginally consistent with the TeV shell thickness. One would need to increase the sub-mm part of the dust emission by two orders of magnitude to reconcile the dominant IC scenario and high magnetic fields. This option is unlikely. Another possibility is to suppose that the large magnetic fields are confined to sharp filaments at the shock and that they rapidly drop inward [48]. If so, the filaments would trace the magnetic radial profile rather than fast synchrotron cooling and the field Figure 4: constraints on the diffusion coefficient (k0 times the Bohm limit) and on the downstream mag- netic field as a function of shock velocity for the three supernova remnants, G347.3-0.5, Vela Jr at 1 kpc, and RCW 86, resolved at TeV energies (see text for the explanation of the different lines and regions). 30TH INTERNATIONAL COSMIC RAY CONFERENCE 2007 strength Bd would not be constrained by the blue areas in Figure 4. This option requires a careful modelling of the diffusion, advection, and radiation cooling to be tested against the observations because of the impor- tance of the projection effects. Electron spectral hard- ening with energy, as expected from modified shocks, should also be taken into account. A convincing IC scenario would also need to explain why the synchro- tron spectral index varies from 1.8 to 2.6 along the shell of G347.3-0.5, whereas the γ-ray index remains uniformly close to 2.0 [41, 44]. The large magnetic field helps in this matter since it implies X-ray elec- trons that have a few times more energy than the γ-ray emitting ones. So, the latter are closer to the cut-off energy (thus harder) than the X-ray ones that belong to the exponential cut-off part of the spectrum. The nearby distance of 200 pc for Vela Jr does not yield a convincing case. The minimum shock velocity neces- sary to sustain the X-ray synchrotron properties is 1.6 times larger than the maximum value set by an age older than 1 kyr and the required field is very large (Bd > 25 nT). It would completely rule out an IC scenario, as noted by the HESS group [42]. Following the constraints in Figure 4 and adopting downstream field strengths of 8, 8, and 6 nT as possible values for G347.3-0.5, Vela Jr (1 kpc), and RCW 86, respectively, we can infer maximum electron energies of 26, 24, and 11 TeV and synchrotron lifetimes of 100, 110, and 380 yr. The maximum proton energy is de- rived by equating the acceleration timescale and the remnant age. Choosing ages of 1620 and 1820 years from the historical supernovae and a comparable age of 2300 years for Vela Jr at 1 kpc that corresponds to a current speed of 5000 km/s in Figure 4, we find that protons can be accelerated to 400 TeV in G347.3-0.5, 500 TeV in Vela Jr, and 50 TeV in RCW 86. These values fall well below the knee energy despite the large magnetic amplification. The region explored by these high-energy protons in their random walk is found to be smaller than the observed TeV shell thickness. The large magnetic amplification favours the interpre- tation of the γ-ray flux in terms of pion decay, but this scenario also faces important difficulties. Large gas densities between 1 and 2 cm-3 are required to match the TeV flux. Yet, the absence of thermal emission in G347.3-0.5 sets an upper limit of 0.02 cm-3 at 1 kpc [44]. This limit can be further lowered for an increased gas compressibility in a modified shock. A low density of 0.03 ff-1/2 cm-3 (for a filling factor ff) has also been found for Vela Jr at 1 kpc [49]. These values cannot be increased to meet the γ-ray requirement. Large densi- ties (300 cm-3) exist on the southwestern side of G347.3-0.5 where the clear correlation between the synchrotron flux and absorbing column-density indi- cates that the shock runs into dense clouds seen in CO. If protons are accelerated as well as electrons, having more high-energy protons and more targets to produce pions, one would expect the pion-decay flux to scale more or less quadratically with the ambient density. Yet, the γ-ray profile linearly follows the X-ray one, both radially and azimuthally. It does not exhibit an enhanced γ-ray to X-ray flux ratio toward the CO clouds [41]. The same difficulty is seen in RCW 86. The TeV map presented at the conference carefully avoids the southwestern region where the shock is known to have slowed down to 800 km/s inside a dense cloud and where ample thermal emission is detected [45]. In addition, the pion-decay modelling of the TeV emission in Vela Jr requires an unsually low electron to proton ratio (< 10-4). Furthermore, the maximum pro- ton energies derived from the amplified magnetic fields chosen above correspond to cut-off energies in the γ- ray spectrum of 70 TeV for G347.3-0.5 and 90 TeV for Vela Jr, at odds with the spectral break detected be- tween 3.7 ± 1.0 TeV and 17.9 ± 3.3 TeV in G347.3-0.5 [41] and possibly several TeV for Vela Jr [42]. One would need to decrease the downstream field by an order of magnitude to explain these breaks from proton-proton collisions. It would then conflict with the observed synchrotron properties and IC emission would become dominant. The spectral break near 9 TeV inferred for RCW 86 would be consistent with the possible 5 TeV break reported at the conference. Fi- nally, it has often been said that a flat spectrum below 1 TeV would sign pion-decay against IC emission, but the reader should be warned that the flattening at high energy of the electron spectra emerging from modified shocks, as well as various dosages of infrared photon targets can easily broaden the IC spectra and increase the overall IC flux for a given magnetic amplitude. Flattened spectra will certainly help to better fit the radio and X-ray data simultaneously, but they will pro- vide less room for pion-decay emission. In conclusion, whereas indirect evidence of ion accel- eration exists in X rays from the modified shock ther- modynamics, no clear picture has emerged yet from the beautiful γ-ray images of shell supernova remnants. Pulsar wind nebulae The case of the HESS J1813-178 source brings forth the problem of separating emission from the supernova remnant and pulsar activities when the telescope reso- lution is limited. The source coincides with the radio THE GALACTIC GAMMA-RAY CLUB shell of the G12.82-0.02 remnant and with a non- thermal X-ray nebula that probably traces the wind of a young and energetic pulsar inside the remnant, al- though searches for radio pulsations have failed so far [50]. Since the 1990s, when the GeV to TeV emission from the Crab nebula was successfully interpreted as synchro-self-Compton emission in the pulsar wind downstream of its terminal shock inside the remnant [51], and when several of us drew attention to the fact that a number of hard X-ray nebulae were seen inside the error boxes of unidentified, possibly variable, EGRET sources (e. g. [52, 53]), pulsar wind nebulae have been firmly established as sources of TeV emis- sion. The identification is based on the correlation be- tween the X-ray and γ-ray images of the diffuse emis- sion. In several cases, the TeV image extends further away from the pulsar, as around PSR B1823-13 (HESS J1825-137, [54]), in the long Vela X tail [55], or along the jet of PSR B1509-58 [56]. For more compact sources, unresolved by HESS, such as in G21.5-0.9, G0.9+0.1, and Kes 75, the proposed identification is based on the lack of non-thermal X-ray emission from the shell or on the weakness of the magnetic field re- quired to explain the TeV flux from electrons acceler- ated at the shell [57, 58]. The number of pulsar wind nebulae emitting synchro- tron X rays has grown rapidly in the past few years. Thanks to the high-resolution images provided by XMM-Newton and Chandra, many have been resolved into a large variety of shapes. The innermost regions of the wind are dominated by polar jets and a toroidal wind that results from the winding of the neutron star magnetic field. A rather complex standing shock forms as the wind decelerates to match the boundary condi- tion imposed by the external medium (the pressure of the supernova ejecta or the ram pressure of the ambient medium if the pulsar motion has become supersonic inside or outside the remnant). The downstream equa- torial flow expands laterally and turns back against the external medium to fill intermediate latitudes and form an elongated bubble, stretched along the pulsar spin axis and jets [59]. The synchrotron emitting part of the wind is confined between the termination shock and the bubble outer boundary since it is assumed that the up- stream charges flow along with the frozen-in field and do not radiate. Doppler boosting of the synchrotron intensity is important for the innermost part of the wind, as in the Crab wisps, but not for the rest of the nebula [59]. The wind electrons can efficiently upscat- ter soft photons to produce TeV γ rays. In the Crab nebula, the average synchrotron infrared and optical energy density in the bubble exceeds the other ambient radiation fields and the modelling of the multi- wavelength spectrum indeed finds that synchro-self- Compton emission dominates in γ rays [51, 60]. The situation appears to be different for the other nebulae seen in γ rays. The second most luminous one after the Crab is in G21.5-0.9. Assuming that the sub-eV syn- chrotron energy flux is comparable to the observed X- ray one (for a flat SED), and considering the size of the synchrotron bubble (0.65’, [61]), one finds an average energy density 7 times lower than that of the cosmo- logical background or of the dust radiation at the solar circle. The latter can easily increase by a factor of a few in the inner Galaxy, so synchro-self-Compton should not dominate in G21.5-0.9. The third most luminous case is found in the left wing of the Kookaburra, around PSR J1420-6048. It is fainter, but more com- pact. Following the same assumptions, the soft syn- chrotron energy density falls 25 times below the inter- stellar or cosmological fields, so we can consider that most of the TeV wind nebulae should mainly upscatter the cosmological background and interstellar radiation. Starlight and warm dust radiation may play a signifi- cant role below 1 TeV if the nebula resides near bright star clusters as in G0.9+0.1 [57] and G12.82-0.02 (HESS J1813-178, [50]). In any case, the GLAST sen- sitivity should prove very useful to constrain the peak of the inverse Compton emission, therefore the particle ageing. Thanks to the X-ray imaging capabilities, one can often separate the diffuse nebular emission from pulsar DC Figure 5: evolution of the 2-10 keV luminosity of pul- sar wind nebulae as a function of the spindown power. The squares mark the observed values, highlighted in blue when the nebula also shines at TeV energies, and the open diamonds give the toy model predictions if 1% of the spindown power is given to the wind. 30TH INTERNATIONAL COSMIC RAY CONFERENCE 2007 Figure 6: evolution of the ratio of the 0.5-10 TeV to the 2-10 keV luminosities of pulsar wind nebulae. The filled squares give the observed values and the open diamonds the prediction from the toy model. The three blue squares with a ratio < 0.1 are the youngest, most compact, wind nebulae in the sample (Crab, Kes 75, and G21.5-0.9). The fourth blue square marks the jet- like nebula from PSR B1509-58. Figure 7: distributions of the photon spectral indices of the pulsar wind nebulae in X rays and at TeV energies, for the whole sample of objects displayed in Figure 5 (upper histogram) and the sub-group of TeV emitting objects (right histogram and scatter plot). emission or the innermost part of the wind. Several authors have noted that the nebular luminosity scales with the pulsar spindown power as LX ∝ Ė psr1.4 ± 0.2 [62, and references therein]. Updating their list for the newly found nebulae and for results found in the litera- ture on more detailed morphological studies of several Figure 8: evolution of the size ratio between the TeV and X-ray emission regions of pulsar wind nebulae as a function of spindown power. nebulae they had listed (details will be given in a forth- coming paper), I have plotted in Figure 5 the evolution of the 2-10 keV luminosity with Ėpsr and I have high- lighted the sources that have been detected at TeV en- ergies. The latter span a small range of ages from a thousand years in the Crab and Kes 75 to 50 kyr in PSR J1809-1917. Their X-ray luminosity is not particularly bright compared to the general scaling relation. The luminosities recorded between 0.5 and 10 TeV show a much narrower dynamical range. Except for Vela X and G21.5-0.9, which have luminosities of 8 1025 W and 6 1026 W, respectively, all other sources gather in the (0.3-1.4) 1028 W range and show little dependence on the pulsar spindown power, if any, and a large disper- sion. This is illustrated in Figure 6 where the ratio of the TeV to X-ray luminosity scales as LTeV/LX ∝ Ė psr-1.9. In fact, most of the apparent evolution is due to the change in X-ray emission. This trend is nicely inde- pendent of distance. Since the TeV emitting electrons have lower energies than the X-ray ones, the evolution of the LTeV/LX ratio with Ėpsr (alias youth) traces how the particles age in the nebula. Particle ageing is also illustrated by the spectral softening of the TeV emitting electrons with respect to the X-ray emitting ones. Figure 7 shows that the γ-ray spectra are often softer than the X-ray ones. Particle ageing is also reflected in the larger extent of the TeV nebula when the source is resolved in both energy bands. Figure 8 shows that the size ratio, meas- ured in terms of radius or length depending on the source morphology, indeed increases with age (i. e. decreasing spindown power) as RTeV/RX ∝ Ė psr-2.0, but it presents a large dispersion. The spectral softening with distance from the pulsar strongly supports an electronic origin of the γ radiation since synchrotron and inverse THE GALACTIC GAMMA-RAY CLUB Compton losses result in a longer lifetime at lower energy. One would expect harder spectra in γ rays than in X rays at large distance in the case of ion emission. The higher energy electrons illuminating the X-ray synchrotron nebula give a more instantaneous image of the wind than the longer-lived electrons shining in γ rays which integrate a fair fraction of the pulsar wind history. Both the evolution of the wind power and of its magnetic field near the shock are linked to the history of the pulsar spindown power. To explore their impact on the recorded luminosities in Figures 5 and 6, I have developed a toy model based on a small set of assump- tions. Given the young age (< 600 kyr) of the X-ray nebulae plotted in Figure 5, one can reasonably assume that the neutron star magnetic field has not decayed, therefore that the product P(n-2).(dP/dt) of the rotational period and its time derivative remains remains constant from birth to now. n notes the pulsar braking index (n = 3 for a magnetic dipole). This implies that the spin- down power evolves as Ėpsr(t) = Ė0 (1 + t/τ0)(n+1)/(n-1). The characteristic timescale τ0 can be retrieved from the present timing measurements and the pulsar period at birth: τ0 = P0(n-1) P(2-n) / (n-1) / dP/dt. A birth period P0 = 15 ms has been adopted to account for the fastest pulsars in the sample. In the absence of pulsar timing data for the HESS J1813-178, Rabbit, and HESS J1640-465 nebulae, a spindown power of 1030 W and ages of 2, 16, and 20 kyr have been chosen respec- tively, following [50, 63, 64]. The pulsar powering the bow shock nebula G189.22+2.90 in IC 443 is probably older (30 kyr) and less energetic (1029 W) [65]. A con- stant fraction of the spindown power Ėpsr(t) is poured into the wind at the terminal shock. It is distributed over the electron spectrum. A constant E-2.1 spectrum has been assumed for the injected electrons above 1 GeV. I have not attempted to reproduce the trend be- tween the X-ray photon index and the spindown power (αX ∝ Ė psr-1/2) that has been observed [66]. The maxi- mum electron energy has been set to the full electrical potential drop eΦopen across the open magnetosphere today. The latter evolves slowly as Ėpsr1/2 and has not changed much over the short lifetime of the X-ray emitting electrons, so it has been kept constant in the calculation. The particles have been traced in time, loosing energy primarily by synchrotron radiation. This should be revised in view of the large LTeV/LX ratios found for middle-aged objects. Adding inverse Comp- ton losses will remove even more particles from the γ- ray window. The magnetic energy density of the wind injected at the shock may scale as Ėpsr(t), so B(t) ∝ Ėpsr(t)1/2. Other scaling laws have been explored (B ∝ Ėpsr, ...), but do not appreciably modify the results given in Figure 5 and 6. The minimum magnetic strength in the wind at birth is set to the maximum value between 30 nT and the value required so that the electrons cur- rently injected with half the maximum energy eΦopen shine at 10 keV which is the upper limit of the X-ray window. This applies only to the older pulsars that have a reduced voltage. The inverse Compton luminosity was calculated for an energy density of 0.53 MeV m-3 of photons with an energy of 2 meV representative of the cosmological background and cold dust radiation. The most drastic assumption is that the injected parti- cles and field remain frozen. In other words, there is no spatial evolution of the wind density or field strength in the toy model and the particles loose energy in the same initial field all their life. This is why the compari- son between the predictions of the toy model and the data allows to guess what is due to the pulsar evolution and what must be due to the wind expansion and struc- tural evolution in the data. The results are presented for each nebula as open dia- monds in Figures 5 and 6. The evolution of the pre- dicted synchrotron luminosity in the 2-10 keV band follows reasonably well the observations given the simplicity of the toy model. The wind power has been set to 1% of the spindown power as a good match to the data. Most of the observed X-ray evolution is driven by the decrease in wind power with time be- cause the short lifetime of the X-ray emitting electrons. They have been injected in the recent past, typically over the last 10% or 20% of the pulsar lifetime in most cases. The situation is quite different in γ rays since the emission integrates the wind history over a large frac- tion of the pulsar age, often more than half its age. The evolution of the predicted inverse Compton emission is much shallower than the data. The modelled LTeV/LX ratio is independent of the power fraction attributed to the wind. The predicted slope in Figure 6 is rather in- sensitive to the initial strength as well as the evolution of the magnetic field B(t). The rapid increase of the TeV to X-ray luminosity ratio with age indicates that synchrotron losses must strongly decrease as the parti- cles move out to keep enough of them shining in γ rays. The four nebulae in the sample for which we ex- pect a minimal influence of the spatial evolution have been highlighted. They consist of the three very young and compact Crab, Kes 75, and G21.5-0.9 nebulae where most of the synchrotron power comes from the inner region where the magnetic pressure has built up to the equipartition value because of the wind slowing down [67]. The jet-like emission from PSR B1509-58 has also been highlighted because the comparison of the tail lengths recorded from 0.5 to 100 keV by RO- SAT, Beppo-SAX, and INTEGRAL was compatible with synchrotron ageing in a uniform magnetic field 30TH INTERNATIONAL COSMIC RAY CONFERENCE 2007 over 20 or 30 pc [56]. The toy model predictions are indeed closest to the data for these cases, especially when keeping in mind that an additional SSC compo- nent for the Crab would move the model prediction up by a factor of 3 or 4. For the others objects, the syn- chrotron burn-off is obviously too strong to account for the TeV flux. Adding inverse Compton losses will in- crease the discrepancy at old age. The model strongly under-predicts the TeV flux and it is unlikely that the ambient interstellar radiation field can be increased to match the data. So, the trend observed in Figure 6 bears valuable information on the spatial structure of the nebulae as they evolve. The fact that the Cherenkov telescopes have detected wind nebulae at different stages of their evolution is of great interest. The three youngest objects quoted above illustrate the early development of the wind when con- fined by the supernova ejecta. HESS J1640-465 may bring another example when its age is known. The elongated nebular shapes in HESS J1813-178 and around PSR B1823-13, well inside their supernova remnant, and the long Vela X tail that expands almost at right angle from the pulsar spin axis give potential examples of winds crushed back by an irregular reverse shock. This may happen when the forward supernova shock slows down at different rates in azimuth because of different mass loading in a non-uniform medium [68]. Later in the evolution, when the pulsar has moved near the edge of the supernova remnant or when it has left it, its supersonic motion confines the wind inside a bow shock. The ram pressure strongly compresses the wind upfront and lets it stretch at the back. It forms a cometary tail trailing behind the pulsar. The particles injected at the front and back regions of the termination shock suffer very different losses before joining in the tail. Modelling the spatial distribution of the particle density and magnetic field strength in the crushed and bow shock configurations is difficult and being able to probe particle ageing in these situations with GeV to TeV γ rays will prove very useful to constrain the mod- els. The nebulae of PSR J1809-1917 and B1800-21 may illustrate this stage when the identification is con- firmed. So will the Rabbit wind when its age is known. Understanding why the X-ray wakes of the relatively young pulsars in IC 443 and in CTB 80 (PSR B1951+32) have not been detected above 100 GeV by MAGIC also needs further investigations [69, 70], as does the case of PSR B1853+01 in W44. Other promising candidates have been reported at GeV energies from EGRET sources and need confirmation by GLAST. The 10-kyr old, 2.2 1030 W pulsar PSR J2229+6114 is a compelling identification for the stable 3EG J2227+6122 source that is confirmed in the re- vised catalogue [71]. It also coincides with a COMP- TEL source in the 0.75-3 MeV band. The compact X- ray nebula, with a possible 14” jet, belongs to an in- complete non-thermal radio shell. CTA 1 should also bring an interesting case. The brightest unidentified EGRET source off the Galactic plane, 3EG J1835+5918, and its X-ray counterpart have long been proposed as a second Geminga [72, 73] because of its hard and stable spectrum cutting off at 2 GeV and be- cause of the lack of radio and optical counterparts down to very low magnitudes. The 1.7 1029 W compact keV nebula can be powered by a 20 kyr radio-quiet pulsar. Another promising case corresponds to the wind of PSR B1046-58. Likely γ-ray pulsations have been found in the signal from 3EGJ1048-5840, but 40 % of the emission above 400 MeV is not pulsed and a faint X-ray nebula has recently been found for this 20-kyr old, 2 1029 W pulsar [74, 75]. The fraction of radio- quiet Geminga pulsars among the Galactic sources may be rather large if the pulsed γ-ray beams are produced at high altitude inside the magnetosphere [76]. Search- ing for wind nebulae with GLAST will therefore need to concentrate at energies above several GeV to benefit from the sharp cut-off expected in the pulsed emission from 10-100 kyr old neutron stars. Gamma-ray binaries The detection of a pulsar wind nebula in a binary sys- tem has opened the possibility of probing the wind structure as a function of compression near periastron [77]. PSR B1259-63 indeed follows an eccentric 3.5-yr long orbit around its massive Be companion and the ram and radiation pressures build up when it crosses the equatorial outflow from the star. A variety of situa- tions have been explored at the interface between the two winds: whether the pulsar wind is still supersonic and bounded by a termination shock at the point of pressure balance with the stellar outflow or not, or whether the Compton drag from the intense stellar radiation can slow down the unshocked wind or not. Inverse Compton components and lightcurves have been estimated [78, and references therein]. Another source of γ rays has been proposed using the termina- tion shock of the stellar outflow against the pulsar wind to accelerate ions and electrons and letting them shine in γ rays by π0 decay, bremsstrahlung, and inverse Compton processes [79]. The TeV detection of two other eccentric and massive binaries, LS 5039 [80] and LSI +61°303 [81], has started a debate between a pul- sar wind or a microquasar jet origin of the emission. The nature of the compact object in LS 5039 and LSI +61°303 was not firmly identified and both types of THE GALACTIC GAMMA-RAY CLUB systems could present striking similarities when seen from Earth. Because of the stellar brightness, a strong orbital modulation is expected in both systems from two-photon pair production [26]. It is essential to model this absorption to be able to explore the intrinsic source variability that may result from a change in accretion rate in the microquasar case or a change in wind compression in the pulsar case, and from the in- crease in inverse Compton emissivity near periastron in both systems. The energy distribution recorded for the three binaries were quite alike and, morphologically, comparable images could be computed from an ex- tended jet and from a cometary wind [82]. Very high resolution radio images of LSI +61°303 showed a tail pointing away from the star along the orbit [83] which strongly suggested the presence of a bow-shock pulsar wind. So, the situation appeared confused until the detection, at the 4σ level, of the canonical black-hole binary Cygnus X-1. It was seen by MAGIC above 100 GeV for a short hour [84]. It has strengthened the case for a genuine high-energy activity from microquasars. Confirming another γ-ray flare will establish high-mass microquasars as γ-ray emitters as well as pulsar wind nebulae. Summarizing the numerous models that pre- dict γ radiation from microquasar jets and how they should vary with aspect angle, precession, accretion rate, and companion type is beyond the scope of this review. The studies of LS 5039 [85, 86] illustrate the various processes at work in massive systems where inverse Compton scattering of the copious stellar pho- tons clearly dominates in the GeV to TeV band. This is probably why only massive binary systems have been detected so far. Above 1 TeV, synchro-self-Compton emission contributes a significant fraction to the overall flux. This process is the only hope of detecting low- mass binaries with current telescopes [87], unless the intensity is magnified by Doppler boosting at small viewing angles from the jet axis. We would then hap- pily discover the long-sought microblazars. In conclusion, with both rotation-powered and accre- tion powered massive binaries, isolated pulsars and the many facets of their magnetospheric and wind activi- ties, champagne bubbles from superstar clusters and SNOBs, shock acceleration in supernova remnants, bubbles and stellar wind collisions, a wealth of GeV and TeV sources still awaiting identification, and higher performance telescopes soon to come, the future of the Galactic γ-ray club looks very bright. Let me thank the organizers for having run such a festive and lively con- ference where many ideas were forged and debated to make this future even brighter. References [1] Hartman, R. C., et al., 1999, ApJS, 123, 79. [2] Casandjian, J. M., & Grenier, I. A., these pro- ceedings. [3] Grenier, I. A., & Casandjian, J. M., Terrier, R. 2005, Science, 307, 1292. [4] Grenier, I. A., 1995, AdSpR, 15, 73. [5] Grenier, I. A., 2000, A&A, 364, L93. [6] Gehrels, N., et al., 2000, Nature, 404, 363. [7] http://www-glast.slac.stanford.edu/software/ IS/glast_lat_performance.htm [8] Aharonian, F. A., et al., 2008, A&A, in press, arXiv:astro-ph/0801.3555v3 [9] Strong, A. W., & Mattox, J. R., 1996, 308, L21. [10] Aharonian, F. A., & Atoyan, A. M., 1996, A&A, 309, 917. [11] Gabici, S., & Aharonian, F. A., 2007, ApJ, 665, L131 [12] Aharonian, F. A., et al., 2006, Nature, 439, 695. [13] Crocker, R. M., et al., 2007, ApJ, 666, 934 [14] Sakano, M., Warwick, R. S., Decourchelle, A., 2006, JPhCS, 54, 133. [15] Aharonian, F. A., et al., 2006, A&A, 467, 1075. [16] Abdo, A. A., et al., 2007, ApJ, 658, L33. [17] 2008, al., et J.,   Albert, arXiv:astro-ph/0801.2391v2 [18] Montmerle, T., 1979, ApJ, 231, 95. [19] Binns, W. R., et al., 2007, SSRv, 130, 439. [20] Binns, W. R., et al. 2005, ApJ, 634, 351. [21] Figer, F. D., McLean, I. S., Morris, M., 1999, ApJ, 514, 202. [22] Figer, F. D., et al., 2002, ApJ, 581, 258. [23] Knödlseder, J., et al., 2000, A&A, 360, 539. [24] Churchwell, E., Whitney, B. A., Babler, B. L., et al. 2004, ApJS, 154, 322 [25] Butt, Y. M., et al., 2003, ApJ, 597, 494. [26] Dubus, G., 2006, A&A, 451, 9. [27] Orlando, E., Strong, A. W., 2007, these pro- ceedings. [28] Hanson, M. M., 2003, ApJ, 597, 957. [29] Piatti, A. E., Bica, E., & Claria, J. J. 1998, A&AS, 127, 423 [30] Cesarsky, C. J., Montmerle, T., 1983, SSRv, 36, 173. 30TH INTERNATIONAL COSMIC RAY CONFERENCE 2007 [31] Bykov, A. M., Toptygin, I. N., 2001, AstL, 27, 625. [32] Reimer, A., Pohl, M., Reimer, O., 2006, ApJ, 644, 1118 [33] Benaglia, P., Romero, G. E., Stevens, I. R., Torres, D. F., 2001, A&A, 366, 605. [34] Völk, H. J., 1983, SSRv, 36, 3. [35] Parizot, E., Marcowith, A., Ballet, J., & Gal- lant, Y. A., 2006, A&A, 453, 387. [36] Bell, A. R., & Lucek, S. G., 2001, MNRAS, 321, 433. [37] Ptuskin, V. S., & Zirakashvili, V. N., 2003, A&A, 403, 1. [38] Ellison, D. C., Decourchelle, A., Ballet, J., 2004, A&A 413, 189. [39] Hughes, J. P., Rakowski, C. E., De- courchelle, A., 2000, ApJ, 543, L61. [40] Decourchelle, A, 2005, Proc. of the X-Ray and Radio Connections, Santa Fe, USA, 2004, http://www.aoc.nrao.edu/events/xraydio [41] Aharonian, F. A., et al., 2006, A&A, 449, 223 [42] Aharonian, F. A., et al., 2007, ApJ, 661, 236. [43] Hoppe et al., 2007, these proceedings. [44] Cassam-Chenaï, G., et al., 2004, A&A, 427, 199. [45] Vink, J., et al., 2006, ApJ, 648, L33. [46] Berezhko, E. G., Ksenofontov, L. T., Völk, H. J., 2002, A&A, 395, 943. [47] Moskalenko, I. V., Porter, T. A., Strong, A. W., 2006, ApJ 640, L155. [48] Pohl, M., Yan, H., & Lazarian, A, 2005, ApJ, 626, L101. [49] Slane, P., et al., 2001, ApJ, 548, 814. [50] Helfand, D. J., et al., 2007, ApJ, 665, 1297. [51] de Jager, O. C., et al., 1996, ApJ, 457, 253. [52] Roberts, M. S. E., Romani, R. W., Johnston, S., Green, A. J, 1999, ApJ, 515, 712. [53] Grenier, I. A., 1999, ICRC, 3, 476. [54] Aharonian, F. A., et al., 2006, A&A, 460, 365 [55] Aharonian, F. A., et al., 2006, A&A, 448, L47. [56] Forot, M., et al., 2006, ApJ, 651, L45. [57] Aharonian, F. A., et al., 2006, A&A, 432, L25. [58] Djannati-Ataï, A., et al., 2007, these proceed- ings. V. , 2 0 0 6 , [59] Komissarov, S. S., Lyubarsky, Y. E., 2004, MNRAS, 349, 779. [60] Aharonian, F. A., et al., 2004, ApJ, 614, 897. [61] Camilo, F., et al., 2006, ApJ, 637, 456. [62] Li, X.-H., Lu, F.-J., Li, Z., 2007, arXiv:astro-ph/0707.4279v1 [63] Ng, C.-Y., Roberts, M. S. E., & Romani, R. W, 2005, ApJ, 627, 904. [64] Funk, S., et al., 2007, ApJ, 662, 517. [65] Gaensler, B. M., et al., 2006, ApJ, 648, 1037. [ 6 6 ] E . G o t t h e l f , arXiv:astro-ph/0610376v1. [67] Kennel, C. F., Coroniti, F. V., 1984, ApJ, 283, 694. [68] Blondin, J.~M., Chevalier, R. A., Frierson, D. M, 2001, ApJ, 563, 806. [69] Albert, J., et al., 2007, ApJ, 664, L87. [70] Albert, J., et al., 2008, ApJ, 669, 1143. [71] Halpern, J. P., et al., 2001, ApJ, 552, L125. [72] Mirabal, N., & Halpern, J.~P, 2001, ApJ, 547, L137. [73] Reimer, O., et al., 2001, MNRAS, 324, 772. [74] Kaspi, V.~M., et al., 2000, ApJ, 528, 445. [75] Gonzales, M., et al., 2006, arXiv:astro-ph/0610523v2. [76] Harding, A. K., Grenier, I. A., & Gonthier, P. L, 2007, Ap&SS, 309, 221. [77] Aharonian, F. A., et al., 2005, A&A, 442, 1. [78] Ball, L., & Dodd, J, 2001, PASA, 18, 98. [79] Kawachi, A., et al., 2004, ApJ, 607, 949. [80] Aharonian, F. A., et al., 2005, Science, 309, 746. [81] Albert, J., et al., 2006, Science, 312, 1771. [82] Dubus, G., 2006, A&A, 456, 801. [83] Dhawan, V., Mioduszewski, A., Rupen, M, 2006, VI Microquasar Workshop: Microquasars and Beyond. [84] Albert, J., et al., 2007, ApJ, 665, L51. [85] Paredes, J. M., Bosch-Ramon, V., Romero, G. E, 2006, A&A, 451, 259. [86] Dermer, C.~D., Böttcher, M, 2006, ApJ, 643, 1081. [87] Grenier, I. A., Kaufman Bernadò, M. M., & Romero, G. E, 2005, Ap&SS, 297, 109.
astro-ph/0411548
1
0411
2004-11-18T19:01:35
Intrinsically faint quasars: evidence for meV axion dark matter in the Universe
[ "astro-ph", "hep-ph" ]
Growing amount of observations indicate presence of intrinsically faint quasar subgroup (a few % of known quasars) with noncosmological quantized redshift. Here we find an analytical solution of Einstein equations describing bubbles made from axions with periodic interaction potential. Such particles are currently considered as one of the leading dark matter candidate. The bubble interior possesses equal gravitational redshift which can have any value between zero and infinity. Quantum pressure supports the bubble against collapse and yields states stable on the scale more then hundreds million years. Our results explain the observed quantization of quasar redshift and suggest that intrinsically faint point-like quasars associated with nearby galaxies are axionic bubbles with masses 10^8-10^9M_{Sun} and radii 10^3-10^4R_{Sun}. They are born in active galaxies and ejected into surrounding space. Properties of such quasars unambiguously indicate presence of axion dark matter in the Universe and yield the axion mass \approx 1 meV, which fits in the open axion mass window constrained by astrophysical and cosmological arguments.
astro-ph
astro-ph
Intrinsically faint quasars: evidence for meV axion dark matter in the Universe Anatoly A. Svidzinsky Department of Physics, Institute for Quantum Studies, Texas A&M University, TX 77843-4242 (Dated: January 25, 2019) Growing amount of observations indicate presence of intrinsically faint quasar subgroup (a few % of known quasars) with noncosmological quantized redshift. Here we find an analytical solution of Einstein equations describing bubbles made from axions with periodic interaction potential. Such particles are currently considered as one of the leading dark matter candidate. The bubble interior possesses equal gravitational redshift which can have any value between zero and infinity. Quantum pressure supports the bubble against collapse and yields states stable on the scale more then hun- dreds million years. Our results explain the observed quantization of quasar redshift and suggest that intrinsically faint point-like quasars associated with nearby galaxies are axionic bubbles with masses 108-109M⊙ and radii 103-104R⊙. They are born in active galaxies and ejected into sur- rounding space. Properties of such quasars unambiguously indicate presence of axion dark matter in the Universe and yield the axion mass m ≈ 1 meV, which fits in the open axion mass window constrained by astrophysical and cosmological arguments. ∼ 50−100 Based on observations, Karlsson [1] has noted division of quasars (QSOs) into two groups with different red- shift properties and concluded the following. If we select QSOs associated with most nearby (distance d < Mpc), galaxies then their redshift is close to certain val- ues (quantized), as shown in Fig. 3 below. Meanwhile, in QSO samples associated with distant galaxies no pe- riodicity in intrinsic redshift is observed. Such a division is supported by later studies of QSOs associated with most nearby galaxies where the quantization was con- firmed [2, 3] and distant (0.01 < zgal < 0.3) galaxies for which absence of any periodicity was claimed [4]. The observations suggest existence of intrinsically faint (op- tical luminosity L = 105 − 107L⊙) QSO subgroup with quantized noncosmological redshift. Being intrinsically faint, such objects are not detected from large distances (which yields disappearance of redshift quantization in distant QSO samples) and constitute only a few % of the known QSO population. Observations indicate that such quasars are ejected from nearby active galaxies or in the process of ejection from the galactic nucleus [5, 6]. Here we show that bubbles of dark matter with pe- riodic interaction potential, masses about 108 − 109M⊙ and radii 103 − 104R⊙ can explain the intrinsically faint quasars. The bubble is supported against collapse by quantum pressure and decays on a time scale more than hundreds million years. Hypothetical axions, one of the leading dark matter candidate, fit well into this picture and can account for the redshift quantization. Usual baryonic matter falls into the bubble interior, heated by the release of the gravitational energy and produce electromagnetic radiation that freely propagate into sur- rounding space. In this Letter we study massive real scalar field ϕ with periodic interaction potential V (ϕ) = V0[1 − cos(ϕ/f )], (1) where V0 > 0. This potential is quite general and de- rived in quantum filed theory in connection with pseudo Nambu-Goldstone bosons (PNGBs) [7]. In all such mod- els, the key ingredients are the scales of global symme- try breaking f and explicit symmetry breaking (V0)1/4. One of the examples of a light hypothetical PNGB is the axion which possess extraordinarily feeble couplings to matter and radiation and is well-motivated dark mat- ter candidate [8]. If the axion exists, astrophysical and cosmological arguments constrain its mass to be in the range of m = 10−6 − 3 × 10−3 eV and the global symmetry-breaking scale to lie in a window f ≈ 107 GeV ×0.62/m(eV) = 2 × 109 − 6 × 1012 GeV [8]. We consider spherically symmetric system with metric ds2 = −N 2dt2 + g2dr2 + r2dΩ2, (2) where g, the radial metric, and N , the lapse, are functions of t and r with r being the circumferential radius. We introduce dimensionless coordinates and define the unit of distance, time and ϕ as r0 = ¯h mc , t0 = ¯h mc2 , ϕ0 = 1 √4πG , (3) where c is the speed of light, G is the gravitational con- stant, m = √V0/f is the particle mass. In dimensionless units the static Klein-Gordon and Einstein equations de- scribing the self-gravitating field ϕ and the metric are [9] ϕ′ g2 (cid:18) g2 + 1 r − 2rg2V(cid:19) + ϕ′′ g2 − ∂V ∂ϕ = 0, (4) N N ′ = g′ = 2 (cid:20) g2 − 1 r g 2 (cid:20) 1 − g2 r + r(cid:0)ϕ′2 − 2g2V(cid:1)(cid:21) , + r(cid:0)ϕ′2 + 2g2V(cid:1)(cid:21) , (5) (6) with boundary conditions g(0) = g(∞) = N (∞) = 1, g′(0) = N ′(0) = ϕ′(0) = 0, V (ϕ(∞)) = 0, where prime denotes ∂/∂r, V = 1 α2 [1 − cos(αϕ)], α = 1 √4πGf = mpl√4πf (7) is the dimensionless potential and the coupling parame- ter respectively, mpl = p¯hc/G = 1.2 × 1019 GeV is the Planck mass. The interaction potential has degenerate minima at ϕ = 2πn/α, where n is an integer number. Here we show that in the limit of strong nonlinearity, (4)-(6) have an approximate static solu- α ≫ 1, Eqs. tion that describes a spherical bubble with surface width much smaller then its radius R. The bubble surface is an interface between two degenerate vacuum states with ϕ = 2πn/α (r < R) and ϕ = 0 (r > R). Outside the bubble Eqs. (4)-(6) lead to the known Schwazschild solution g2 = 1 1 − 2M/r , N 2 = 1 − 2M r , (8) where M is the bubble mass in units of m2 pl/m. Let us assume that R ≫ α ≫ 1. Then, near the surface one can omit terms with 1/r in Eqs. (4)-(6) and take r ≈ R, we obtain − 2Rϕ′V + ϕ′′ g2 − ∂V ∂ϕ = 0, N ′ = N R 2 (cid:0)ϕ′2 − 2g2V(cid:1) , g′ = gR 2 (cid:0)ϕ′2 + 2g2V(cid:1) . (9) (10) (11) Eqs. (9)-(11) can be solved analytically. Their first inte- gral is N = const, ϕ′2 = 2g2V, g′ = Rgϕ′2. (12) We assume ϕ(0) = 2πn/α, where n = 1, 2, 3, . . . is the number of kinks at the bubble surface, and ϕ(r) mono- tonically decreases with r. Eqs. (12) yield 1 g = 1 − RZ ϕ(0) ϕ √2V dϕ, ϕ′ = − √2V 1 − RR ϕ(0) ϕ . √2V dϕ (13) For V (ϕ) given by Eq. (7) the final solution is 4R α2 ln sin(αϕ/2)+(cid:20)1 − 4R α2 (2m − 1)(cid:21) arctanh[cos(αϕ/2)] 2 FIG. 1: Scalar field ϕ as a function of distance r to the bubble center for bubbles with equal radius and different quantum numbers n = 1, 2, 3. The unit of length is ¯h/mc. Note, we plot the field ϕ only in the vicinity of the bubble surface where it undergoes variation. where Rm is a position of the mth kink, m = 1, 2, . . . , n. When the coordinate r passes through the point Rm the scalar field ϕ(r) changes from 2π(n − m + 1)/α to 2π(n − m)/α (see Fig. 1). Eq. (13) yields the following expression for g as a function of ϕ inside the bubble: 1 g = 1 − 4R α2 [2m − 1 + cos(αϕ/2)]. (15) Outside the bubble ϕ = 0, m = n and 1/g = 1 − 8nR/α2. The solution is valid if 1/g > 0, that is R < Rmax = α2/8n. Match of the inner solution (13) with the Schwazschild solution (8) determines the mass- radius relation M = 4πnuR2 − 8π2n2u2R3, (16) cosine potential (7) u = 2/πα2. where u is the surface energy density given by an integral over one potential period u = R √2V dϕ/4π. For the Redshift of the bubble interior z = 1/N − 1 can be found by matching the inner N = const and the outer (8) solutions: 1 1 1 − 4πnuR − 1. (17) z = p1 − 2M/R − 1 = The internal redshift monotonically increases from zero to infinity when the bubble radius R changes from zero to Rmax. Fig. 2 shows the redshift of space as a function of the distance r to the bubble center. The redshift is con- stant in the bubble interior and monotonically decreases outside the bubble. = sign[sin(αϕ/2)](r − Rm), ϕ ∈ [2π(n − m + 1)/α, 2π(n − m)/α], (14) Let us make rescaling M → M/4πu, R → R/4πu, then Eqs. (16), (17) yield M = nR2 − n2R3/2, z = 1 1 − nR − 1. (18) 3 n R, in α2¯h/mc 1 2 3 4 5 6 7 8 0.0329 0.0241 0.0204 0.0182 0.0168 0.0159 0.0153 0.0151 z 0.357 0.629 0.96 1.40 2.06 3.24 6.11 26.6 TABLE I: Redshift of the bubble interior z and its radius R for M = 0.00752α2 m2 pl/m and different kink numbers n. FIG. 2: Redshift z of space as a function of distance r to the bubble center for bubbles shown in Fig. 1. For a given M the redshift depends on the integer number n, which implies the redshift is quantized. In early samples of QSOs associated with nearby spiral galaxies, Karlsson showed that the redshift distribution has a periodicity log(1 + zn+1) − log(1 + zn) = 0.089, where n = 0, 1, 2, . . . and z0 = 0.061 [10]. It has been later confirmed by other groups [2, 11]. In a recent pa- per, Burbidge and Napier [3] tested for the occurrence of this periodicity in new QSO samples and found it to be present at a high confidence level. The peaks were found at z ≈ 0.30, 0.60, 0.96, 1.41 and 1.96 in agreement with Karlsson's empirical formula. The formula also includes the peak at z0 = 0.061, however, this peak does not occur for quasars, but for morphologically related objects. The redshift periodicity is observed only in QSO sam- ples satisfying certain selection criteria, in particular, the galaxies which are assumed to be paired to the QSOs must be most nearby spirals [1, 12]. This implies that redshift quantization is a property of intrinsically faint QSOs which are not detected from large distances. It is naturally to assume that QSOs born in the same type of galaxies have approximately equal masses because their formation mechanism must be similar. Such phe- nomenon is well known for type Ia supernovae or neu- tron stars: practically all measured neutron star masses cluster around the value of 1.4M⊙ with only a few per- cent deviation [13]. If dark matter bubbles are born with equal masses then, according to Eqs. (18), their red- shift must be quantized. For M = 0.0601 (in dimension units M = 0.00752α2m2 pl/m) Eqs. (18) have solutions for n = 1, 2, . . . , 8, they are given in Table 1. In Fig. 3 we plot the most recent histogram of the redshift distribution from Ref. [12] in which five peaks are clearly seen. The solid lines show the redshifts from our Table 1, they match well the observed peaks. The agreement is remarkable because the theory has only one free parameter, the bubble mass M . Such coincidence strongly suggests that the point-like quasars associated FIG. 3: Histogram of the redshift distribution of QSOs close to bright nearby active spiral galaxies or multiple QSOs with small angular separation from Ref. [12]. The solid lines rep- resent position of the peaks from Table 1. with nearby galaxies are dark matter bubbles composed of scalar particles with periodic interaction potential. One should mention an alternative possibility of quasar evolution. Bubbles can be originally born with the same mass and number of kinks n = 5 that corresponds to the 5th peak. During evolution the kinks tunnel to the bubble center and quasars sequentially decay into states with smaller n but the same mass. For axions with m = 0.1 − 3 meV and f = 2 × 109 − 6 × 1010 GeV Eq. (7) yields α = 5.6 × 107 − 1.7 × 109. Hence, an axion bubble with the internal redshift z = 0.36 and n = 1 would have the mass M = 3 × 107 − 109M⊙ and the radius R = 3× 102− 104R⊙. Such radius range agrees with the size of the emission region expected for the intrinsically faint QSOs. Indeed, for Seyfert 1 galaxies the size of the broad-line region is R ∼ 10 − 100 light days. The luminosity L of the QSOs is 5-6 orders smaller. Based on the empirical relation for Seyfert 1 galaxies R ∝ L1/2 [14], we obtain for the quasars R ∼ 104R⊙. Now we discuss the bubble life time. Under the in- fluence of surface tension and gravitational attraction an initially static bubble starts to collapse. In the thin-wall approximation the initial acceleration is given by [15] R = − 2N 3 R − N 2M (1 + N )R2 , (19) where N = p1 − 2M/R. For one kink thin-wall con- tracting bubble the conserved mass is [15] energy mc2 by a moving bubble surface is governed by the Boltzmann factor exp(−mc2/Teff ), where Teff = a/c is the effective temperature and a is the acceleration of the surface [19]. For the bubble a ≈ c2/R(t) and the Boltz- mann factor reduces to exp(−R(t)/l) where l = ¯h/mc is the surface width. Hence, emission of scalar particles is exponentially suppressed apart from small regions where R(t) < ∼ l. As a result, during one period of oscillation, tc ∼ R/c, the energy loss is ∆E ∼ (l/R)E, which yields the bubble life time 4 M = 4πuR2 p1 − (dR/dτ )2 − 8π2u2R3, (20) where τ is the interior coordinate time. Based on Eqs. (19), (20) one can expect a continuous contraction of the bubble to the origin on an astronomically short time scale R/c ≪ 1yr. However, so far we treated the scalar field as classical. Quantum corrections suppress the collapse and result in appearance of long-lived bubbles, stable on a scale more then hundreds million years. To include quantum effects it has been suggested that the expres- sion (20) be interpreted as the canonical hamiltonian of the bubble at the quantum level [16, 17, 18]. The bub- ble wave function Ψ(R) satisfies the following stationary quantum mechanical equation in one dimension (¯h = 1) [18]: (cid:20)(cid:0)E + 8π2u2R3(cid:1)2 + ∂2 ∂R2 − 16π2u2R4(cid:21) Ψ(R) = 0. (21) This equation possesses stationary solutions that are not possible in the classical model. Bubbles of non-negligible redshift correspond to highly excited stationary states for which the energy spectrum can be treated as quasi- continuous. At the quantum level the collapse is pre- vented by quantum pressure that balances the surface tension and gravitational attraction producing station- ary configurations. Let us estimate the decay time of an excited station- ary state of the quantum bubble. The decay occurs by means of scalar particle emission. We estimate the de- cay time using the Bohr correspondence principle as the time of energy loss by the classical bubble with the ra- dius R(t) oscillating between the turning points R(t) = R and R(t) = 0, where R is determined by Eq. (16). In the quantum picture, however, there are no such oscil- lations. The probability of creation a particle with the R l R2 cl . t ∼ (22) For an axionic bubble with R > 102R⊙ and l < 0.7 cm Eq. (22) yields t > 108 yrs which is the time we need to account for the phenomenon of quasars. tc = Properties of the intrinsically faint point-like QSOs, combined with equations for the bubble mass M = 0.00752α2m2 pl/m = 2.94m(eV) × 1011M⊙ and the radius R = 0.0329α2¯h/mc = 2.73m(eV) × 106R⊙, allow us to determine the axion mass m. The quasar luminosity sug- gests that the bubble radius is larger then 103R⊙ which yields m > 0.4 meV and M > 108M⊙. From the other hand, the quasar ejection from active galaxies implies that the bubble mass M must be much smaller then the galactic mass. It is reasonable to constrain M < 109M⊙ which leads to m < 3 meV and R < 104R⊙. We con- clude, the axion mass is m = 0.4− 3 meV. This value fits in the open window for the axion mass constrained by as- trophysical and cosmological arguments [8], which unam- biguously points towards the axionic nature of dark mat- ter composing the intrinsically faint point-like quasars. Current cavity search experiments in Livermore [20] and Kyoto University [8] are looking for the axion in the mass range 1 − 10 µeV which deviates by two orders of mag- nitude from our result. Probably now, when the axion mass is established from quasar observations, the axion has a better chance to be discovered. We mention that data on central "black hole" masses in small companion galaxies allow us to determine the axion mass more accurately and yield m = 1.0 − 1.9 meV. We will discuss this elsewhere. Moreover, observa- tions show that apart from the intrinsically faint point- like objects considered here there is a subgroup of bright quasars which probably also possess noncosmological red- shift. Tachyons, another dark matter candidate, can ex- plain their nature. We discuss this in a detail paper [21]. [1] K.G. Karlsson, A&A 239, 50 (1990). [2] H. Arp et al., A&A 239, 33 (1990). [3] G. Burbidge and W.M. Napier, AJ 121, 21 (2001). [4] E. Hawkins, S.J. Maddox and M.R. Merrifield, MNRAS 336, L13 (2002). [5] G.R. Burbidge et al., ApJS 74, 675 (1990); E.M. Bur- bidge et al., ApJ 591, 690 (2003); H. Arp et al., A&A 391, 833 (2002); A&A 418, 877 (2004). [6] H. Arp, Quasars, redshifts and controversies, Interstellar Media, Berkeley, 1987; Seeing red: redshifts, cosmology and academic science, Apeiron, Montreal, 1998. [7] C.T. Hill and G.G. Ross, Nucl. Phys. B311, 253 (1988). [8] R. Bradley et al., Rev. Mod. Phys. 75, 777 (2003). [9] E. Seidel and W. M. Suen, Phys. Rev. D 42, 384 (1990). [10] K.G. Karlsson, A&A 13, 333 (1971); 58, 237 (1977). [11] J.M. Barnothy and M.F. Barnothy, PASP 88, 837 (1976). [12] W.M. Napier and G. Burbidge, MNRAS 342, 601 (2003). [13] N.K. Glendenning, "Compact Stars: Nuclear Physics, Particle Physics, and General Relativity", Springer Ver- lag; New York, 2nd edition, (2000). [14] T.G. Wang and X.G. Zhang, MNRAS 340, 793 (2003). [15] S. K. Blau et al., Phys. Rev. D 35, 1747 (1987). [16] V.A. Berezin et al., Phys. Lett. B 212, 415 (1988). [17] A. Aurilia and E. Spallucci, Phys. Lett. B 251, 39 (1990). [18] A. Aurilia et al., Phys. Lett. B 262, 222 (1991). [19] A. Gorsky and K. Selivanov, Phys. Rev. D 62, 071702 (2000). [20] S.J. Asztalos et al., Phys. Rev. D 69, 011101(R) (2004). [21] A.A. Svidzinsky, astro-ph/0409064. 5
astro-ph/0508083
2
0508
2005-11-21T21:39:29
Definitive Identification of the Transition between Small- to Large-Scale Clustering for Lyman Break Galaxies
[ "astro-ph" ]
We report angular correlation function (ACF) of Lyman Break Galaxies (LBGs) with unprecedented statistical quality on the basis of 16,920 LBGs at z=4 detected in the 1 deg^2 sky of the Subaru/XMM-Newton Deep Field. The ACF significantly departs from a power law, and shows an excess on small scale. Particularly, the ACF of LBGs with i'<27.5 have a clear break between the small and large-scale regimes at the angular separation of ~7'' whose projected length corresponds to the virial radius of dark halos with a mass of 10^11-12 Mo, indicating multiple LBGs residing in a single dark halo. Both on small (2''<theta<3'') and large (40''<theta<400'') scales, clustering amplitudes monotonically increase with luminosity for the magnitude range of i'=24.5-27.5, and the small-scale clustering shows a stronger luminosity dependence than the large-scale clustering. The small-scale bias reaches b~10-50, and the outskirts of small-scale excess extend to a larger angular separation for brighter LBGs. The ACF and number density of LBGs can be explained by the cold dark matter model.
astro-ph
astro-ph
To Appear in the Astrophysical Journal Letters Preprint typeset using LATEX style emulateapj v. 6/22/04 5 0 0 2 v o N 1 2 2 v 3 8 0 8 0 5 0 / h p - o r t s a : v i X r a DEFINITIVE IDENTIFICATION OF THE TRANSITION BETWEEN SMALL- TO LARGE-SCALE CLUSTERING FOR LYMAN BREAK GALAXIES1 Masami Ouchi 2,3, Takashi Hamana 4, Kazuhiro Shimasaku 5, Toru Yamada 4, Masayuki Akiyama 6, Nobunari Kashikawa 4, Makiko Yoshida 5, Kentaro Aoki 6, Masanori Iye 4, Tomoki Saito 5, Toshiyuki Sasaki 6, Chris Simpson 7, and Michitoshi Yoshida 8 To Appear in the Astrophysical Journal Letters ABSTRACT We report angular correlation function (ACF) of Lyman Break Galaxies (LBGs) with unprecedented statistical quality on the basis of 16,920 LBGs at z = 4 detected in the 1 deg2 sky of the Subaru/XMM- Newton Deep Field. The ACF significantly departs from a power law, and shows an excess on small scale. Particularly, the ACF of LBGs with i′ < 27.5 have a clear break between the small and large-scale regimes at the angular separation of ≃ 7′′ whose projected length corresponds to the virial radius of dark halos with a mass of 1011−12M⊙, indicating multiple LBGs residing in a single dark halo. Both on small (2′′ < θ < 3′′) and large (40′′ < θ < 400′′) scales, clustering amplitudes monotonically increase with luminosity for the magnitude range of i′ = 24.5 − 27.5, and the small-scale clustering shows a stronger luminosity dependence than the large-scale clustering. The small-scale bias reaches b ≃ 10 − 50, and the outskirts of small-scale excess extend to a larger angular separation for brighter LBGs. The ACF and number density of LBGs can be explained by the cold dark matter model. Subject headings: large-scale structure of universe -- galaxies: formation -- galaxies: high-redshift 1. INTRODUCTION Recent observational studies have been found strong clustering in two-point angular correlation function (ACF) of Lyman break galaxies (LBGs) at z = 3−5, (e.g. Giavalisco & Dickinson 2001; Ouchi et al. 2001; Foucaud et al. 2003; Adelberger et al. 2003; Ouchi et al. 2004b; Hildebrandt et al. 2004; Allen et al. 2005), red galaxies at z = 3 (Daddi et al. 2003), and Lyα emitters (LAEs; Ouchi et al. 2003; Shimasaku et al. 2004) at z = 5. Even at z = 6, there is a piece of evidence for filamentary large (100 Mpc)-scale structures of LAEs (Ouchi et al. 2005). The distribution of high-z galaxies is fairly inho- mogeneous and highly biased against matter distribution predicted by the cold dark matter (CDM) model. The estimated bias is b ≃ 2−8, depending on luminosity/type and redshift of galaxies. However, the shape of the ACF for high-z galaxies is not well constrained. Ouchi et al. (2001) report a 3σ excess of the ACF at θ < 5′′ for z ∼ 4 LBGs, while Porciani & Giavalisco (2002) found a pos- sible deficit of the ACF at 10′′ . θ . 30′′ for bright LBGs which they interpret as the halo exclusion effect on hosting halos with a mass of 1012M⊙. In the local universe, the correlation function shows a departure from a power law (e.g. Hawkins et al. 2003; Zehavi et al. 2004). The departure is reproduced in the 1 Based on data collected at Subaru Telescope, which is oper- ated by the National Astronomical Observatory of Japan. 2 Space Telescope Science Institute, 3700 San Martin Drive, Bal- timore, MD 21218, USA; [email protected]. 3 Hubble Fellow 4 National Astronomical Observatory, Tokyo 181-8588, Japan 5 Department of Astronomy, School of Science, University of Tokyo, Tokyo 113-0033, Japan 6 Subaru Telescope, National Astronomical Observatory, 650 N.A'ohoku Place, Hilo, HI 96720, USA 7 Department of Physics, University of Durham, South Road, Durham DH1 3LE, UK 8 Okayama Astrophysical Observatory, National Astronomical Observatory, Kamogata, Okayama 719-0232, Japan framework of the halo occupation distribution (HOD) and the related halo models in the CDM cosmology (van den Bosch et al. 2003; Magliocchetti & Porciani 2003; Ze- havi et al. 2004; Benson et al. 2001; Berlind et al. 2003), and is explained by two sources contributing to the corre- lation function; one for galaxy pairs residing in the same halo (1-halo term) and the other for galaxies hosted by different halos (2-halo term; see, e.g., Zehavi et al. 2004). The HOD has been also applied to clustering of galaxies at high-z (Bullock et al. 2002; Moustakas & Somerville 2002; Hamana et al. 2004). However, parameters of the models have not been constrained with similar accuracy as at low-z because of the small sample (100−2000 galax- ies) and surveyed area (0.01 − 0.1deg2). In this paper, we present ACF of z = 4 LBGs with unprecedented statistical quality, on the basis of 16,920 LBGs obtained in the 1 deg2 sky of Subaru/XMM- Newton Deep Field (SXDF; Sekiguchi et al. 2004). Throughout this paper, magnitudes are in the AB sys- tem, and we adopt H0 = 70h70km s−1 Mpc−1 and [Ωm, ΩΛ, n, σ8] = [0.3, 0.7, 1.0, 0.9]. To facilitate compar- ison with previous results, we express r0 using h100, the Hubble constant in units of 100 km s−1 Mpc−1. 2. DATA AND SAMPLE We carried out deep optical broad-band imaging with Subaru/Suprime-Cam in the 1 deg2 sky of the SXDF. Our broad-band images reach B ≃ 28.3, V ≃ 27.3, R ≃ 27.6, i′ ≃ 27.5, and z′ ≃ 26.5 with a 2′′-diameter cir- cular aperture at the 3σ level (Furusawa et al. in prepa- ration). Typical seeing sizes (FWHM) of these images are 0′′.8. We use the i′-band selected source catalog of the SXDS V er1.0 produced with SExtractor (Bertin & Arnouts 1996), which is composed of 0.7 million objects with i′ < 27.5. We select LBGs at z = 4.0 ± 0.5 on the basis of the color criteria of Ouchi et al. (2004a), i.e., B − R > 1.2, R − i′ < 0.7, and B − R > 1.6(R − i′) + 1.9, 2 Ouchi et al. Fig. 1. -- The distribution of LBGs at z = 4.0 ± 0.5 in the SXDF. The red, blue, and black points denote the positions of the LBGs with i′ < 24.5 (bright), 24.5 ≤ i′ < 26.0 (interme- diate), and 26.0 ≤ i′ < 27.5 (faint), respectively. The gray ar- eas present masked regions where we did not use for our analy- sis. The scale on the map is marked in both degrees and (co- moving) megaparsecs for projected distance at z = 4.0. This figure is degraded. This paper with the original figure can be downloaded from http : //www − int.stsci.edu/ ∼ ouchi/work/astroph/sxds z4LBG/ouchi highres.pdf which were determined with the results of spectroscopy and Monte-Carlo simulations. We visually inspect all the candidates and mask areas contaminated with halos of bright stars and CCD blooming. Our final catalog in- cludes 16,920 LBGs in a 1.00 deg2 area (Table 1). Figure 1 shows the sky distribution of our LBGs. Our spec- troscopic follow-up observations show that 60 out of 63 identified candidates are real LBGs at z = 3.5 − 4.5; i.e., 17 out of 17 and 43 out of 46 are LBGs in the SXDF (Akiyama M. in preparation) and in the Subaru Deep Field, respectively, where the latter LBG sample is made with the same color criteria as ours (Yoshida 2005). Thus, the contamination rate of our LBG sample is estimated to be (63 − 60)/63 = 5%. 3. RESULTS AND DISCUSSION 3.1. Definitive Identification of Clustering Transition We derive the ACF, ω(θ), by the formula of Landy & Szalay (1993) with random samples composed of 200,000 sources, and estimate bootstrap errors (Ling et al. 1986). Since clustering properties of our 5% contaminants are not clear, we do not apply a correction for contaminants with the assumption of random distribution (c.f. Ouchi et al. 2004b). However, this correction changes ω(θ) and bias only by 10% or less. Figure 2 presents the ACF of LBGs (top panel), residuals of a power-law fit (middle panel), and galaxy-dark matter bias (bottom panel) de- fined as b(θ) ≡ pω(θ)/ωdm(θ), where ωdm(θ) is the ACF predicted by the non-linear model of Peacock & Dodds (1996). In the top and middle panels of Figure 2 the ACF of LBGs shows a significant excess on small scale, and indicates that a power law, Aωθ−β, does not fit the Fig. 2. -- T op : The ACF, ω(θ), of LBGs. The filled and open squares indicate the ACF with 1 σ bootstrap errors, while the open squares mean for ACF on very small scale which may include additional errors in source deblending and confusion. The solid line is the best-fit power law (Aωθ−β ) for 2′′ − 1000′′. The open circles are the ACF with IC correction under the assumption of the conventional power-law approximation, and the dashed lines are the best-fit power law for these open circles. The dotted curve is the ACF of dark matter predicted by the non-linear model of Peacock & Dodds (1996). The scale on the top axis denotes the projected distance in comoving megaparsecs at z = 4.0. The ticks labeled with R(1E10), R(1E11), R(1E12), R(1E13), and R(1E14) correspond to the predicted virial radii of dark halos, r200, with a mass of 1 × 1010, 1011, 1012, 1013, and 1014 h−1 70 M⊙, respectively. M iddle : The ratios of the ACF to the best-fit power law for our LBGs (squares), together with those for local galaxies (crosses; Zehavi et al. 2004). Bottom : The galaxy-dark matter bias, b, of LBGs as a function of separation. The dashed curve presents bias of local galaxies (Zehavi et al. 2004). The ticks with b(1E11), b(1E12), and b(1E13) show linear biases of dark halos with a mass of 1 × 1011, 1012, and 1013 h−1 70 M⊙, respectively, predicted by the CDM model of Sheth & Tormen (1999). data. This is the definitive identification of the depar- ture from a power law for the ACF of LBGs at z = 4. With a visual inspection, we confirm that all close-pairs of LBGs are not false detections. We also plot histogram of galaxy sizes for LBG pairs. We find that most of our LBGs have FWHM≃ 1′′ for pairs with any separations down to, at least, ≃ 2′′, and that extended LBGs do not boost small-scale ACF by producing false pairs. Uncer- tainties in source deblending and photometry can hardly account for the small-scale excess at & 2′′. In fact, a similar small-scale excess of ACF for z = 4 − 5 LBGs is also found by a recent study on high-resolution (∼ 0′′.1) HST images (Lee et al. 2005). Comparing our ACF with the one of dark matter, we find that the small-scale excess extends up to ≃ 7′′, i.e. 0.24h−1 70 Mpc, which is comparable to virial radius, r200, of dark halos with a mass of 1011−12M⊙ (see the ticks in the top panel of Figure 2), where r200 is a sphere of radius within which the mean enclosed density is 200 times the Transition of Clustering for LBGs 3 Fig. 3. -- The ACFs of magnitude-limited subsamples of LBGs at z = 4.0. In the top to third-top panels, the filled symbols are the ACFs of our LBGs with the limiting magnitude indicated in the legend. Each of these panels shows the ACF of i′ < 27.5 LBGs with open squares. The dotted curves are the ACF of dark matter predicted by the non-linear model of Peacock & Dodds (1996). The thick solid and dashed lines indicate the best-fit ACFs of the halo model and the breakdown of 1-halo and 2-halo terms for each subsample, while the thin lines are for i′ < 27.5 LBGs. In the bottom panels, biases of LBGs for the each magnitude-limited subsample are presented with the symbols which correspond to those marks found in the top to third-top panels. The plots of large-scale biases are magnified in the inserted boxes. mean cosmic value (Mo & White 2002). Interestingly, the large-scale average bias at 40′′ < θ < 400′′ is estimated to be 2.9±0.2 which is also comparable to linear bias of dark halos with a mass of 1011−12M⊙ (b = 2.2 − 3.5) predicted by the CDM (Sheth & Tormen 1999; see the ticks in the bottom panel of Figure 2). This coincidence of the dark- halo mass strongly supports that typical z = 4 LBGs reside in dark halos with a mass of 1011−12M⊙. More- over, these pieces of evidence suggest that multiple LBGs occupy a single dark halo. The middle panel of Figure 2 also plots residuals of a power-law fit for local galax- ies (Zehavi et al. 2004), which is comparable to those of z = 4 LBGs on large scale, but significantly larger than LBGs on intermediate scale (0.2 − 1.0 Mpc) correspond- ing to the radius of 1012−14M⊙ dark halos. According to the halo mass function (Sheth & Tormen 1999), the ratio of galaxy-sized halos (1010−12M⊙) to group/cluster-sized halos (1012−14M⊙) is about 10 times higher in number density at z = 4 than at z = 0. This relative deficit of group/cluster-sized halos at high-z would be the cause of the clearer break between small and large-scale ACFs at z = 4 than z = 0. Although multiple occupation of LBGs explains very consistently both the angular scale of transition and the amplitude of large-scale bias, there remains the possibility that the small-scale excess is en- hanced or produced by brightening of pair galaxies due to interactions. Fig. 4. -- Bias and slope of z = 4 LBGs as a function of limiting- absolute magnitude calculated from i′ − M1500 = 46.0. Top and middle panels present bias of small-scale (2′′ < θ < 3′′) and large- scale (40′′ < θ < 400′′) clustering. Filled circles plot for our LBGs, whose bias is directly measured from the ACFs. Open pentagons, squares, triangles, and diamonds present the large-scale (≃ 8h−1 100) bias estimated from the conventional power-law fit by Allen et al. (2005); Ouchi et al. (2004b, 2001); Arnouts et al. (2002). In the middle panel, the right-hand vertical scale means mass of dark halos corresponding to the linear bias (Sheth & Tormen 1999). The upper abscissa axis ticks number densities of our LBGs. Bottom panel shows the slope of a power law for the ACFs. Filled and open circles indicate the slopes for large-scale (40′′ < θ < 400′′) with no IC correction (βL), and for all scales (2′′ < θ < 1000′′) with IC correction (β; see Table 1), respectively. 3.2. Luminosity Dependence of Clustering We calculate ω(θ) and b for six subsamples with lim- iting magnitudes of i′ < 24.5, 25.0, 25.5, 26.0, 26.5, and 27.0 (Figure 3). We define the large- and small-scale bi- ases as the biases in the range 40′′ < θ < 400′′ (1−10h−1 100 Mpc) and 2′′ < θ < 3′′ (0.05 − 0.07h−1 100 Mpc), respec- tively, and show the biases in Figure 4. Although the angular range for the small-scale bias is somewhat arbi- trary, the range defined here is beyond the internal struc- tures of galaxies and below the radii of dark halos with 1011M⊙, and thus is sensitive to a multiplicity of LBGs in a halo. Luminosity segregation of large-scale clustering is reported for z = 4 LBGs (Ouchi et al. 2004b; Allen et al. 2005; Lee et al. 2005). In Figures 3 and 4, we find that ACFs and biases monotonically decrease from i′ < 24.5 to 27.5 on small scale as well as large scale. Interestingly, Figure 4 shows that the small-scale bias has a stronger dependence on luminosity (b ≃ 10 − 50) than the large- scale bias (b ≃ 3 − 4), and the bottom panels of Figure 4 Ouchi et al. 3 indicate that outskirts of small-scale excess extend to θ ∼ 10′′ for bright (i′ < 24.5 − 25.5) LBGs. All the fea- tures of luminosity dependence suggest that bright LBGs reside in more massive dark halos, since massive dark ha- los have not only a high large-scale bias, but also a high small-scale bias (i.e. high probability of pair galaxies in a massive halo) and an extended outskirt of bias due to a large halo size. Although the ACFs depart from a power law, we ap- proximate the ACFs with a power law, in order to com- pare our results with previous results. We fit the ACF over 2−1000′′ with ω(θ) = Aω(θ−β −IC/Aω), where IC is the integral constraint (Groth & Peebles 1977). The bot- tom panel of Figure 4 presents the best-fit slopes, β, as a function of magnitude. The slopes, β, become flatter at faint magnitudes (see also Kashikawa et al. 2005). This luminosity dependence of β is explained by the strong luminosity dependence of small-scale clustering as dis- cussed above. Then we calculate the Limber equation with the redshift distribution function of Ouchi et al. (2004b), and estimate the correlation lengths, r0, of spa- tial two-point correlation function, ξ = (r/r0)−γ, where γ = β + 1. Table 1 presents the best-fit parameters, Aω and β, together with r0. These r0 are consistent with those obtained by Ouchi et al. (2004b) as well as by Hildebrandt et al. (2004) and Kashikawa et al. (2005). However, our results are not consistent with those of small-sky surveys in HDF, if we assume that ACF does not significantly evolve between z = 3 and 4. For exam- ples, Giavalisco & Dickinson (2001) find a small corre- lation length of r0 = 1.0+0.8 100 Mpc for z = 3 LBGs, while we find a larger value, r0 = 3.8+0.2 100 Mpc, for our i′ < 27.5 LBGs whose number density is comparable to that of Giavalisco & Dickinson (2001). We restrict our power-law fitting to the same narrow range as Giavalisco & Dickinson (2001) (1′′ . θ . 20′′), and then we obtain the consistent results within errors, i.e. r0 = 1.3±0.3h−1 100 −0.7h−1 −0.2h−1 Mpc and β = 1.9 ± 0.2, due to the fitting only to the small-scale excess of the ACF (see Kravtsov et al. 2004). Similarly, the correlation length of red galaxies in HDF-S (r0 = 8h−1 100 Mpc; Daddi et al. 2003) is probably overes- timated by the extrapolation from a bump of small-scale ACF with a relatively flat slope of β = 0.8, which is also claimed by the model of Zheng (2004). 3.3. Comparison with a Halo Model We fit the halo model of Hamana et al. (2004) to the observed ω(θ) and number density, n, simultaneously. This model predicts the ω(θ) and n of galaxies con- tributed by a combination of the 1-halo and 2-halo terms in the framework of the CDM model. The best-fit mod- els are shown in Figure 3 and Table 1. These models account for the overall shape of our ACFs, i.e., the small- scale excess as well as the large-scale clustering (see also Lee et al. 2005), although there remain large residuals 9 (e.g. χ2/dof = 3.0 for i′ < 27.5 LBGs). Reducing these residuals results in a decrease in the combined likelihood (ω(θ) + n) from the best-fit value. The large residuals imply that we need a more precise model for our LBGs. For implications of the model fitting, Table 1 summarizes the average number of LBGs in a halo, hNgi, and the av- erage masses of the halo, hMhi, (Hamana et al. 2004) for the best-fit models. The average mass of hosting halos monotonically decreases from 2 × 1012h−1 70 M⊙ (i′ < 24.5) to 6 × 1011h−1 70 M⊙ (i′ < 27.5). The average number of LBGs in a halo is less than unity, hNgi ≃ 0.2 − 0.7, while the model ACF in Figure 3 shows a significant 1-halo term produced by multiple LBGs in one halo. This im- plies that majority of halos with an average mass have no LBG and only some halos host one or multiple LBG(s). We thank M. Fall, M. Giavalisco, K. Lee, S. Okamura, and Z. Zheng for helpful comments and discussion. 9 Note that bootstrap errors of the ACF are assumed to be independent. REFERENCES Adelberger, K. L., Steidel, C. C., Shapley, A. E., & Pettini, M. 2003, ApJ, 584, 45 Allen, P. D., Moustakas, L. A., Dalton, G., MacDonald, E., Blake, C., Clewley, L., Heymans, C., & Wegner, G. 2005, MNRAS, 523 Arnouts, S., et al. 2002, MNRAS, 329, 355 Bullock, J. S., Wechsler, R. H., & Somerville, R. S. 2002, MNRAS, Landy, S. D., & Szalay, A. S. 1993, ApJ, 412, 64 Lee, K. et al. 2005, submitted to ApJ Ling, E. N., Barrow, J. D., & Frenk, C. S. 1986, MNRAS, 223, 21P Magliocchetti, M., & Porciani, C. 2003, MNRAS, 346, 186 Mo, H. J. & White, S. D. M. 2002, MNRAS, 336, 112 Moustakas, L. A., & Somerville, R. S. 2002, ApJ, 577, 1 2002, 329, 246 Benson, A. J., Frenk, C. S., Baugh, C. M., Cole, S., & Lacey, C. G. 2001, MNRAS, 327, 1041 Berlind, A. A., et al. 2003, ApJ, 593, 1 Bertin, E. & Arnouts, S. 1996, A&AS, 117, 393 al. 2002, ApJ, 579, 42 Daddi, E., et al. 2003, ApJ, 588, 50 Foucaud, S., McCracken, H. J., Le F`evre, O., Arnouts, S., Brodwin, M., Lilly, S. J., Crampton, D., & Mellier, Y. 2003, A&A, 409, 835 Giavalisco, M. & Dickinson, M. 2001, ApJ, 550, 177 Giavalisco, M. 2005, "Wide-Field Imaging from Space", Eds. T. McKay, A. Fruchter, and E. Linder. Elsevier, in press Groth, E. J. & Peebles, P. J. E. 1977, ApJ, 217, 385 Hamana, T., Ouchi, M., Shimasaku, K., Kayo, I., & Suto, Y. 2004, MNRAS, 347, 813 Hawkins, E., et al. 2003, MNRAS, 346, 78 Hildebrandt, H., et al. 2004, ArXiv Astrophysics e-prints, arXiv:astro-ph/0412375 Kashikawa, N. et al. 2005, submitted to ApJ Kravtsov, A. V., Berlind, A. A., Wechsler, R. H., Klypin, A. A., Gottlober, S., Allgood, B., & Primack, J. R. 2004, ApJ, 609, 35 MNRAS, 332, 827 Ouchi, M. et al. 2001, ApJ, 558, L83 Ouchi, M. et al. 2003, ApJ, 582, 60 Ouchi, M., et al. 2004a, ApJ, 611, 660 Ouchi, M., et al. 2004b, ApJ, 611, 685 Ouchi, M., et al. 2005, ApJ, 620, L1 Peacock, J. A., & Dodds, S. J. 1996, MNRAS, 280, L19 Porciani, C., & Giavalisco, M. 2002, ApJ, 565, 24 Sekiguchi, K. et al. 2004, Astrophysics and Space Science Library, 301, 169 Sheth, R. K. & Tormen, G. 1999, MNRAS, 308, 119 Shimasaku, K., et al. 2004, ApJ, 605, L93 van den Bosch, F. C., Yang, X., & Mo, H. J. 2003, MNRAS, 340, 771 Yoshida, M. 2005, Master Thesis, University of Tokyo Zehavi, I., et al. 2004, ApJ, 608, 16 Zheng, Z. 2004, ApJ, 610, 61 Transition of Clustering for LBGs 5 Summary of Clustering Properties TABLE 1 Conventional Power-Law Approx. Modele i′ AB (mag) N (< i′)a n(< i′)b 70 Mpc−3) (h3 24.5 25.0 25.5 26.0 26.5 27.0 27.5 239 808 2231 4891 8639 12921 16920 9.8 ± 1.6 × 10−5 2.8 ± 0.3 × 10−4 6.4 ± 0.6 × 10−4 1.3 ± 0.1 × 10−3 2.2 ± 0.3 × 10−3 3.7 ± 0.7 × 10−3 5.8 ± 1.4 × 10−3 c Aω (arcsecβ ) 10.5 ± 8.2 5.0 ± 9.1 3.1 ± 1.6 2.6 ± 0.6 0.6 ± 0.1 0.8 ± 0.1 0.5 ± 0.1 βc c r0 d βL hN gi (h−1 100 Mpc) 4.9+4.3 −4.1 5.5+1.7 −2.1 5.0+0.7 −0.8 5.0+0.4 −0.4 4.8+0.2 −0.3 4.4+0.1 −0.2 3.8+0.2 −0.2 1.1 ± 0.4 0.9 ± 0.3 0.8 ± 0.1 0.8 ± 0.1 0.5 ± 0.1 0.6 ± 0.1 0.5 ± 0.1 ≃ 1.6 0.9 ± 0.6 0.8 ± 0.4 1.0 ± 0.2 0.7 ± 0.2 0.7 ± 0.1 0.6 ± 0.1 0.2+0.2 −0.2 0.3+0.4 −0.3 0.6+0.1 −0.5 0.6+0.1 −0.1 0.6+0.1 −0.1 0.6+0.1 −0.2 0.7+0.2 −0.1 log hMhi (h−1 70 M⊙) 12.3+0.1 −0.6 12.3+0.1 −0.2 12.1+0.1 −0.1 12.0+0.1 −0.1 11.9+0.05 −0.05 11.8+0.07 −0.04 11.8+0.02 −0.05 aCumulative numbers. Differential surface densities are 0.002 ± 0.001, 0.014 ± 0.002, 0.049 ± 0.004, 0.158 ± 0.007, 0.395 ± 0.011, 0.739 ± 0.014, 1.041 ± 0.017, 1.189 ± 0.018, and 1.111 ± 0.018 arcmin−2 (0.5mag)−1 for i′ = 23.25, 23.75, 24.25, 24.75, 25.25, 25.75, 26.25, 26.75, and 27.25, respectively, which are consistent with previous measurements (e.g. Ouchi et al. 2004a). bCumulative number density calculated from luminosity function of Giavalisco (2005). cResults from the conventional power-law approximation, i.e. ω(θ) = Aω(θ−β − IC/Aω), over 2′′ − 1000′′. For integral constraints, IC, we apply IC/Aω = [3, 14, 21, 28, 364, 154, 293] × 10−4 for i′ = [24.5, 25.0, 25.5, 26.0, 26.5, 27.0, 27.5]. dPower-law slope for the fit of ω = AωLθ−βL over 40′′ − 400′′ with no IC correction. e χ2/dof = [0.7, 0.4, 2.5, 6.9, 8.3, 7.1, 3.0] for i′ = [24.5, 25.0, 25.5, 26.0, 26.5, 27.0, 27.5].
astro-ph/0211575
1
0211
2002-11-26T16:02:27
Rayleigh - Taylor Gravity Waves and Quasiperiodic Oscillation Phenomenon in X-ray Binaries
[ "astro-ph" ]
Accretion onto compact objects in X-ray binaries [black hole, neutron star (NS), white dwarf] is characterized by non-uniform flow density profiles. Such an effect of heterogeneity in presence of gravitational forces and pressure gradients exhibits Raylegh-Taylor gravity waves (RTGW). They should be seen as quasioperiodic wave oscillations (QPO). In this paper I show that the main QPO frequency, which is very close to the Keplerian frequency, is split into separate frequencies (hybrid and low branch) under the influence of the gravitational forces in the rotational frame of reference. The observed low and high QPO frequencies are an intrinsic signature of the RTGW. I elaborate the conditions for the density profile when the RTGW oscillations are stable. A comparison of the inferred QPO frequencies with QPO observations is presented. I find that hectohertz frequencies detected from NS binaries can be identified as the RTGW low branch frequencies. I also predict that an observer can see the double NS spin frequency during the NS long (super) burst events when the pressure gradients and buoyant forces are suppressed. The Coriolis force is the only force which acts in the rotational frame of reference and its presence causes perfect coherent pulsations with a frequency twice of the NS spin.
astro-ph
astro-ph
Rayleigh - Taylor Gravity Waves and Quasiperiodic Oscillation Phenomenon in X-ray Binaries Lev Titarchuk 1,2 ABSTRACT Accretion onto compact objects in X-ray binaries (black hole, neutron star (NS), white dwarf) is characterized by non-uniform flow density profiles. Such an effect of heterogeneity in presence of gravitational forces and pressure gra- dients exhibits Raylegh-Taylor gravity waves (RTGW). They should be seen as quasioperiodic wave oscillations (QPO) of the accretion flow in the transition (boundary) layer between the Keplerian disk and the central object. In this pa- per I show that the main QPO frequency, which is very close to the Keplerian frequency, is split into separate frequencies (hybrid and low branch) under the influence of the gravitational forces in the rotational frame of reference. The RTGWs must be present and the related QPOs should be detected in any sys- tem where the gravity, buoynancy and Coriolis force effects cannot be excluded (even in the Earth and solar environments). The observed low and high QPO frequencies are an intrinsic signature of the RTGW. I elaborate the conditions for the density profile when the RTGW oscillations are stable. A comparison of the inferred QPO frequencies with QPO observations is presented. I find that hectohertz frequencies detected from NS binaries can be identified as the RTGW low branch frequencies. I also predict that an observer can see the double NS spin frequency during the NS long (super) burst events when the pressure gradients and buoyant forces are suppressed. The Coriolis force is the only force which acts in the rotational frame of reference and its presence causes perfect coherent pulsations with a frequency twice of the NS spin. The QPO observations of neu- tron binaries have established that the high QPO frequencies do not go beyond of the certain upper limit. I explain this observational effect as a result of the density profile inversions. Also I demonstrate that a particular problem of the gravity waves in the rotational frame of reference in the approximation of very small pressure gradients is reduced to the problem of the classical oscillator in 1George Mason University/CEOSR/NRL; [email protected] 2NASA Goddard Space Flight Center, code 661, Laboratory for High Energy Astrophysics, Greenbelt MD 20771; [email protected] -- 2 -- the rotational frame of reference which was previously introduced and applied for the interpretation of kHZ QPO observation by Osherovich & Titarchuk. Subject headings: Accretion, accretion disks -- (magnetohydrodynamics:) MHD -- stars:oscillations (including pulsations) -- stars: neutron -- X-ray:binaries 1. Introduction The theory of oscillations of rotating fluids of variable density [Rayleigh- Taylor (R-T) effect] was developed in detail by Chandrasekhar (1961), hereafter C61. A large variety of the magnitohydrodynamic (MHD) problems, including the stability of inviscid and viscous fluids in the case of two uniform layers separated by a horizontal boundary with and without rotation as well as the magnetic field effects were analyzed using perturbation technique. The simplest case of a one-dimensional gravitational force was studied. A similar analysis was also implemented for the case of an exponentially varying density. It follows from C61 that the quasiperiodic oscillations (QPO) with the twin "kilohertz" frequencies represent the main frequencies of the stable gravity waves in the rotational frame of reference. It also as follows from C61 that the twin "kilohertz" frequencies should be on the order of the Keplerian frequency. In this Paper I present the results of the study of the R-T effect for a particular case of fluid oscillations in an accretion flow under influence of a central gravitational force. This particular R-T analysis is important in view of the high and low frequency detection by the Rossi X-ray Timing Explorer (RXTE) in a number of low mass X-ray binaries (Strohmayer et al. 1996, van der Klis et al. 1996), black hole candidate sources (Morgan, Remillard & Greiner 1997; Strohmayer 2001a,b; Remillard 2002) and by Extreme Ultraviolet Explorer, Chandra X-ray observatory and optical observations in white dwarfs (Mauche 2002 and Woudt & Warner 2002). The presence of two observed peaks with frequencies ν1 and ν2 in the upper part of the power spectrum became a natural starting point in modeling the phenomena. In NS binaries, for example Sco X-1, the lower frequency part of the power spectrum, contains two horizontal branch oscillation (HBO) frequencies νHBO ∼ 45 Hz and ν2HBO ∼ 90 Hz (probably the second harmonic of νHBO) which slowly increase with the increase of ν1 and ν2 (van der Klis et al. 2000). Any plausible model faces the challenging task of describing the dependences of the peak separation ∆ν = ν2 − ν1 on ν1 and ν2 . Attempts have been made to relate ν1 and ν2 and the peak separation ∆ν = ν2 − ν1 with In the sonic point beat frequency model by Miller, Lamb & the neutron star (NS) spin. Psaltis (1998) the kHz peak separation ∆ν is considered to be close to the NS spin frequency and thus ∆ν is predicted to be constant. However observations of kHz QPOs in a number -- 3 -- of binaries (Sco X-1, 4U 1728-34, 4U 1608-52, 4U 1702-429 and etc) show that the peak separation decreases systematically when the high [kilohertz (kHz)] frequencies increase (for a recent review see van der Klis 2000, hereafter VDK). For Sco X-1 VDK found that the peak separation of kHz QPO frequencies changes from 320 Hz to 220 Hz when the lower kHz peak ν1 changes from 500 Hz to 850 Hz. The correlation between high frequency (lower kHz frequency) and low frequency (broad noise component) QPOs previously found by Psaltis, Belloni & van der Klis (1999) for black hole (BH) and neutron star (NS) systems has been recently extended over two orders of magnitude by Mauche (2002) to white dwarf (WD) binaries. Accepting the reasonable assumption that the same mechanism produces the QPO in WD, NS and BH binaries, one can argue that the data exclude relativistic models and, beat frequency models as well as any model requiring either the presence or absence of a stellar surface or a strong magnetic field. The transition layer model (TLM) was introduced by Titarchuk, Lapidus & Muslimov (1998), hereafter TLM98, to explain the dynamical adjustment of a Keplerian disk to the innermost sub-Keplerian boundary conditions (it is at the star surface for NSs and WDs). TLM98 argued that a shock should occur where the Keplerian disk adjusts to the sub- Keplerian flow. Thus the transition layer bounded between the sub-Keplerian boundary and the adjustment radius can undergo various type of oscillations under the influence of the gas, radiation, magnetic pressure and gravitational force. Osherovich & Titarchuk (1999), hereafter OT99, suggested that the phenomological model of a one-dimensional classical oscillator in the rotational frame of reference could explain the observed correlations between twin kHz frequencies ν1, ν2 and the HBO frequencies. They further suggested that the oscillations of the fluid element that bounced from the disk shock region (at the adjustment radius) would be seen as two independent oscillations parallel and perpendicular to the disk plane respectively. This is due to the presence of a Coriolis force in the magnetospheric rotational frame of reference. In this paper I show that when the pressure gradients can be neglected, the problem of the Rayleigh-Taylor wave oscillations (buoynancy effect) in the rotational frame of reference is reduced to the OT99 formulation (see §3). This result provides a solid basis for application of the OT99 model for interpretation of the QPO phenomena observed in X-ray binaries. The main goal of this Paper is to demonstrate that there is an inevitable effect of the gravity wave oscillations in the heterogeneous fluid of the accretion flow near compact objects. In §2, I formulate the problem of the gravity wave propagation in the bounded medium in the rotational frame of reference. In §3 I present analysis and solutions of this problem using a perturbation method in the context of three dimensional periodic waves with assigned wave numbers. Applications of the gravity modes and their relations with the -- 4 -- QPO observations is presented in §4. Summary and conclusions are drawn in §5. 2. The Problem of Gravity Wave Propagation in a Bounded Medium Lord Rayleigh (1883) has treated the non-rotating inviscid case of the present problem. He developed a general theory for any density configuration ρ0(z) (the z axis being the upward drawn vertical). Rayleigh's treatment of inviscid superposed fluids was extended by Bjerkness et al. (1933) to include the influence of rotation. With an assumption of the gravitational force directed vertically, Hide (1956) developed the theory for any density ρ0(z) and viscosity µ0(z) profiles and any δ (where δ is an angle between the rotational axis and the vertical). In this study we consider a case of an inviscid and incompressible fluid. The equation of relative motion appropriate to the problem is ρ ∂ui ∂t + ρui ∂ ∂xj ui − 2ρΩǫijkujsk = − ∂p ∂xj − gρei, (1) where the fluid is supposed to rotate uniformly about an axis whose direction is specified by the unit vector s = Ω/Ω. The tensor notation follows the summation convention and the unit vector e in the radial direction is introduced. In equation (1) ρ denotes the density, Ω the angular velocity of rotation, ui the ith component of the (Eulerian) velocity vector, p is the pressure and g is the acceleration due to gravity. This term g can represent the effective gravity which includes the centrifugal and radiation pressure forces. For an incompressible fluid the continuity equation is ∂ui ∂xi = 0. (2) Because (similar to C61)diffusion effects are ignored in this analysis, an individual fluid element retains the same density throughout its motion. Hence it is required that Dρ Dt = ∂ρ ∂t + uj ∂ρ ∂xj = 0. (3) Because the equilibrium situation in the comoving frame is a static one it is characterized by ui = 0. We now assume the equilibrium situation to be slightly disturbed, so that ui 6= 0. So we shall write ρ = ρ0(r, z) + δρ(x, y, z), p = p0(z) + δp(x, y, z, t). (4) and treat ui, δρ, and δp as quantities of the first order of smallness so that products of such quantities can be ignored. -- 5 -- We are interested in the study of the gravity wave oscillations in the disk transition layer where the accretion flow is sub-Keplerian and the temperature of the flow is of order 5 keV or more (see cartoon diagram of the system in Titarchuk, Osherovich & Kuznetsov 1999, hereafter TOK, Fig. 1). These temperatures of order of 5 keV are a representative values for plasma temperatures inferred from X-ray spectra during kHz QPO events (see more in Titarchuk, Bradshaw & Wood 2001 and Titarchuk & Wood 2002). At this stage we introduce the Cartesian coordinate system. We take the z axis to be in the direction of the (upward) vertical and the x axis to be such that the (x, z) plane contains the angular velocity vector Ω which then has components (Ωx, 0, Ωz). In the general case the gravitational force has components g(x/R, y/R, z/R) where R = (x2 + y2 + z2)1/2. If (u, v, w) are the components of the velocity u, then on substituting in equations (1), (2) and (3) we find ρ0 ∂u ∂t − 2ρ0Ωzv = − ∂v ∂t − 2ρ0Ωxw + 2ρ0Ωzu = − ρ0 δp δx − gxδρ, δp δy − gyδρ, δp δz − gzδρ, + 2ρ0Ωxv = − ∂v = 0, + ∂y ∂u ∂x + ρ0 ∂w ∂t ∂δρ ∂t + u ∂ρ0 ∂x + v ∂ρ0 ∂y + w = 0. ∂w ∂z ∂ρ0 ∂z (5) (6) (7) (8) (9) One can see that the variation of ρ is ignored in all terms except for the ones representing the buoyancy force (see C61 for details). 3. Analysis of various types of the solutions of the problem In order to illustrate the buoyancy (R-T) effect we consider the simplest solutions of the problem including those which already exist in the literature. Case A In the case when we allow the vector Ω to rotate around the z axis at angle δ, we introduce the cylindrical coordinate system, namely taking the z axis in the direction of the (upward) vertical, the x axis as the radial axis and the y axis as the azimuthal axis in the horizontal plane. Then the vector Ω has components (Ω sin δ, 0, Ω cos δ) and gravitational force vector has components g(r/R, 0, z/R) where r = (x2 + y2)1/2. -- 6 -- (I) With assumptions that ρ0 is only a function of R the scale height of ρ0 is order of R we obtain that δρ = [(dρ/dx)dx + (dρ/dz)dz) ∼ ρ0/R[(r/R)χ + (z/R)ζ] where χ = dx, ζ = dz are the radial and vertical components of displacement respectively and conse- quently the radial and vertical components of the perturbation velocity are u = dχ/dt, w = dζ/dt. Thus the vector G = δρ(gx, 0, gz) is transformed into G = ρ0(g/R)[(r/R)(rχ/R + zζ/R), 0, (z/R)(rχ/R + zζ/R]. Furthermore if we neglect the effects of pressure gradients, the set of equations (5-7) can be reduced to the system where the right hand side consists of only the vector G. Then, for radial, azimuthal and vertical displacements χ, Υ, ζ this system takes the form: χ − 2Ω cos δ Υ = −(g/R)(r/R)(rχ/R + zζ/R), Υ + 2Ω cos δ χ − 2Ω sin δ ζ = 0, ζ + 2Ω sin δ Υ = −(g/R)(z/R)(rχ/R + zζ/R). (10) (11) (12) OT99 have already analyzed the solution of Eqs. (10-12) (see Eqs. 2-4, in OT99) in the case of z/R ≪ 0 and δ ≪ 1. They found that in the rotational frame of reference, the radial oscillation with the main frequency ωK = (g/R)1/2 is split to the oscillations taking K + 4Ω2)1/2 place near the horizontal (disk) plane (χΥ) with the hybrid frequency ωh = (ω2 and to oscillations taking place near the vertical plane (Υζ) with the low branch frequency ωL = 2Ω(ωK/ωh) sin δ. For z ∼ r the dispersion equation for the frequency ω ω2[ω4 − 2(ω2 ∗ + 2ω2)ω2 + ω4 ∗ + 4Ω2ω2 ∗ + 2Ω2ω2 ∗ sin 2δ] = 0 (13) besides the nonoscillatory mode (ω = 0), describes two oscillatory eigenmodes. For δ ≪ 1 they are ω1 = ω∗ = (g/2R)1/2 = ωK/√2 and ω2 = (ω2 ∗ + 4Ω2)1/2 = (ω2 K/2 + 4Ω2)1/2. The relation of the model eigenfrequencies νL = ωL/2π, νK = ωK/2π and νh = ωh/2π with the QPO observed frequencies: horizontal branch oscillation freqiencies νHBO, kHz frequencies were studied by OT99; TOK, Kuznetsov & Titarchuk (2002); and Titarchuk (2002), hereafter T02. It is worth noting that the relation between νK, Ω, νh and νL predicts the existence of the invariant quantity δ. (e.g. Titarchuk & Osherovich 2001). T02 calculated δ and its uncertainty of δ finding that the inferred δ−values are consistent with being constant at least for four Z sources, Sco X-1, GX 340+0, GX 5-1, GX 17+2 (see more on this issue in section §4). (II). With assumptions that g∆ρ ≪ 1 and ∇p ≪ 1 the set of equations for χ, Υ, ζ is similar to Eqs (10-12) where the right hand side vector is the zero vector, 0 = (0, 0, 0). In this case the dispersion equation for ω ( for small δ << 1) ω2(ω2 − 4Ω2) = 0 (14) -- 7 -- has the only one nontrivial root ω = 2Ω which is related to the eigenmode oscillations taking place parallel to the disk (χΥ)-plane. We discuss the application of this solution to the observation in section §4. Case B If we do not neglect the pressure gradient and assume the vertical gravitational force and the vector Ω directed along the vertical (δ = 00), then the entire problem (Eqs. 5-9) is reduced to set of equations which have already been analyzed by Chandrasekhar in C61. He studied two cases: (1) In the case of two uniform fluids separated by a horizontal boundary he showed that for two adjacent, hydrostatic, inviscid fluids, the low fluid having density ρ1 and the upper layer having density ρ2 the eigenfrequency νR−T was νR−T = [2(Ω/2π)2 + (4(Ω/2π)4 + ν 4 0 )1/2]1/2 (15) where ν0 is the frequency in the absence of rotation. The frequency ν0 is of order of νK = (g/R)1/2 if the k-wave number (ν0 depends on k and density difference ∆ρ = ρ2 − ρ1) of order R−1 and ∆ρ/(ρ1 + ρ2) ∼ (0.5 − 1). For a stratified medium of density ρ = C0 exp(βz), where 1/β is the scale height, the R-T instability occurs for positive β, whereas stable gravity waves occur for negative β. The 0 + 4(Ω/2π)2]1/2 if one assumes that the wave number k is R-T frequency νR−T = νh = [ν 2 of order of d−1 ∼ H −1 (where d is a layer size). It is worthwhile to emphasize that even in the simplest case of the vertical gravitational force (C61) the hybrid frequency νh is an eigenfrequency (which is also true for the case with no pressure gradient effects, see above). Formula (10) for νR−T can also be reduced to νh if one assumes that ν0 ∼ 2(Ω/2π). Case C Now we consider the case when the vector Ω rotates uniformly around the vertical at angle δ, then the gravitational vector G = δρ(gr/R, 0, zg/R), also taking into account the pressure effects. The general case of G = δρ(gx, gy, gz) can be analyzed in a similar way and it will be presented elsewhere. Following the usual practice in problems of this kind, we seek solutions of equations (5-9) which are of form (see e.g. C61) u, v, w, δρ. δp = constant × exp(ikxx + ikyy + ikzz + nt), (16) where kx, ky and kz are the horizontal and vertical wave numbers of the harmonic perturba- tions respectively. We also assume that ρ0 = f (x)ϕ(z) is a function of x (or r) and z which the scale heights are 1/γ and 1/β for radial and vertical density profiles respectively, r ≈ R. Upon substituting for u, v, w, δρ, δp in the form (16) equations (5) to (8) become nu − 2Ωzv = −ikx(δp/ρ0) + guγ/n + gwβ/n, (17) -- 8 -- nv − 2Ωxw + 2Ωzu = −iky(δp/ρ0), nw − 2Ωxv = −ikz(δp/ρ0) + (zg/R)uγ/n + (zg/R)wβ/n, kxu + kyv + kzw = 0. (18) (19) (20) We also use the relation nδρ + (γu + βw)ρ0 = 0 which follows from Eqs. (9) and (16) in order to express δρ through ρ0, u and w. The set of equation (17-20) assumes a nontrivial solution only if the determinant of the system D = 0. This equation provides the dispersion relation for the determination of n: P (n) = a4n4 + a2n2 + a1n + a0 = 0, (21) where a0 = k2 a2 = 4(Ωxkx + Ωzkz)2 − gβ[−(γ/β)k2 k2 = k2 y + k2 z . yg2γ 2(kz/kx − z/R)(β/γ − kz/kx), a1 = −βkyg(2Ωxkx + 2Ωzkz)(1 − γz/βR), y(γ/β + z/R)], a4 = x(β/γ − kz/kx)(kz/kx − z/R) + k2 x + k2 3.1. Stable gravity modes The specific wave values kx, ky, kz are determined by the conditions imposed on the oscillatory domain boundary. Thus for a given set of boundary conditions the analysis of the R-T instability is reduced to the analysis of the roots of of fourth order algebraic equation (21), which depends on the main parameters of the atmosphere, β, γ and ratio of the wave numbers kz/kx. We assume that z/R < 1 and kx, ky ∝ 1/R, kz ∝ 1/z is the case of interest. Figure 1 illustrates the specific behavior of the polynomial P(n) which should help one to understand the presence (or absence) of its roots for given coefficient a0, a1, a2 (a4 > 0, a3 = 0). For example, if a0 < 0, a1 < 0 and a2 > 0 (see case i below) then P (n) has two complex conjugate roots n1,2 = ζ ± iη and two real roots n3, n4 (in fact, d2P/dn2 > 0). Because a1 < 0, the absolute value of the positive root n4 is larger than that of the negative one n3. But a3 is a sum of the polynomial (real and complex conjugate) roots, and because of a3 = n1+n2+n3+n4 = 0 one can come to conclusion that 2ζ = −(n3+n4) < 0. Therefore (in this case) the stable oscillatory mode exists and n1,2 are related to these damped oscillations. Below we present all cases with the oscillatory stable solutions: Case i: β, β/γ > 0, β/γ − kz/kx < 0 and a2 > 0 (see Fig. 1, curve i). For such conditions a0 < 0 and a1 < 0. Then equation (21) has two real roots which relate to the unstable (growing) and stable (decaying) modes with one pair of complex conjugate roots corresponding to the stable oscillatory mode: n1,2 ≈ a1/2d ± iωh, (22) -- 9 -- 2 − 4a4a0)1/2. These roots of equation (21) are where ωh = [(a2 + d)/2a4]1/2 and d = (a2 found using the sequential approximation method: first we solve equation (21) with a1 = 0, n and then in the next stage we look for roots of equation (21) as n = n + α. Case ia: β, β/γ > 0, β/γ − kz/kx < 0 and a2 < 0. For such conditions a0 < 0 and a1 < 0. Equation (21) has just one pair of the complex conjugate roots and two real roots which relate to the unstable (growing) and stable (decaying) modes (see Fig. 1, curve ia). The oscillatory mode n1,2 is stable for which we have n1,2 ≈ a1/2d ± iωL, (23) where ωL = [(−a2 + d)/2a4]1/2. Case ib: β > 0, β/γ < 0, β/γ − kz/kx < 0. For such conditions a0 < 0 and a1 < 0 and a2 > 0. This case is similar to case i (see Fig 1, curve i) when equation (21) has two real roots which relate to the unstable (growing) and stable (decaying) modes and complex conjugate roots which correspond to the stable oscillatory mode: n1,2 ≈ a1/2d ± iωh. (24) Case ii: β > 0, β/γ − kz/kx > 0. For such conditions a0 > 0 and a1 < 0 and a2 > 0. Equation (21) has a pair of complex conjugate roots (see Fig. 1, curve ii) and one of them n1,2 is related to the stable oscillatory mode: n1,2 ≈ a1/2d ± iωh. (25) Case iii: β < 0, β/γ − kz/kx > 0, and a2, d > 0. For such conditions a0 > 0 and a1 > 0. Equation (21) has a pair of complex conjugate roots (see Fig. 1, curve iii) and one of them n3,4 is related to the stable oscillatory mode: n1,2 ≈ −a1/2d ± iωL, (26) where ωL = [(a2 − d)/2a4]1/2. Finally, we single out the case when the effective gravitational force goes to zero. This can happen during a burst event, when the gravitational forces are compensated by the radiation pressure forces (e.g. Titarchuk 1994). For such conditions, a0, a1 → 0 and a2/a4 = 4[(kx/k)Ωx/k + (kz/k)Ωz]2. Equation (21) has zero roots n = 0 and conjugate complex roots n1.2 = ±i(a2/a4)1/2 = ±2i[(kx/k)Ωx + (kz/k)Ωz] (27) which are related to a pure harmonic mode. In fact, this result also follows from treatments detailed in C61 and OT99. The hybrid frequency then becomes simply νh = 2(Ω/2π) for -- 10 -- g = 0. This case of the perfect coherent oscillations would also occur if the density profile is quasi-uniform i.e. when β, γ → 0 but β/γ = O(1). These two conditions (on either the effective gravity force or the density profile) can be realized during the burst event. In this section the main goal is to reveal all cases when stable gravity modes exist and when the stability breaks down. This analysis is particularly important in a view of the transient nature of QPO features (see e.g. Zhang et al. 1998). With an increase of bolometric luminosity (presumably in mass accretion rate) the kHz QPO frequency increases and then entirely disappears! At low rates it appears once again! The stability analysis presented here leads to conditions for the existence and the destruction of gravity modes in terms of density profiles scale heights β −1, γ −1 and boundary conditions (kx, ky kz). In fact, the strong dependence of the gravity wave stability on the density profile was a central point of Chandrasekhar's analysis (C61). For the accretion disk cases, we can give an example where stable gravity modes are followed by instabilities (the gravity mode destruction). In case (ib) we have a stable mode with ωh as a QPO frequency. When the density profile changes, over z−coordinate from γ < 0 to γ > 0 then the stable g-mode with ωh can still be sustaned [see case (ia)], But if the density profile over z coordinate is stabilized (β < 0) and that over r coordinate is inverted (i.e. from γ < 0 to γ > 0) then the QPO oscillations are no longer stable. 4. Gravity modes and their relation to QPO phenomenon As we have seen in the previous section, there are two stable oscillatory gravity modes: one is associated with the hybrid frequency ωh (see cases i, ib, ii) and another with the low branch frequency ωL (cases ia and iii). We also estimate the decay rate for oscillations as λ = a1/d and the QPO quality value Q = ω/2λ. The presence and absence of these modes depend on the atmospheric structure (scale height inverses β and γ) and on the imposed boundary conditions (wave numbers kx, ky and kz). Furthermore, it is possible to restore the related boundary conditions if one compares the observed QPO features with the calculated mode frequencies and Q-values. The analysis made in section 3 is also necessary in order to control an accuracy of numerical calculations of the set of hydrodynamical equations (5-9). To illustrate the results obtained in §3 we should specify the orders of the introduced quantities kx, ky, kz, β, γ. Namely we assume that kx ∼ 1/R, ky ∼ 1/R, kz ∼ 1/z, γ ∼ q/R (where q < 1) and z/R < 1. If we also assume that β/γ−kz/kx ∼ R/z then β ∼ 2q/z. With these assumptions expressions for the polynomial coefficients aj (see Eq. 21) are significantly K(zΩx/R + Ωz), a2/a4 ∼ 4(zΩx/R + Ωz)2 + qω2 simplified: a0/a4 ∼ q2ω4 K. K, a1/a4 ∼ −2qω2 -- 11 -- Because q < 1 and ωK < ωh we can calculate ωL as and ωL ≈ q(ωK/ωh)ωK QL = ωL/2(a1/2d) ≈ ωh/2Ω ≥ 1/2. (28) (29) TOK introduced the classification scheme for the QPO features and related the observed high and low kHz frequencies to the hybrid and Keplerian frequencies, νh = ωh/2π and νK = ωK/2π respectively. They also attributed the hectohertz frequencies detected in the atoll source 4U 1728-34 (Ford & van der Klis 1999) to the low branch. The angle δ was found to be almost twice that found in Sco X-1 (see also T02 for details of δ− determination). The low frequencies in atoll sources (as a rule) are three times higher than those in the Z-sources. The hectohertz frequencies have been identified in several other neutron star LMXBs (4U 0614+09; van Straaten et al. 2000, 2002; SAX J1808.4-3658, 4U 1705-44; Wijnands & van der Klis 1998). They weakly depend on the kHz frequencies, their ratio to low kHz frequency being about (4-5) (van Straaten et al. 2002, hereafter S02). The hectohertz frequency Lorentzian profile are very broad with Q-values around 1 or even less. We suggest that these observed frequencies can be identified as the low branch frequencies (see formulas 28-29). Taking into account resonance effects (T02), we correct the error bars of the hectohertz frequencies presented in S02. In Figure 2 the best fit to the data (S02) is presented using formula (28) which includes one fit parameter q. The best-fit value of q = 0.3 for which χ2 ∼ 1/2 are also in agreement with the observed values of Q (S02, red = 0.76. The estimated Q > Table 2). The harmonic modes ωhm = 2(kxΩx + kzΩz) can be related to a coherent oscillation frequency of 582 Hz which is observed in 4U 1636-53 during the superburst (Strohmayer & Markwardt 2002). If kz/kx ≫ 1 (which is our case) ωhm is a double frequency of the NS spin frequency. Furthermore, we have already found the same eigenmode with the frequency 2Ω in case (AII) when effects of the buoynant forces and pressure gradients are neglected (see §3). In fact, Miller (1999) has reported the detection of the coherent pulsations of frequency ∼ 291 Hz during the burst development in 4U 1636-53. Thus the NS spin frequency Ω/2π = 291 Hz and the double NS spin frequency 2Ω/2π = 582 Hz have probably been detected in 4U 1636-53. It is also worth noting that the RXTE observations of neutron binaries establish that the kHz QPO frequencies do not seem to exceed beyond a certain upper limit (Zhang et al. 1998). This observational effect may be a result of the density profile. Our stability analysis presented in §3 clearly indicates such a possibility (see also C61). -- 12 -- 5. Conclusions I have presented a detailed study of the Rayleigh-Taylor (R-T) instability in the accre- tion flow. To summarize I have :(1) put forth arguments to explain the QPO phenomena, as a result of the R-T effect in the rotational frame of reference. (2) formulated and solved the mathematical problem of the gravity wave propagation (R-T effect) in the accretion flow. (3) concluded that the stable gravity modes in the rotational frame of reference are related to the hybrid and low branch frequencies. (4) demonstrated that the particular problem of the gravity waves in the rotational frame of reference, in the approximation of very small pressure gradients, is reduced to the problem of the classical oscillator in the rotational frame of reference which was previously introduced and applied for the interpretation of kHz QPO observation by OT99. I demonstrate that these frequencies are intrinsic features of the R-T effect. They ap- pear in various configurations of the accretion flow depending on assumptions regarding the density profiles, the boundary conditions and the effects of the pressure forces. It is not by chance the high and low frequencies phenomenon has common observational appearances for a wide range of objects classes from black hole sources down to white dwarfs (Mauche 2002). (5) Investigated the conditions for the density profile and the wave numbers (boundary con- ditions) when the gravity modes are stable. (6) Identified the observed QPO frequencies seen in the power density spectra of NS LMXBs using the inferred gravity mode frequencies. In particular, I found that the inferred low branch frequencies and their Q-values are consis- tent with the QPO hectohertz frequencies observed in the atoll sources 4U 1728-34 and 4U 0614+19. (7) During the NS long (super) burst event, I find that the observer should see oscillations at double NS spin frequency. The Coriolis force is the only force which acts in the rotational frame of references and its presence causes perfect coherent pulsations with a frequency twice of the NS spin frequency. Finally one can conclude that the R-T gravity wave oscullations must be present and the related QPOs should be detected in any system where the gravity, buoynancy and Coriolis force effects cannot be excluded (even in the Earth and solar environments). L.T. acknowledges fruitful discussions with Chris Shrader and Kent Wood. I also ap- preciate the fruitful discussions with the referee and his/her constructive evaluation of the manuscript. -- 13 -- REFERENCES Bjerkness, V. Bjerkness, J., Solberberg, H. & Bergron, T. 1933, Physikalish Hydrodynamik, Berlin: Springer Chandrasekhar, S. 1961, Hydrodynamics and Hydromagnetic Stability, Oxford: Oxford at the Caredon Press (C61) Ford, E., & van der Klis, M. 1998, ApJ, 506, L39 Hide, R. 1956, Quart. J. Math. Appl. Math. 9, 35 Kuznetsov, S.I. & Titarchuk, L.G. 2002, ApJ, 571, L137 Mauche, C. W. 2002, ApJ in press (astro-ph/0207508) Miller, M. C. 1999, ApJ, 554, L77 Miller, M.C., Lamb, F.K., Psaltis, D. 1998, ApJ 508, 791 Morgan, E.H., Remillard, R.A., & Greiner, J. 1997, ApJ. 482, 993 Osherovich, V., & Titarchuk, L. 1999, ApJ, 522, L113 (OT99) Psaltis, D., Belloni, & van der Klis, M. 1999, ApJ, 520, 262 Rayleigh, L. 1883, Proc. London Math. Soc. 14, 170 Remillard, R. 2002, astro-ph/0202305 Strohmayer, T.E. & Markwardt, C. B. 2002, ApJ, 577, 337 Strohmayer, T.E. 2001a, ApJ, 552, L49 Strohmayer, T.E. 2001b, ApJ, 554, L169 Strohmayer, T.E., et al. 1996, ApJ, 469, L9 Titarchuk, L.G. 2002, ApJ, 578, L71 (T02) Titarchuk, L., & Wood, K. 2002, ApJ, 577, L23 Titarchuk, L., Bradshaw, C.F. & Wood, K. 2001, ApJ, 560, L55 Titarchuk, L.G., Lapidus, I.I. & Muslimov, A. 1998, ApJ, 499, 315 (TLM98) Titarchuk, L., & Osherovich, V. 2001, ApJ, 555, L55 -- 14 -- Titarchuk, L., Osherovich, V., & Kuznetsov, S.I. 1999, ApJ, 525, L129 (TOK) van der Klis, M. 2000, ARA&A, 38, 717 (VDK) van der Klis, et al. 1996, ApJ, 469, L1 van Straaten, S., Ford, E., van der Klis, M. & Belloni, T. 2002, ApJ, 568, 912 van Straaten, S., Ford, E., van der Klis, M., Mendez, M. & Kaaret, Ph. 2000, ApJ, 540, 1049 Wijnands, R., & van der Klis, M. 1998, ApJ, 507, L63 Woudt, P.A. & Warner, B. 2002, MNRAS, 2002, 333, 411 Zhang, W., Smale, A.P., Strohmayer, T.E., & Swank, J.H. 1998, ApJ, 500, L171 This preprint was prepared with the AAS LATEX macros v5.0. -- 15 -- Fig. 1. -- Behavior of the dispersion polynomial P(n) for stable gravity modes (see text). -- 16 -- Fig. 2. -- Correlation between the lower kHz frequency and hectohertz frequency for 4U 1728-34 (star), 4U 0614+09 (triangles) (van Straaten et al. 2002) and the best-fit curve of low branch frequency vs Keplerian frequency (solid line). χ2 red = 0.76 for q = 0.3.
astro-ph/0408418
2
0408
2005-03-14T18:27:26
Non-parametric inversion of strong lensing systems
[ "astro-ph" ]
We revisit the issue of non-parametric gravitational lens reconstruction and present a new method to obtain the cluster mass distribution using strong lensing data without using any prior information on the underlying mass. The method relies on the decomposition of the lens plane into individual cells. We show how the problem in this approximation can be expressed as a system of linear equations for which a solution can be found. Moreover, we propose to include information about the null space. That is, make use of the pixels where we know there are no arcs above the sky noise. The only prior information is an estimation of the physical size of the sources. No priors on the luminosity of the cluster or shape of the halos are needed thus making the results very robust. In order to test the accuracy and bias of the method we make use of simulated strong lensing data. We find that the method reproduces accurately both the lens mass and source positions and provide error estimates.
astro-ph
astro-ph
Mon. Not. R. Astron. Soc. 000, 1 -- 16 (0000) Printed 16 November 2018 (MN LATEX style file v1.4) Non-parametric inversion of strong lensing systems. J.M. Diego1, P. Protopapas2, H.B. Sandvik3, M. Tegmark4. University of Pennsylvania. 209S, 33rd St, Department of Physics & Astronomy, Philadelphia PA 19104, USA. [email protected] [email protected] [email protected] [email protected] Draft version 16 November 2018 ABSTRACT We revisit the issue of non-parametric gravitational lens reconstruction and present a new method to obtain the cluster mass distribution using strong lensing data without using any prior information on the underlying mass. The method relies on the de- composition of the lens plane into individual cells. We show how the problem in this approximation can be expressed as a system of linear equations for which a solution can be found. Moreover, we propose to include information about the null space. That is, make use of the pixels where we know there are no arcs above the sky noise. The only prior information is an estimation of the physical size of the sources. No priors on the luminosity of the cluster or shape of the halos are needed thus making the results very robust. In order to test the accuracy and bias of the method we make use of simulated strong lensing data. We find that the method reproduces accurately both the lens mass and source positions and provide error estimates. Key words: galaxies:clusters:general; methods:data analysis; dark matter 1 INTRODUCTION Analysis of strong lensing images in galaxy clusters is one of the more subjective and intuitive fields of modern astron- omy (see Blandford & Narayan 1992, Schneider, Ehlers & Falco 1993, Wambsganss 1998, Narayan & Bartelmann 1999, Kneib 2002, for comprehensive reviews on gravitational lens- ing). The common use of parametric models means making educated choices about the cluster mass distribution, for instance that the dark matter haloes follow the luminous matter in the cluster or that galaxy profiles posses certain symmetries. Once these choices are made, one can only hope they are the right ones. The recent improvement in strong lensing data however, is impressive and should increasingly allow us to extract information using fewer assumptions. For instance, it should allow us to relax assumptions about the mass distributions of the cluster and instead enable us to test this. The approach of testing rather than assuming the underlying physics has now been used in other branches of astrophysics and cosmology (e.g. Tegmark 2002) but has so far been largely absent in strong lensing analyses. There are two relatively different fields within the area of strong lensing, namely lensing by galaxies and galaxy clusters. These are different both in appearance as well as in abundance. Galaxy-cluster systems are few and far be- tween, only the most abnormally dense clusters have surface densities greater than the critical density for lensing. How- c(cid:13) 0000 RAS ever, they are much more spectacular than their galaxy lens counterparts. The most massive clusters are able to create multiple images of extended objects such as distant galax- ies with image separations of up to ∼ 1 arcminute. Arguably the most impressive system, A1689 (Broadhurst et al. 2005), boasts more than 100 multiple images of some 30+ sources. Galaxy lens systems are of course on a much smaller scale with image separations of a few arcseconds. They are far more abundant, but less impressive. The lensed objects that we are able to observe are typically high redshift quasars due to their high luminosity and point like structure, al- though galaxy-galaxy lensing has been observed ( Brainerd, Blandford & Smail 1996, Hudson et al. 1998, Guzik & Seljak 2002). The classification of strong lensing systems into these two groups is not merely a matter of scale. The baryons in galaxies have had time to cool and form the visible galaxy, thereby giving a cuspy profile, suitable for lensing. The cool- ing time for galaxy clusters exceed the Hubble time and the density profile is therefore far less cuspy, making them less ideal lenses. Although cluster lens systems are scarce, the im- pressive number of lensed images in each system, mean they still contain a lot of information, particularly regarding clus- ter mass profiles. The information is harder to extract due to the relatively complicated gravitational potentials, but if this challenge can be overcome they could be highly useful probes, certainly of cluster physics and potentially also for 2 Diego et al. cosmology (Yamamoto & Futamase 2001, Yamamoto et al. 2001, Chiba & Takahashi 2002, Golse et al. 2002, Sereno 2002, Meneghetti et al 2005) The intention of this paper is to improve our methods for extracting this information. Although alternative approaches have been suggested to recover the density field in both the weak and strong lensing regime, (see for instance Kaiser & Squires 1993, Broadhurst et al. 1995, Kaiser 1995, Schneider 1995, Schneider & Seitz 1995, Seitz & Schneider 1995, Bartelmann et al. 1996, Taylor et al. 1998, Tyson et al. 1998, Bridle et al. 1998, Marshall et al. 1998), the standard approach to modeling strong cluster lenses is using parametric methods. This is motivated by the fact that the data usually do not contain more than a few arcs. This is not enough to constrain the mass distribution without the help of a parametrization. Parametric methods rely heavily on assumptions or priors on the mass distribu- tion (Kochanek & Blandford 1991, Kneib et al. 1993, 1995, 1996, 2003, Colley et al. 1996, Tyson et al. 1998, Broad- hurst et al. 2000, 2005, Sand et al. 2002). A common prior is the assumption that there is a smooth dark matter com- ponent which is correlated spatially with the centroid of the luminous matter in the cluster. The mass is then ususally modeled by a large smooth dark matter halo placed on top of the central galaxy or the centroid of the luminous matter, as well as smaller dark matter haloes located in the positions of the other luminous galaxies. The parameters of each halo are then adjusted to best reproduce the observations. There is plenty of subjectivity involved in this process, particularly in the addition of the dominant dark matter component to the cluster. The assumption that the dark matter follows the luminosity is necessary but remains the Achilles heel of parametric lens modelling. For large clus- ters the number of parameters in the parametric lens model quickly becomes large but there is still no guarantee that the parametric model used, is in fact capable of reproducing well the mass distribution. It is not hard to envisage complica- tions like dark matter substructure, asymmetric galaxy pro- files, interactions between individual galaxies and the cluster or even dark matter haloes without significant luminosity all of which would not be well represented by the typical para- metric methods. In these cases, where the number of pa- rameters is large, we may want to consider alternative non- parametric methods where all the previous problems do not have any effect on any of the assumptions. Also is in these situations where the number of parameters in both para- metric and non-parametric methods is comparable. When the number of parameters is comparable in both cases, it is interesing to explore non-parametric methods since they do not rely on the same assumptions. This paper does not pretend to be an attack to paramet- ric methods but a defense of non-parametric alternatives. Parametric methods are usually the best way to obtain in- formation about the gravitational potential (few multiple images, simple lens) and are not affeted by resolution prob- lems, specially in the center of the lens which can play a key role in reproducing the radial arcs. They however suffer of some potential problems, some of which have been out- lined above. These problems combined with the impressive quality of recent and upcoming data lead us in this paper to explore the potential to constrain the mass distribution without imposing priors on it. In other words we want to know what strong gravitational lensing images can tell us about cluster mass profiles whilst pretending we know noth- ing about the luminosity. Upcoming images of strong lens- ing in galaxy clusters will contain of the order of a hundred arcs and should make this a manageable task (Diego et al. 2004). A non-parametric approach will provide an impor- tant consistency check, since concurring results would lend strength to the parametric approach, whereas any resulting differences would need to be addressed. Accepting this challenge we present in this paper a new method which makes use of all the available information in super-high quality strong lensing systems. The method has also been thoroughly tested with simulated lensing data with very good results. We thus address the crucial issue of how well we can reproduce both the cluster mass profile and the positions and shapes of the background galaxies. We want to stress that ours is by no means the first work proposing use of non-parametric methods in strong lensing. Among the non-parametric methods already available, are the pixellization methods of Saha et al. (1997, 2000), Ab- delsalam et al. (1998a, 1998b), Williams et al. (2001), who first established many of the ideas revisited in this work, as and also the multi-pole approach of Kochanek & Blandford (1991), Trotter et al. (2000). Naturally, our work shares many similarities with this former work, but there are also important and interesting differences. Firstly, in Saha et al. (1997) (and subsequent papers), the authors divide the lens plane into a grid similar to the method presented here, but with the important dif- ference that their grid is fixed while our grid is dynamical. Our grid adapts to the new estimated mass at each step. This has important implications for the solution since high density regions will be sampled more heavily. These dense regions play a key role, particularly in the positions of the radial arcs. Not sampling these regions properly will neces- sarily lead to a biased mass distribution. Another important difference with Saha et al (1997) is that they make use of a prior on the mass distribution, pe- nalizing deviations from the distribution of luminous matter. They claim this prior does not play an important role but we find this questionable. It is hard to quantify the effect since the method apparently was not tested on simulated lensing data. In contrast, as mentioned above, our method is thoroughly tested with simulations to quantify how well the mass distribution is recovered. We do not use any prior other than a physical prior on the sizes of the sources. This prior is proved to be weak provided it is chosen with a min- imum of wisdom. A third important difference between the work pre- sented in this paper and others is that we show how to speed up the algorithm significantly by adopting techniques com- monly used in optimization problems. Moreover as an added novelty, our algorithm also in- cludes for the first time information about the null space. Rather than using only the information in the lensed arcs, we use information which has hitherto been overlooked; the areas in the sky where no arcs are observed. The paper is laid out as follows. In the next section we present the lens inversion problem, and provide a linear approximation. We then in section 3 present the simulated lensing data used to test the method, before we go on to present the adaptive gridification of the lens-plane, section 4. In section 5 we describe the various inversion algorithms c(cid:13) 0000 RAS, MNRAS 000, 1 -- 16 and finally, in section 6 we introduce the use of the comple- mentary (null) space. 2 THE PROBLEM FORMULATED IN ITS BASIC LINEAR FORM The fundamental problem in lens modelling is the following: Given the positions of lensed images, ~θ, what are the po- sitions of the corresponding background galaxies ~β and the mass distribution of the lens, M (~θ). Mathematically this en- tails inverting the lens equation ~β = ~θ − ~α(~θ, M (~θ)) (1) where ~α(~θ) is the deflection angle created by the lens which depends on the observed positions, ~θ. From now on we will omit the vector notation unless otherwise noted. The data points of our problem, ~θ, are given by the x and y positions of each of the pixels forming the observed arcs. The deflection angle, α, at the position θ, is found by integrating the contributions from the whole mass distribu- tion. α(θ) = 4G c2 Dls DsDl Z M (θ′) θ − θ′ θ − θ′2 dθ′ (2) where Dls, Dl, and Ds are the angular distances from the lens to the source galaxy, the distance from the observer to the lens and the distance from the observer to the source galaxy respectively. In equation 2 we have made the usual thin lens approximation so the mass M (θ′) is the projected mass along the line of sight θ′. Due to the (non-linear) de- pendency of the deflection angle, α on the position in the sky, θ, this problem is usually regarded as a typical exam- ple of a non-linear problem. We will see that this is only partially true. The next approximation we make is to split the lens plane in Nc small regions (hereafter cells) over which the projected mass is more or less constant. We can then rewrite equation (2) as; α(θ) = 4G c2 Dls DsDl XNc mi θ − θi θ − θi2 (3) The first point we want to make here is that the deflection angle α may be thought of as the net contribution of many small masses mi in the positions θi, each one pulling the deflection in the direction of (θ − θi) and with a magnitude which is proportional to mi/(θ − θi). If we divide the lens plane in a grid with Nc cells, the masses mi can be consid- ered as the mass contained in the cell i (i = 1, ..., Nc). If the cells are sufficiently small then the above pixellization of the mass plane will give a good approximation to the real mass distribution. Our second point is that the problem is non-linear in one direction but linear in the other. That is, given a position in the sky θ (and given a lens) there is only one β which satisfy equation 1 but given a position of the background galaxy β, there may be more than one position in the sky (θ) satisfying equation 1 or equivalently, the source galaxy may appear lensed in more than one position in the sky. The linear nature of the problem is evident when one realizes that the only non-linear variable, namely the θ positions, are fixed by the observation and that the problem depends c(cid:13) 0000 RAS, MNRAS 000, 1 -- 16 Inverting the lens 3 Figure 1. Original simulated mass profile. The total projected mass is 1.119 × 1015 h−1 M⊙ in the field of view (0.1 degrees across) and the cluster is at z = 0.18. The units of the grey scale map are 1015h−1M⊙/pixel where each pixel is 0.494 arcsec2. Also shown are the radial (small) and tangential (large) critical curves for a source at redshift z = 3. linearly on the unknowns (β positions and masses, mi, in the cells). Let us now assume that we have a data set consist- ing of a system of radial and tangential strong lensing arcs which are spread over Nθ pixels in the image. We will also assume that we know which arcs originate from the same sources. Since both the data and the mass distribution has been discretized we can rewrite equation (1) as a simple matrix equation: β = θ − ΥM (4) where θ and β are now 2Nθ element vectors containing x and y values of the observed positions and the (unknown) source positions respectively. M is the mass vector containing all Nc mass cells, and Υ is the (2Nθ × Nc) matrix casting the mass vector into a vector of displacement angles. A more technical account of the make-up of the Υ matrix can be found in the appendix. Equation 4 clearly demonstrates the linear nature of the problem when formulated in this manner. The problem has now been reduced to a set of 2Nθ linear equations with 2Nθ + Nc unknowns (lens masses and source galaxy positions). Notice that when the problem is formulated in this form, there are more unknowns than equa- tions which means it is an underdetermined system with an infinite number of solutions. In order to identify a suitable solution for such a system, we need to add extra information or impose constraints. One way of doing this is by reducing the number of un- knowns, for instance by removing the source positions from the unknown category. This can be achieved by minimizing the dispersion in the source plane, i.e. demanding that the pixels in the source plane be as concentrated as possible for each source. In this case we are left with only Nc unknowns. 4 Diego et al. allows for a complete, linear solution to a problem usually considered non-linear. 3 SIMULATIONS Before proceeding to invert the system of linear equations (5), it is instructive to take a closer look at the simulations which are going to be used to test the inversion algorithms. Given that most methods rely on the luminous mass dis- tribution, this is particularly important should the galaxies not accurately trace the dark matter. Our simulations consist of three elements. The first el- ement of the simulation is the projected mass distribution (M ) in the lens plane. We simulate a generic mass distribu- tion of a cluster with a total projected mass of 1.119 × 1015 h−1 M⊙ in the field of view located at redshift z = 0.18. The field of view of our simulation is 0.1 degrees The mass profile is built from a superposition of 20 NFW profiles with added ellipticity. The halos' masses vary from 0.25 × 1015 h−1 M⊙ to 2 × 1012 h−1 M⊙. In the context of lensing, NFW halos seem to reproduce well the shear profile of massive clus- ters up to several Mpc (Dahle et al. 2003, Kneib et al. 2003, Broadhurst et al. 2005). The final mass distribution is shown in figure 1. Also in the same figure we show the two critical curves. The interior one is the radial critical curve and the exterior one the tangential critical curve. Both curves have been calculated assuming the source is at redshift z = 3. The tangential critical curve is usually associated with the Einstein radius. Its large size (radius ≈ 2 arcmin) is due to the unusually high projected mass of our cluster plus its rel- atively low redshift (z = 0.18). Although the deduced Ein- stein radius is larger than the one observed in clusters, the simulation presented here will serve the purpose of testing the different methods discussed here. More realistic simula- tions with proper mass distribution and source density will be presented in a future paper. The second element in our simulation are the sources (β), for which we extract 13 sources from the HUDF (Hubble Ultra Deep Field, Beckwith et al. 2003). We assign them a redshift (z ∈ [1, 6]) and size and place them in different positions behind the cluster plane. The third element are the lensed images (θ) which are calculated from the first two elements (M and β). This is done through a simple ray-tracing procedure. For each posi- tion θ in the image, we calculate the deflection angle, α, and then the corresponding source plane position,β, according to the lens equation (equation 1). If the calculated β coincides with one of the original sources, we assign to the lensed im- age the value (colour) of that source. Otherwise that point in the lensed image is left dark (value 0 ). We repeat the op- eration for all the pixels (θs) to produce a complete image. Also, since the sources are small compared with our pixel size and to avoid missing some sources, we oversample our θ pixels by subdividing them and checking each pixel at dif- ferent locations (θ + ∆θ with ∆θ < 1 and θ ∈ [1, 512]) The original sources are plotted in figure 2 and the corresponding θs in the left panel of figure 3. c(cid:13) 0000 RAS, MNRAS 000, 1 -- 16 Figure 2. The 13 sources at redshift z ∈ (1, 6). The field of view is 0.1 degrees across and is centered in the same point as figure 1. The solid line marcs the position of the caustics at z = 3 correponding to the radial (dotted line) and tangential (solid line) critical curves. Another way of constraining the system is by assum- ing the Ns sources are well approximated by point sources, which reduces the number of unknowns to Nc + 2Ns. This means effectively demanding that all observed θs for arcs corresponding to the same source, can be traced back through the lens to Ns single points. With this assumption we can rewrite the lens equation in the compact form of θ = ΓX. (5) Γ is now a matrix of dimension 2Nθ × (Nc + 2Ns) and X is the vector of dimension (Nc + 2Ns) containing all the unknowns in our problem (see appendix A), the mass ele- ments and the 2Ns central (x- and y) coordinates of the Ns sources. Now the system is matematically overdetermined (more data points that unknowns) and has a unique point source solution. This unique solution can be found by nu- merical methods as we will see later. The linearization of the problem means that it is in principle solvable by both matrix inversion and simple lin- ear programming routines. In practice, the problem quickly becomes ill-conditioned and too large for direct matrix in- version, and approximate numerical methods are more suit- able. The main problem with the linearization is that we do not know if the obtained linearized solution creates artifi- cial tangential or radial arcs. Checking this requires forward solving of the lens equation which is non-linear due to the complicated dependence of the deflection angle on θ. We suggest a novel approach to this problem, by using all the available information in the images, i.e. the informa- tion inherent in the dark areas (pixels containing no arcs) as well as the observed arcs. By pixellization the dark areas, tracing these pixels back through the lens and imposing that they fall outside the sources it is possible to find the true solution without over-predicting arcs. This use of the null space is to our knowledge unprecedented, and in principle Inverting the lens 5 Figure 3. The sources lensed by the mass distribution of figure 1. The right image is the lensed image using 325 cells in the dynamical grid (see text). The left image is the exact solution using no griding. Note the differences in the radial arcs. 4 GRIDIFYING THE MASS DISTRIBUTION 4.1 The Υ matrix As laid out in section 2 the basis of our non-parametric reconstruction method is the assumption that the real mass can be well approximated by a pixelized mass distribution. The Υ matrix is then the matrix which casts the mass vector into the vector of deflection angles αi = ΥijMj (6) For convenience rather than taking the mass distribu- tion in each cell to be constant we assume it follows a Gaus- sian distribution centred in the centre of the cell with some dispersion. This allows us to calculate analytically the lens- ing contribution from each mass cell, saving valuable com- puter time. We use a dispersion of 2a where a is the size of the cell, and we have confirmed with simulations that the lensed image generated with this choice agrees well with the true lensed image using a relatively small number of cells. The detailed structure of the Υ matrix will be explained in appendix A. 4.2 Multi resolution mass-grid Rather than taking a uniform grid, it is better to construct a dynamical- or multi-resolution grid. By sampling dense re- gions more heavily, it is possible to reduce drastically the number of cells needed to accurately reproduce the lens- ing properties of the cluster. In other words we choose an adaptive grid which samples the dense cluster centre better than the outer regions. Since we do not actually know the density profile of the cluster, this multi-resolution grid must be obtained through an iterative procedure. An example of c(cid:13) 0000 RAS, MNRAS 000, 1 -- 16 this process can be found on the SLAP⋆ (Strong Lensing Analysis Package) webpage. Given a mass estimate (a first mass estimate can be ob- tained with a coarse regular grid), we split a given cell into four sub-cells if the mass in the cell exceeds some threshold value. The lower this threshold, the higher the number of divisions and consequently the higher the final number of cells. The obtained grid can then be used for the next mass estimate, and the process can be repeated as necessary. Typ- ically the mass estimate will improve with each iterative step as this dynamical grid allows for the relevant regions of the cluster to become resolved. In figure 4 we show an example of a gridded version of the true mass in figure 1 for a threshold of Mthr = 8.0×1012 h−1 M⊙. This grid has 325 cells. The corresponding (true) mass in the grid is shown in figure 5. The parameter Mthr or equivalently the number of cells, Nc, can be viewed as a free parameter but also as a prior. Fixing it means we are fixing the minimum scale or mass we are sensitive too. In a paralell paper (Diego et al. 2004) we explored the role of Nc and found that it can have a negative effect on the results if it takes very large values. A natural upper limit for Nc is 2 times the number of pixels in the θ vector minus 2 times the number of sources (the factor 2 accounts for the x and y positions). Our simulations show that a good election for Nc is normally 1/4 of this upper limit. For the shake of clarity it is useful to give some practical numbers. The image on the left of figure 3 has 5122 pixels from which 2156 pixels are part of one of the ≈ 35 arcs so Nθ = 2156. These arcs are coming from 13 sources (Ns = 13) and the lens plane is tipically divided in a few hundred cells (Nc ∼ 300). ⋆ see http://darwin.cfa.harvard.edu/SLAP/ 6 Diego et al. Figure 4. Dynamical grid for the mass in figure 1. This case corresponds to Mthr = 8.0 × 1012 h−1 M⊙. Figure 5. Masses in the cells of the dynamical grid of figure 4. 5 INVERSION METHODS In this section we will describe some inversion methods which can be applied to solve the problem. Most of these methods can be found in popular books like Numerical Recipes (Press et al. 1997). All these algorithms are be- ing implemented in the package SLAP⋆ which will be made available soon. Once we have the problem formulated in its linear form with all the unknowns on one side it is tempting to try a direct inversion of equation 5. Although the Γ matrix is not square, one can find its inverse, Γ−1, by decomposing Γ into orthogonal matrices. This is similar to finding the eigenval- ues and eigenvectors of Γ. This approximation has its ad- vantages as well as its drawbacks and we will explore this possibility later. However, we anticipate that degeneracies between neighboring pixels in the arcs as well as neighboring cells in the lens plane (not to mention the compact nature of the sources) will result in a system of linear equations which is not well behaved. The rank of the matrix Γ will be normally smaller than its smaller dimension. Calculating the inverse in this situation is not a trivial task. A second approach is rotating our system of linear equa- tions using a transformation which is just ΓT . This trans- forms Γ into a square, symmetric and positive definite ma- trix of dimension (2Ns + Nc) × (2Ns + Nc), A = ΓT Γ which is better behaved than the original Γ matrix. However, the rank of A is in generally smaller than its dimension and its inverse does not exist. The hope in this case must therefore be to find an approximate rather than exact solution to the system. The third approach is the simplest and will be explored first. We assume we know nothing about the sources other than their redshift and that they are much smaller than the strong lensing arcs. This is the same as saying that the lens has to be such that it focuses the arcs into compact sources at the desired (known) redshift. This simple argument alone will turn out to be powerful enough to get a quick but good first estimate of the mass distribution in the lens plane. We explore all these approaches below in reverse order. 5.1 A first approach: Minimizing dispersion in the source plane In this subsection we will discuss the simplest (although ef- fective) method to get a fast estimation of the mass using no prior information on the lens nor the sources. The prob- lem then contains two sets of unknowns: The mass vector we want to determine, M , and the β positions of the sources. In general, for a finite number of observed arcs, there are sev- eral combinations of β and M which can reproduce the ob- servations. The most obvious unphysical solution is the null solution, where the mass is zero and the sources identical to the observed arcs. The easiest way to avoid such unphysical solutions is to minimize the variance of the β positions. This is equivalent to imposing that the vector M really acts as a true lens: We require big arcs with large magnifications and multiple images separated by arcsec or even arc-min to focus into a rather compact region in the source plane. This min- imization process assumes that we are able to associate the multiple lensed images with a particular source. This can be achieved with either spectroscopy or with morphology and the multicolor imaging of the arcs. To minimize the variance in the source plane it is illus- trative to follow the steepest descent path although other more effective minimization algorithms can be used (see be- low). Given an initial guess for the mass vector, one can calculate the derivative of the variance in the source plane as a function of the mass and minimize in the direction of the derivative. Once a minimum is found, we calculate the derivative in the new mass position and minimize again in an iterative procedure. The quantity to be minimized is : f (M ) = Xs σ2 s (7) c(cid:13) 0000 RAS, MNRAS 000, 1 -- 16 Inverting the lens 7 Figure 7. Recovered sources after minimizing the variance. The real positions of the sources are shown as crosses. Note that the recovered sources are farther away than the real ones. This com- pensates for the fact that the recovered mass is lower than the the true value. the sources. In fact, since the source positions can be esti- mated from equation 4, the minimization of the variance also provides us with an initial guess for these. The drawback is the slow convergence of the algorithm. A typical minimiza- tion may take several hours on a 1 GHz processor. In the next sub-section we will go a step further and we will include the β positions in the minimization as well as speed up the convergence by orders of magnitude. 5.2 Biconjugate Gradient Inversion of linear systems where the matrix dimensions are of order 103, is a numerically trivial problem for today's computers provided the matrix is well behaved. If the ma- trix has null or negative eigenvalues, a direct inversion is not feasible and one has to aim to solve for some approximated solution. Our system of linear equations is a good example of an ill-conditioned one. Direct inversion of the matrix is not possible due to negative eigenvalues. However there is another important reason why we do not want to solve ex- actly (or invert) the system of equations. An exact solution means that we will recover a mass distribution which puts the arcs into delta function sources. As we will see later, this solution will be unphysical. Instead, we are interested in an approximated solution which does not solve exactly the sys- tem of equations and which has a residual. This residual will have the physical meaning of the extension of the sources or the difference between the point-like sources and the real, extended ones. The biconjugate gradient (Press et al. 1997) will be a useful way to regularize our problem. The biconjugate gradient algorithm is one of the fastest and most powerful algorithms to solve for systems of linear equations. It is also extremely useful for finding approximate solutions for systems where no exact solutions exist or where Figure 6. Smooth version of the recovered mass after minimizing the variance. The total recovered mass is 1.01 × 1015 h−1 M⊙. Compare this mass with the original one in figure 1. where the sum is over the number of identified sources and σ2 s is the variance of the source s in the source plane. That is; σ2 s =< β2 >s − < β >2 (8) s where the β's are calculated from equation 4 and the aver- age is over the β's corresponding to source s. By combining equations 4 and 8 is easy to compute the derivative of σ2 s with respect to M . ∂σ2 s ∂Mj = 2 < β >s< Υj >s −2 < βΥj >s (9) x + σ2 where Υj is the column j of the Υ matrix and the average is made only on the elements associated with the source s. We should note that all equations involving the vectors β, θ or α = ΥM have two components, x and y so there will be in fact two equations like equation 9. One for the x component of β and the other one for the y component. At the end, the quantity we aim to minimize is σ2 = σ2 y. As we already mentioned, the minimization can be done following the path of steepest descent given by equation 9. This path will end in a minimum at the new mass M j . The process can be repeated by evaluating the new path at the new mass position until the variance is smaller than certain ǫ. A good choice for ǫ is to take a few times the expected variance for a population of Ns galaxies at the measured Ns redshifts. specific values for ǫ will be discussed later. In practical terms the minimization is done through a series of iterations, gradually improving the dynamical grid as laid out in section 4.2. For each iteration the ability of the mass distribution to focus the βs into compact sources is improved. Equation 8 can be improved by weighting it with the amplification of the different images, assuming the errors in the image plane are similar from images to images. This point will not be explored here. The minimization of the variance is a powerful and robust method for finding a first guess for the mass vector without making assumptions about c(cid:13) 0000 RAS, MNRAS 000, 1 -- 16 8 Diego et al. the exact solution is not the one we are interested in. The latter will be our case. Given a system of linear equations; Ax = b (10) a solution of this system can be found by minimizing the following function, f (x) = c − bx + 1 2 xtAx (11) where c is a constant. When the function f (x) is minimized, its gradient is zero. ∇f (x) = Ax − b = 0 (12) That is, at the position of the minimum of the function f (x) we find the solution of equation (10). In most cases, finding the minimum of equation 11 is much easier than finding the solution of the system in 10 especially when no exact solution exists for 10 or A does not have an inverse. The biconjugate gradient finds the minimum of equa- tion 11 (or equivalently, the solution of equation 10) by fol- lowing an iterative process which minimizes the function f (x) in a series of steps no longer than the dimension of the problem. The beauty of the algorithm is that the successive minimizations are carried out on a series of orthogonal con- jugate directions , pk, with respect to the metric A. That is, piApj = 0 j < i (13) This condition is useful when minimizing in a multidimen- sional space since it guarantees that successive minimiza- tions do not spoil the minimizations in previous steps. Let us now turn to the system we want to solve, namely equation 5. The biconjugate gradient method assumes that the matrix Γ ( matrix A in equation 10) is square. For our case this does not hold since we typically have Nθ >> (Nc + Ns). Instead we build a new quantity, called the square of the residual,R2: Figure 8. Recovered mass after minimizing R2 using the bicon- jugate gradient algorithm. The mass has been smoothed with a Gaussian filter. The total recovered mass is 1.17 × 1015 h−1 M⊙. R2 = (θ − ΓX)T (θ − ΓX) 1 2 θT θ − ΓT θX + = 2( 1 2 X T ΓT ΓX) (14) (15) By comparing equations 15 and 11 is easy to identify the 2 θT θ, b = ΓT θ and A = ΓT Γ. Minimizing the terms, c = 1 quantity R2 is equivalent to solving equation 5. To see this we only have to realize that b − AX = ΓT (θ − ΓX) = ΓT R (16) If an exact solution for equation 5 does not exist, the min- imum of R2 will give the better approximated solution to the system. The minimum can be now found easily with the biconjugate gradient (Press et al. 1997). For the case of sym- metric matrices A, the algorithm constructs two sequences of vectors rk and pk and two constants, αk and βk; αk = rT k rk pT k Apk rk+1 = rk − αkApk βk = rT k+1rk+1 rT k rk pk+1 = rk+1 + βkpk (17) (18) (19) (20) Figure 9. Recovered β's after minimizing R2. Again, crosses rep- resent the true position of the sources. The recovered βo falls in the middle of its corresponding cloud β points. At every iteration, an improved estimate of the solution is found by; Xk+1 = Xk + αkpk (21) The algorithm starts with an initial guess for the solution, X1, and chooses the residual and search direction in the first iteration to be; r1 = p1 = b − AX1 (22) Note that p1 is nothing but ∇R2. Thus the algorithm chooses as a first minimization direction the gradient of the c(cid:13) 0000 RAS, MNRAS 000, 1 -- 16 function to be minimized at the position of the first guess. Then it minimizes in directions which are conjugate to the previous ones until it reaches the minimum or the square of the residual R2 is smaller than certain value ǫ. The method has one potential pathological behavior when applied to our problem. One can not choose ǫ to be arbitrarily small. If one chooses a very small ǫ the algorithm will try to find a solution which focuses the arcs in Ns sources which are delta functions. This is not surprising as we are assuming that all the 2Nθ unknown βs are reduced to just 2Ns βs, i.e the point source solution (see figures 10 and 11 ) . The mass distribution which accomplishes this is usu- ally very much biased compared to the right one. It shows a lot of substructure and it has large fluctuations in the lens plane. One therefore has to choose ǫ with a wise criteria. Since the algorithm will stop when R2 < ǫ we should choose ǫ to be an estimate of the expected dispersion of the sources at the specified redshifts. This is the only prior which has to be given to the method. However, we will see later how the specific value of ǫ is not very critical as long as it is within a factor of a few of the right source dispersion (see figure 18 below). Instead of defining ǫ in terms of R2 is better to define it in terms of the residual of the conjugate gradi- ent algorithm, r2 k. This will speed the minimization process significantly since we do not need to calculate the real dis- persion at each step but to use the already estimated rk. Both residuals are connected by the relation, rk = ΓT R (23) Imposing a prior on the sizes of the sources means that we expect the residual of the lens equation, R, to take typical values of the order of the expected dispersion of the sources at the measured redshifts. Hence we can define an Rprior of the form; Inverting the lens 9 focuses the arcs into sources which are larger than the real sources. We should take this into account when fixing σi. The reader can argue that a more clever way of includ- ing this prior information in the algorithm is by perturbing the β elements in equation 5 (or similarly, equation 5 in Ap- pendix A). This is done by adding some noise to the 1s in the two 1 matrices in equation 37 (see Appendix A). One could for instance add Gaussian noise with a dispersion similar to the expected dispersion of the source at redshift z. The reality however is that the quadratic nature of R2 cancels out any symmetric perturbation added to the elements of Γ. Thus, the result is similar if we perturb Γ or not and we still have to include the prior and fix ǫ to be large enough so we do not recover the point source solution. This also tell us that this method is not very promising if one wants to include parity information in the recovery of the mass and sources. In the next subsection we will show a different approach which can include this parity information. 5.3 Singular Value Decomposition The Singular Value Decomposition (hereafter SVD) algo- rithm allows for decomposition of a generic m × n (with m >= n) matrix A into the product of 3 matrices, two or- thogonal and one diagonal (e.g Press et al. 1997). A = U W V T (26) where U is an m × n orthogonal matrix, W is an n × n diag- onal matrix whose elements in the diagonal are the singular values and V T is the transpose of an n × n orthogonal ma- trix. When A is symmetric, the SVD reduces to finding the eigenvectors and eigenvalues of A. The advantage of this decomposition is that the inverse of A is given by ; Ri prior = σi ∗ RN D (24) A−1 = V W −1U T (27) where the index i runs from 1 to Nθ and σi is the disper- sion (prior) assumed for the source associated to pixel i and RN D is a random number uniformly distributed over -1 and 1. Then, we can estimate ǫ as; ǫ = rT k rk = RT priorΓΓT Rprior (25) When calculating ǫ is is better to consider only the first Nc columns of Γ and discard the last 2Ns. This is recommended to avoid that the 1s in this part of the Γ matrix do not cancel out when multiplied by the random number and dominate the much smaller Γij elements corresponding to the mass components. The last Nc columns of Γ should give 0 con- tribution when multiplied by the random Ri prior vector. If we chose as a prior that the sources are Gaussians with a σ = 30h−1 kpc located at the measured redshift, this ren- ders ǫ ≈ 2 × 10−10. The reader will note that the chosen σ is a few times larger than the one we would assign to a typical galaxy. We will discuss this point later. The final result is shown in figures 8 and 9. One has to be careful in not choosing the σi very small. In fact, they should be larger than the real dispersion of the source. Only when the number of grid points, Nc, is large enough, can the gridded version of the true mass focus the arcs into sources which are similar in size to the real ones. If Nc is not large enough, the gridded version of the true mass c(cid:13) 0000 RAS, MNRAS 000, 1 -- 16 where W −1 is another diagonal matrix whose elements are just the inverse of the elements of W , that is W −1 jj = 1.0/Wjj . The proof A−1A = I follows from the property that U and V T are orthogonal (U T U = V V T = I). In our case, we can use SVD to calculate the inverse of the Γ matrix and find the solution, X, directly by inverting equation (5) X = Γ−1θ = V W −1U T θ (28) Although the SVD allows us to invert the problem by calcu- lating Γ−1, its full power lies in its ability to solve a system approximately. The level of approximation can be controlled by setting a threshold in the matrix W −1. In our problem, there will be many equations which are strongly correlated in the sense that most of the θ positions in a single arc will come from the same source (that is, they will have almost the same β). Also we have to keep in mind that we are us- ing all the pixels in our data. This means that two equations corresponding to two neighboring pixels will look almost ex- actly the same. When one computes the SVD of the matrix Γ, these two facts translate into a matrix W with elements in the diagonal which are 0 or close to 0. The inverse of W would be dominated by these small numbers and the solu- tion will look very noisy. The good news about using SVD is that the most relevant information in the Γ matrix is packed 10 Diego et al. Figure 10. Recovered mass after minimizing R2 using the bicon- jugate gradient algorithm and with a very small ǫ (point source solution). The total recovered mass is 2.43 × 1015 h−1 M⊙ but there are also regions with negative masses. Figure 11. Recovered β's after minimizing R2 for the point source solution. The real source positions are shown as crosses. into the first few values (the largest values in W ) while the small single values in W will contain little information or the redundant information coming from neighboring pixels. One can just approximate W by another matrix W ′ where all the elements in its diagonal smaller than certain thresh- old are set to 0. Also, in the inverse of W ′, these elements are set to 0. The magic of this trick is that the solution found with this approximation will contain the main trend or main components of the mass distribution. Figure 12. Recovered mass after SVD (no prior). The total re- covered mass is 1.01 × 1015 h−1 M⊙. The degree of accuracy is controlled by setting the thresh- old in the singular values of the matrix W . Those elements in W below the threshold are set to 0 and the same in its inverse. The threshold is usually set after looking at the sin- gular values. The first ones will normally stay in some kind of plateau and after then the singular values will decrease rapidly. The threshold should be normally set immediately after the plateau. In figures 12 and 13 we show the result af- ter decomposing Γ in its SVD decomposition and calculating its inverse with (27). Like the two previous algorithms, the SVD has its own pathologies. Using standard subroutines to find the SVD of Γ usually return a no convergence error. This error comes from the nearly degenerate nature of Γ. One has then to in- crease the number of maximum iterations on these subrou- tines or use a coarse version of Γ where only a small fraction of the θ positions (or equivalently, Γ rows) are considered. Another solution is inverting the preconditioned system of equations where we previously multiply for ΓT . This allows to use all the θ pixels and find the SVD of ΓT Γ in a small number of iterations. However, using the SVD of Γ instead of ΓT Γ has a very interesting feature. It allows to introduce parity information in an effective way. Contrary to the quadratic cases of minimization of the variance or the square of the residual R2, the SVD allows us to include parity information in the Γ matrix which will not disappear when we look for the best solution. Since no ΓT θ or ΓT Γ operations are involved, parity information will not cancel out. SVD could be an interesting way of fine tuning the solution by including the extra information coming from the parity of the arcs. 6 INCORPORATING THE NULL SPACE Another advantage of using the SVD algorithm is that in this case no prior on the extension of the sources is needed. So far we have made use of the information contained in the observed strong lensing arcs. This gives us a solution which c(cid:13) 0000 RAS, MNRAS 000, 1 -- 16 Inverting the lens 11 predicting β falling near the β positions which minimize R2. As a penalizing function we will choose an asymptotically divergent Gaussian g(fk) = 1 efk − µ with, fk = ( β(k)x − βo x)2 + ( β(k)y − βo y)2 σ2 k (31) (32) Where β(k)x is the x component of β for the pixel k in the null space. βo x is our estimated value of β (x component) and similarly for the y component. There are as many constraints of the form fk as there are pixels, θ, in the null space. The parameter µ controls the degree of divergence of the Gaussian function. When µ = 0 we recover the classi- cal Gaussian but as µ approaches 1 the Gaussian becomes more and more sharply without increasing its dispersion. For µ = 1 the Gaussian is infinite at fk = 0 (see figure 14. By minimizing the function φ(X) with increasing values of µ we will find that in the limit µ → 1 the solution will push away those β which originally were falling in the region defined by the set of β. Ideally we want to include all the dark or empty pixels in the θ space but this is, in general, found to be a waste of memory and computing time. The fact is that only the θ pixels which are hitting (or close to hit) one of the Ns esti- mated β positions of the sources will have some impact on the solution. Most of the θ pixels in the null space already satisfy all the constraints for the actual solution X. For this reason we will include only those θ for which the solution X predicts their corresponding β are close to hitting (or actu- ally hitting) a source. The θ space will include the observed θ's as a subspace. We have to exclude this subspace from β before minimizing equation 30. This is, again, just an opti- mization process. In practice, one can include all the pixels in the image (excluding those containing part of an arc) in the null space. After the minimization process is finished, the new so- lution will have the arcs falling in compact regions around β0 while the extra-arcs produced by the previous solution will fall in regions outside areas around the βo. ¿From our simulation we have seen that addition of the null space induces small changes in the mass plane and it tends to stabilize the solution in the sense that it makes the recovered mass profile independent of the threshold ǫ. This is in fact an interesting bonus which comes with the addition of the null space. The new function to be minimized, φ, can be minimized until the true minimum is found. In this case, there is no equivalent of a point source solution. Since some of the θ are in fact very close to the observed θ, the solution which minimizes φ will focus those θ and θ in neighboring regions in the source plane. When minimizing φ there will be two competing effects. One will tend to increase the mass so it minimizes R2 (point source solution). The other will tend to reduce the mass so the β will be pushed away from the β positions. The outcome will be a balanced situation between the β trying to collapse into compact sources and the β trying to escape the wells in the beta positions. To accurately quantify the effect more simulations are needed. This will be done in a future paper but as an illustra- tion we show in figure 15 the recovered mass after imposing Figure 13. Recovered sources after SVD (no prior). The real positions of the sources are shown as crosses explains the data in the sense that it predicts arcs in the right positions but it could well happen that the solution over-predicts arcs. An example of this can bee seen by com- paring the reproduced arcs in fig. (16) with the true arcs in fig. (3). Some of the reproduced arcs are slightly larger and span a larger area than the true ones, although the sources are perfectly reproduced in their true positions and the re- covered mass is very close to the true mass distribution. To avoid this we propose for the first time to include information from the null space, θ. That is, the part of the image which does not contain any arc. This space tell us where an arc should not appear but it does not tell us where the hypothetical β should be. The only fair thing we can do is to impose that none of the pixels in the null space fall into the estimated β of our solution. By achieving this we will have a solution which predicts the right arcs while not over- predicting arcs. The solution will be then fully consistent with the observed data. The null space is connected to the solution X by; β = θ − ΥM (29) It is evident that we want the new solution X (M and βo) to be such that the new β do not fall within a circle of radius p(k) centered in each one of the βo where p(k) is the prior with information on how extended are the sources. The null space will perturb the solution X in such a way that the new solution, X ′ = X + ∆ X is an approximated solution of equation 5 and satisfies all the constraints of the form β ∋ β. The way we incorporate the constraints is by adopt- ing an approach commonly used in penalized quadratic pro- gramming. We will minimize the new function φ(X) = R2 + λXk g(fk) (30) where the λ is a constant which guarantees that in the first iterations, the second term in equation 30 does not dominate the first and gk is a function which will penalize those models c(cid:13) 0000 RAS, MNRAS 000, 1 -- 16 12 Diego et al. Figure 14. Penalizing function showing values of µ = 0.8, 0.9, 0.95, 0.97, and 0.98 and σ = 1. Note that the width of the curve does not change when increasing µ. Figure 15. Recovered mass (smoothed) after minimizing φ. The crosses show the position of the three main halos in the cluster. The size of the cross is proportional to the mass of the halo. The total recovered mass is 1.1055 × 1015h−1M⊙. the constraints in the θ space. The total mass is now only 1.2 % lower than the true mass. The new mass also contains more structure and starts showing the internal distribution of the main components of the cluster. In figure 16 we show the predicted position for the arcs after combining the best mass together with the best esti- mate for the position of the sources. To compute the arcs we have assumed that each source is a circle with a radius 15h−1kpc centred in the estimated best source positions and at the measured redshifts. By comparing figures 16 and fig- ure 3 (left) we see that the predicted arcs matches very well the observed (simulated) data with the exception of some of the arcs near the center of the image. Figure 16. The plot shows the predicted arcs according to the best mass and source solution. We have assumed that the sources are circles with radius 15h−1kpc centred in the source position and at the measured redshifts. This result should be compared with the true arcs seen in figure 3. The match is almost perfect. 7 DISCUSSION In this paper we have presented several approaches to non- parametric lens modelling. Using no information about lu- minosity distributions whatsoever, these methods perform remarkably well on simulated strong lensing data. One of the main conclusions of this work is simply that it works; It is possible to recover information about the mass distribu- tion without using prior information on the same if one has a sufficiently large number of arcs. This can be seen in fig- ure 17 where we show the recovered mass profiles compared with the original one. The recovered mass traces the real mass distribution up to furthest data point in the simulated image. Beyond this point, the recovered profile is insensitive to the data. When the minimization stops, the cells in the outer regions stay with a mass close to the one they had in the first step of the minimization. This point is discussed in more detail in Diego et al. (2004). We are optimistic about the performance of these meth- ods when used on future data, but we would also like to emphasize the potential pathologies. We will now discuss the major issues. As we have seen in section (2) the inver- sion algorithms rely on a linearization of the lens equation. We achieve that by decomposing the lens plane into small mass-cells and assuming that the gridded mass is a fair rep- resentation of the true underlying mass. This is true when the number of cells in the grid is large enough, but for a uni- form grid this may mean using several thousand mass-cells, making the problem ill conditioned and underdetermined. By inverting the lens equation in a series of iterations we introduce an adaptive grid which optimizes the number of cells by sampling dense regions more heavily and using larger cells for underdense regions. This allows for good sampling of the lens without a huge number of cells. However, the grid- ded mass plane is still an approximation to the true mass c(cid:13) 0000 RAS, MNRAS 000, 1 -- 16 Inverting the lens 13 Introducing such a prior begs the question: How sensi- tive is the solution to the specific guess of prior? The answer is that it depends on exactly "how bad" that choice is. As long as we assume a source size significantly larger than the true sizes the actual solution does not change much and re- sembles well the true mass distribution. However, when try- ing to approach the true source size, the solutions changes rapidly away from the realistic model. The situation is shown in fig. (18). As shown in section 5.2 a given physical size cor- responds to a given threshold, ǫ, and as long as this threshold is sufficiently large, ≥ 10−10, (corresponding to a physical size of ≥ 30h−1 kpc) the total mass is well behaved. If we instead demand that the physical size of the reconstructed sources be more realistic, say ∼ 12h−1 kpc, we get a thresh- old of ∼ 10−11 for which the mass distribution is already starting to diverge away from the true mass. We therefore want the recovered sources to be larger than the true ones. In other words we want to recover a short-sighted cluster! In previous sections we have demonstrated non-parametric lens modelling using several different algorithms with promising results. However we have not yet addressed the question of uniqueness. Is there a unique solution, and if not, how dif- ferent are the possible solutions? The good news is that minimizing quadratic functions like the variance or the square of the residual guarantees that there is only one absolute minimum. This absolute minimum corresponds to the point source solution. The bad news is that this is not the solution we are looking for. As shown above, trying to focus the sources too much introduces arti- facts in the mass distribution, so we need to stop the mini- mization at some step before the absolute minimum. In two dimensions it is easy to visualize that the quadratic func- tion will have the shape of a valley, and stopping the min- imization at some point before the actual minimum means choosing an ellipse on which all solutions are equally good. In many dimensions this is harder to visualize but we expect our obtainable solutions to lie on an N-dimensional ellipsoid around the minimum. To get a quantitative grasp on how much these solutions differ in mass and source positions, we solve the equations for a range of random initial conditions. This is a manage- able task since our minimization algorithm is extremely fast, taking only about ∼ 1 second on a 1GHz processor. Also for speed purposes, we fix the grid to the one corresponding to the solution shown in figure 8. Fixing the grid speeds the process significantly since the ΓT Γ matrix and the ΓT θ vec- tor do not need to be recalculated in each minimization. In Diego et al. (2004) a similar process is followed carrying multiple minimizations but this time the grid is changed dy- namically based on the solution found in the previous step. The authors found on that paper that the dispersion of the solutions increases due to the extra variability of the grid. The result is shown in figure 19. The minimization pro- cess described in section 5.2 will stop at a different point in the N-dimensional ellipsoid for each set of different initial conditions. If the total mass of the initial mass distribution is very low the minimization will stop at solutions with mass below the true mass and β positions further away from the center of the cluster than the real ones (dots). If instead we start with a total mass much larger than the true value, the minimization process returns higher masses and β po- sitions closer to the center of the potential (crosses). The Figure 17. Original profile (solid line) compared with the re- covered ones. Dashed line is the biconjugate gradient case, dot- dashed the SVD case, dotted the case is when include the null space. The minimum of the variance is similar but with less mass in the tails. All profiles have been normalized by the critical sur- face density for a source at redshift, zs = 3. Figure 18. Total mass as a function of the threshold ǫ used in the biconjugate gradient minimization. The true mass is 1.119 × 1015h−1M⊙. plane so we expect the solution to be an approximation to the true solution as well. Since a solution comprises not only the masses but the positions and extents of the sources as well, this means that we should expect these to also be ap- proximations to the true ones. This has already been iden- tified as one of the potential pathologies of the algorithm. Namely, if we try to focus the sources into very compact regions, with sizes comparable to the typical galaxy sizes at the relevant redshifts, the obtained mass is in general differ- ent to the true mass. In fact the best results are obtained when the mass plane focuses the arcs into regions which are a few times larger than the extent of the true sources. As we pointed out before, this reconstructed mass corresponds to a short-sighted version of the lens. This problem can be over- come by requiring the minimization algorithms stop once the recovered sources are a few times larger than the true sources. The extent of the true sources can be guessed from the redshift. c(cid:13) 0000 RAS, MNRAS 000, 1 -- 16 14 Diego et al. situation improves when we impose that the total mass of the initial distribution have a reasonable value. In this case, the minimization stops in a region close to both the right mass and the right βs. This argument motivates an iterative minimization where successive estimations for the mass are obtained at each step and used in the next. Among the three algorithms presented in this work, the minimization of the variance is the less powerfull due to its low convergence. It is however interesting from a pedagog- ical point of view since it shows the existing degeneracies between the total mass of the cluster and the positions of the sources. The biconjugate gradient is orders of magnitude faster and is capable of finding the point source solution in a few seconds. The point source solution is however unphysical and a prior associated to the size of the sources is needed in order to stop the minimization process at the proper place. The good news are that the algorithm shows a weak sen- sitivity to this prior provided it is chosen with a minimum of wisdom. Since the point where the minimization stops depends on where the minimization starts, this also allows to study the range of possible models consistent with the data by minimizing many times while changing the initial conditions. This feature makes the biconjugate gradient very attractive for studying the space of solutions. This space can be reduced by including the extra-information contained in the null space, that is, the areas in the sky with no observed arcs. We have seen how this information can be naturally in- cluded in the minimization process by introducing a penalty function. Finally, the SVD has the interesting feature that it al- lows to add extra information regarding the parity or re- solved features in the arcs. This possibility has not been studied in detail in this paper but is definitely worth ex- ploring since it would allow to recover smaller details in the mass distribution. The main drawbacks of the SVD is that the decomposition fails to converge when all the informa- tion is used and one has to use a coarse version of the data instead. The second drawback is that some of the singular values in the matrix W are very small. These values will dominate the inverse if they are not masked by a threshold in the matrix W . Choosing the value for this threshold is not very critical as long as its amplitude is large enough to mask the small singular values in W . The accuracy of the methods presented in this paper al- low for high precision non-parametric mass reconstructions and direct mapping of the dark matter distribution of clus- ters. Previous works have suggested some discrepancy be- tween mass estimates derived from different data (Wu & Fang 1997). Accuracy, combined with the speed of the algo- rithm opens the door to cosmological studies. Strong lens- ing analyses were predicted to yield interesting cosmologi- cal constraints, but due to the uncertainties in our under- standing of galaxy clusters they have yet to live up to these predictions. A fast, per-cent level determination of cluster masses from lensing observations, could allow for sufficient statistical sampling to provide information about cosmolog- ical parameters. An interesting follow up to this work would be to establish what number of strong lensing systems, and what quality of data is necessary in order to make interesting constraints using our methods. However the methods presented here should not be applied indiscriminately. For instance, if the reconstructed Figure 19. Dispersion in the solutions for three different sets of initial conditions. Upper left (dots), we start with a random realization of small masses. Asterisks in the center, starting from a random realization but with a total mass of M ≈ 1.06. Crosses (bottom right), starting with high random masses. In all cases, the threshold was set to ǫ = 2 × 10−10 and the starting β positions where chosen as random in a box of 100 × 100 centered in the center of the image. mass is systematically biased toward recovering less cuspy central regions this reduces the possibility of making con- clusions about mechanisms for dark matter annihilation (Spergel & Steinhardt 2000, Wyithe et al. 2001, Boehm et al. 2004). If one is interested in using our methods to discuss this, that bias must be quantified. Due to lens resolution is- sues we anticipate that our reconstruction algorithms indeed have a bias in this direction and it is likely that other algo- rithms may suffer from similar problems. These and other poetential systematic errors shall be investigated in future works. 7.1 Future work This paper is not intended to be an exhaustive exploration of the accuracy of non-parametric mass reconstruction meth- ods but rather to help re-ignite debate and competition in the field, and demonstrate the feasibility and power of such methods in the face of new data. Specific issues such as mag- nitudes of the mass profile bias, source morphologies, sur- face brightness of the arcs, parity, projection effects as well as other relevant issues discussed above should be explored in order to improve the results. Special emphasis should be paid to the effects the grididification of the lens plane has on the results. An attempt to study this issue has been made on a second paper (Diego et al 2004) where the authors have shown that the grid may play an important role. In this paper we have presented the results using a very specific simulation. Although the algorithm has been tested with different simulations with positive results, it would be de- sirable to make a more detailed study of how the result is affected by issues like the structure of the lens, its symme- try or the number of sources and accuracy in their redshifts. Such a study would be even more interesting if a compara- tive study between parametric and non-parametric methods is done using the same simulations. Also interesting is to explore the effects of adding constraints in the mass of the c(cid:13) 0000 RAS, MNRAS 000, 1 -- 16 type Mi > 0 which may help to stabilize the solution even in the absence of a prior on the source sizes. This can be accomplished by adopting the techniques used in quadratic programming. All theses issues, although very interesting, are beyond the scope of this paper and will be studied in subsequent papers. Although not discussed in this paper, a very interesting piece of work would be about the potentiality of an accu- rate strong lensing analysis as a cosmological tool (e.g Link & Pierce 1998, Yamamoto et al. 2001, Golse et al. 2002, Soucail et al. 2004). Previous works (Yamamoto et al. 2001, Chiba & Takahashi 2002, Sereno 2002, Dalal et al. 2004) suggest that the lensing observables are primarily dependent on the lens model while the dependency in the cosmologi- cal parameters is minor. Constraining the lens model with accuracy can open the window to do cosmology with strong lensing images. Is easy to imagine that a single image of a cluster with dozens of arcs coming from sources at differ- ent redshifts will constrain the lens model rather accurately and then, it will have something to say about the cosmology given the large number of distances involved in the analysis. All these issues are of great interest and they are intended to be studied in subsequent papers. 8 ACKNOWLEDGMENTS This work was supported by NSF CAREER grant AST- 0134999, NASA grant NAG5-11099, the David and Lucile Packard Foundation and the Cottrell Foundation. The work was also partially supported by B. Jain through lousy poker playing. We thank D. Rusin and G. Bernstein for helpful dis- cussions. We would like also to thank J.P Kneib for helpful suggestions. REFERENCES Abdelsalam H.M., Saha P. & Williams, L.L.R., 1998, MNRAS, 294, 734. Inverting the lens 15 Guzik J., Seljak U., 2002, MNRAS, 335, 311. Hudson M.J., Gwyn S.D.J., Dahle H., Kaiser N., 1998, ApJ, 503, 531. Kaiser N. & Squires 1993, ApJ, 404, 441. Kaiser N. 1995, ApJ, 439, L1. Kneib J.-P., Mellier Y., Fort B., Mathez G., 1993,A&A, 273, 367. Kneib J.-P., Mellier Y., Pello R., Miralda-Escud´e J., Le Borgne J.-F.,Boehringer H., & Picat J.-P. 1995, A&A, 303, 27. Kneib J.-P., Ellis R.S., Smail I.R., Couch W., & Sharples R. 1996, ApJ, 471, 643. Kneib J.-P., 2002, The shapes of galaxies and their dark halos, Proceedings of the Yale Cosmology Workshop "The Shapes of Galaxies and Their Dark Matter Halos", New Haven, Connecticut, USA, 28-30 May 2001. Edited by Priyam- vada Natarajan. Singapore: World Scientific, 2002, ISBN 9810248482, p.50 Kneib J.-P., Hudelot P., Ellis R.S., Treu T., Smith G.P., Marshall P., Czoske O., Smail I.R., Natarajan P., 2003, ApJ, 598, 804. Kochanek C.S. & Blandford R.D., 1991, ApJ, 375, 492. Link R. & Pierce M.J., 1998, ApJ, 502, 63. Marshall P.J., Hobson M.P., Gull S.F., Bridle S.L, 2002, MNRAS, 335, 1037. Meneghetti M., Jain B., Bartelmann M., Dolag K., 2004, astro- ph/0409030. Narayan R. & Bartelmann M., Formation of Structure in the Universe, Proceedings of the 1995 Jerusalem Winter School. Edited by A. Dekel and J.P. Ostriker; Cambridge University Press 1999. p. 360. Press W.H., Teukolsky S.A., Vetterling W.T., Flannery B.P., 1997, Numerical Recipes in Fortran 77. Cambrdige Univer- sity Press. Saha, P., Williams, L.L.R., 1997, MNRAS, 292, 148. Saha P. 2000, AJ, 120, 1654. Sand D.J., Treu T. Ellis R.S. 2002, ApJ, 574, 129. Schneider P., Ehlers J., Falco E.E., 1993, Book Review: Gravita- tional lenses. Springer-Verlag, 1993. Schneider P., 1995, A&A...302..639S. Schneider P., & Seitz C. 1995, A&A, 294, 411. Seitz C., & Schneider P. 1995, A&A, 297, 287. Sereno M., 2002, A&A, 393, 757. Soucail G., Kneib J.-P., Golse G. 2004, A&A, 417, L33. Spergel D.N. & Steinhardt P.J., 2000, PhRvL, 84, 3760. Taylor A.N., Dye S., Broadhurst T.J., Benitez N., van Kampen Abdelsalam H.M, Saha P. & Williams, L.L.R., 1998, AJ, 116, E. 1998, ApJ, 501, 539. 1541. Tyson J.A., Kochanski G.P., Dell'Antonio I., 1998, ApJL, 498, Bartelmann M., Narayan R., Seitz S. Schneider P., 1996, ApJ, 107. 464, L115. Beckwith S.V.W. et al. 2003, AAS, 202,1705 Blandford R.D., Narayan R. 1992, ARA&A, 30, 311. Boehm C., Hooper D., Silk J., Casse M., Paul J., 2004, PhRvL, 92, 1301. Brainerd T.G, Blandford R.D, Smail I, 1996, ApJ, 466, 623. Bridle S.L, Hobson M.P., Lasenby A.N., Saunders R., 1998, MN- RAS, 299, 895. Broadhurst T.J., Taylor A.N., Peacock J.A., 1995, ApJ, 438, 49. Broadhurst T.J., Huang X., Frye B., Ellis R., 2000, ApJL, 534, 15. Broadhurst T.J. et.al. 2005, ApJ, 621, 53. Chiba T., Takahashi R. 2002, Progress of Theoretical Physics, 107, 625. Colley W.N., W.N, Tyson J.A., Turner E.L., 1996, ApJL, 461, 83. Dahle H., Hannestad S.; Sommer-Larsen J. 2003, ApJ, 588, 73 Dalal N., Holder G., Hennawi, J.F. 2004, ApJ, 609, 50. Diego J.M., Sandvik H.B., Protopapas P., Tegmark M., Ben- itez N., Broadhurst T., 2004, submitted to MNRAS, preprint astro-ph/0412191 Golse G., Kneib J.-P., Soucail G., 2002, A&A, 387, 788. c(cid:13) 0000 RAS, MNRAS 000, 1 -- 16 Tegmark M., 2002, PhRvD, 66, 3507. Trotter C.S., Winn J.N., Hewitt J. N., 2000, ApJ, 535, 671. Wambsganss J., 1998, Living Reviews in Relativity, 1, 12. Wyithe J.S.B., Turner E.L., Spergel D.N. 2001, ApJ, 555, 504. Wu X.-P., & Fang L.-Z. 1997, ApJ, 483, 62. Yamamoto K., & Futamase T., 2001, Progress of Theoretical Physics, 105, 5. Yamamoto K., Kadoya Y., Murata T., Futamase T., 2001, Progress of Theoretical Physics, 106, 917. APPENDIX A: THE Υ AND Γ MATRICIES The Υ matrix contains the information of how each mass element j affects the ith deflection angle. That is αi = ΥijMj. (33) Precisely how the matrix is constructed is a matter of convenience. Our specific choice is to put both x and y components of all the arc pixels into vectors with 2Nθ elements. The resulting Υ is then a 2Nθ × Ncells matrix. 16 Diego et al. In fact the structure of the Υ matrix and the vec- tors are irrelevant as long as they combine to correctly represent the lens equation as βi = θi − ΥijMj. (34) Each element in the Υ matrix is computed as fol- lows. Υx(i, j) = λ[1 − exp(−δ/2σ2)] δx δ2 where λ = 1015M⊙ 4G c2 Dls DlDs (35) (36) positions of the Ns sources. Similarly, ~βy the central y positions. o will contain This paper has been produced using the Royal Astronomical Society/Blackwell Science LATEX style file. y(j) and δ = qδ2 The index i runs over the Nθ observed θ pixels, in- dex j runs over the Nc elements in the mass vector, M . The factor δx in equation (35) is just the differ- ence (in radians) between the x position in the arc (or x of pixel θi) and the x position of the cell j in the mass grid (δx = θx(i) − θ′ x(j)). Similarly we can define δy = θy(i) − θ′ y. Also, for Υy we only have to change δx by δy. Since we include the fac- tor 1015M⊙ in λ (see equation 36), the mass vector M in equation 4 will be given in 1015 h−1 M⊙ units. The h−1 dependency comes from the fact that in λ we have the ratio Dls/(DlDs) which goes as h. Also we calcu- late Υx and Υy separately, the final Υ matrix entering in equation 4 contains both, Υx and Υy (the same holds for the vectors β and θ). One can rearrange the x and y components in any order. x + δ2 The structure of the Γ matrix is identical to the matrix Υ but with the difference that it has 2Ns ad- ditional columns (the location of the extra-columns is irrelevant as long as it is consistent with the location of the 2Ns βo unknowns in the X vector). Is easy to see that each one of these extra columns (with dimension 2 ∗ Nθ) corresponds to one of the Ns sources. Since the Γ matrix has to contain both the x and y component, the first/second half of each one of the extra columns will be all 0's depending on if it corresponds to the y/x component of Bo. The other half will be full of 0's and 1's, the 1's being in the positions associated with that particular source, the 0's elsewhere. That is, the lens equation can be written explicitly as (a denotes matrix and ~a denotes vector); (cid:18) ~θx ~θy (cid:19) = (cid:18) Υx 1 0 Υy 0 1 (cid:19)  ~M ~βx o ~βy o   (37) Where again, ~θx and ~θy are two Nθ dimensional vectors containing the x and y positions, respectively, of the pixels in the observed arcs. The two (Nθ × Nc matrices Υx and Υy contain the x and y lensing effect of the cell j on the θ pixel i. The Nθ × Ns dimensional matrices 1 and 0 are full of 0's (the 0 matrix) or contain 1's (the 1 matrix) in the i positions (i ∈ [1, Nθ]) where the ith θ pixel comes from the j source (j ∈ [1, Ns]) and 0 elsewhere. The vector ~M contains the Nc gridded masses we want to estimate. ~βx o contains the central x c(cid:13) 0000 RAS, MNRAS 000, 1 -- 16
astro-ph/0507396
1
0507
2005-07-18T04:30:26
GALEX Observations of an Energetic Ultraviolet Flare on the dM4e Star GJ 3685A
[ "astro-ph" ]
The Galaxy Evolution Explorer (GALEX) satellite has obtained high time resolution ultraviolet photometry during a large flare on the M4 dwarf star GJ 3685A. Simultaneous NUV (1750 - 2800A) and FUV (1350 - 1750A) time-tagged photometry with time resolution better than 0.1 s shows that the overall brightness in the FUV band increased by a factor of 1000 in 200 s. Under the assumption that the NUV emission is mostly due to a stellar continuum, and that the FUV flux is shared equally between emission lines and continuum, then there is evidence for two distinct flare components for this event. The first flare type is characterized by an exponential increase in flux with little or no increase in temperature. The other involves rapid increases in both temperature and flux. While the decay time for the first flare component may be several hours, the second flare event decayed over less than 1 minute, suggesting that there was little or no confinement of the heated plasma.
astro-ph
astro-ph
Astrophysical Journal accepted, 2005 July 11 Preprint typeset using LATEX style emulateapj v. 6/22/04 GALEX OBSERVATIONS OF AN ENERGETIC ULTRAVIOLET FLARE ON THE DM4E STAR GJ 3685A Richard D. Robinson,1 Jonathan M. Wheatley,2 Barry Y. Welsh,2 Karl Forster,3 Patrick Morrissey,3 Mark Seibert,3 R. Michael Rich,4 Samir Salim,4 Tom A. Barlow,3 Luciana Bianchi,5 Yong-Ik Byun,6 Jose Donas,7 Peter G. Friedman,3 Timothy M. Heckman,5 Patrick N. Jelinsky,2 Young-Wook Lee,6 Barry F. Madore,8 Roger F. Malina,7 D. Christopher Martin,3 Bruno Milliard,7 Susan G. Neff,9 David Schiminovich,3 Oswald H. W. Siegmund,2 Todd Small,3 Alex S. Szalay,5 and Ted K. Wyder3 5 0 0 2 l u J 8 1 1 v 6 9 3 7 0 5 0 / h p - o r t s a : v i X r a Astrophysical Journal accepted, 2005 July 11 ABSTRACT The Galaxy Evolution Explorer (GALEX ) satellite has obtained high time resolution ultraviolet photometry during a large flare on the M4 dwarf star GJ 3685A. Simultaneous NUV (1750 - 2800 A) and FUV (1350 - 1750 A) time-tagged photometry with time resolution better than 0.1 s shows that the overall brightness in the FUV band increased by a factor of 1000 in 200 s. Under the assumption that the NUV emission is mostly due to a stellar continuum, and that the FUV flux is shared equally between emission lines and continuum, then there is evidence for two distinct flare components for this event. The first flare type is characterized by an exponential increase in flux with little or no increase in temperature. The other involves rapid increases in both temperature and flux. While the decay time for the first flare component may be several hours, the second flare event decayed over less than 1 minute, suggesting that there was little or no confinement of the heated plasma. Subject headings: stars: flare -- stars: late-type -- stars: individual (GJ 3685A) -- stars: variables: other (UV Ceti) -- ultraviolet: stars 1. INTRODUCTION UV Ceti stars exhibit many of the activity phenom- ena observed on the Sun, such as flares, dark spots and variable emission from the chromosphere and corona (Thomas & Weiss 2004). Such surface activity is linked to their magnetic field strength and surface coverage, which can be several times larger than that found on the Sun. When quiescent, UV Ceti flare stars are low luminosity (MV > 8), low temperature (T = 2500 - 4000 K) M-type dwarfs. During outbursts, these stars brighten dramatically over timescales of a few sec- onds to hours at all wavelengths from X-ray to radio (Haisch, Strong, & Rodono 1991). The vast majority of stellar flare observations involve photometric monitoring using the Johnson UBV filters (Lacy et al. 1976). Normally this involves high speed sampling using a single transmission filter. Occasionally, however, observations are obtained in multiple filters by sequencing using time steps of a few seconds to about 1 minute. This technique has the advantage of obtaining 1 Institute for Astrophysics and Computational Sciences, Catholic University of America, 200 Hannan Hall, Washington, DC 20064; [email protected] 2 Experimental Astrophysics Group, Space Sciences Labora- tory, University of California, 7 Gauss Way, Berkeley, CA 94720; [email protected], [email protected] 3 California Institute of Technology, MC 405-47, 1200 East Cal- ifornia Boulevard, Pasadena, CA 91125 4 Department of Physics and Astronomy, University of Califor- nia, Los Angeles, CA 90095 5 Center for Astrophysical Sciences, The Johns Hopkins Univer- sity, 3400 N. Charles St., Baltimore, MD 21218 6 Center for Space Astrophysics, Yonsei University, Seoul 120- 749, Korea 7 Laboratoire d'Astrophysique de Marseille, BP 8, Traverse du Siphon, 13376 Marseille Cedex 12, France 8 Observatories of the Carnegie Institution of Washington, 813 Santa Barbara St., Pasadena, CA 91101 9 Laboratory for Astronomy and Solar Physics, NASA Goddard Space Flight Center, Greenbelt, MD 20771 flare colors, allowing estimates of temperatures and sizes, and is reasonably valid during the long duration decay phase of the events. However, during the impulsive phase the fluxes can change on time scales of seconds, so the color data becomes highly uncertain. A second limitation is that the flare energy distribution peaks at wavelengths shorter than 4000A (van den Oord et al. 1996), so that the Johnson photometry only samples the wing of this distribution, which is contaminated by the stellar photo- sphere. The GALEX satellite (Martin et al. 2005) provides the opportunity for obtaining vastly improved photometric measurements of stellar flares. Using a dichroic beam splitter and two photon counting detectors, this telescope is able to simultaneously monitor a 1.25o field of view in the FUV (1350 - 1750 A) and NUV (1750 - 2800A) regions with 5 arcsec and 6.5 arcsec angular resolution respectively. This allows the accurate determination of UV fluxes and colors at a time resolution which is lim- ited only by the count rate and the required S/N values. In this paper we describe a serendipitous observation of a large flare seen by the GALEX telescope on the dM4e star GJ 3685A. Some of the physical properties of this event are discussed and we develop a schematic model which is compared with models developed from other flare observations. 2. OBSERVATIONS AND DATA REDUCTION The flare event on GJ 3685A was detected on 2004 April 20 during a simultaneous FUV and NUV ob- servation of the Medium Imaging Survey (MIS) field MISDR1 13062 0283. The exposure began at 19:42:06 UT and lasted for 1244s. At present, the standard GALEX Data Analysis Pipeline is only designed to pro- duce a calibrated image of the field and thus does not retain timing information. To construct a photometric time sequence, we began with the original time-tagged 2 ROBINSON, et al. photon lists and integrated all photons within a 1.8 ar- cmin aperture centered on the source. The background count rate was measured within an annulus extending from 1.8 to 2.4 arcmin around the star, during the first 500 s of the observation, in order to avoid contamination by scattered light when the flare was at its brightest. The FUV and NUV background rates are 3.2 cts s−1 and 38.5 cts s−1, respectively. The GALEX microchannel plate detectors exhibit a lo- cal non-linearity of 10% affecting point sources with in- put count rates of 90 cts s−1 in FUV and 470 cts s−1 in NUV (Morrissey et al. 2005). In this paper, we use Mor- rissey's published local non-linearity corrections, which have been calibrated with white dwarf standard stars up to 3300 cts s−1 in FUV and 8400 cts s−1 in NUV. The GJ 3685A flare exceeded this calibrated range in the FUV for a few seconds during the brightest parts of the flare: we have omitted these data from the analysis because the uncertainty in the true count rate is unknown in this high count-rate regime. The results of the analysis are given in Figure 1, which shows both the flux variations and the ratio of the FUV to NUV fluxes. In these plots the photon count rates have been converted to fluxes (in erg cm−2 s−1 A−1) using the conversion factors of 1.4 × 10−15 and 2.6 × 10−16 for the FUV and NUV, respectively (Morrissey et al. 2005). Initially, the star showed a 'quiescent' flux of ∼ 10−15 erg cm−2 s−1 A−1 and a flux ratio (or color) of 1. The start of the stellar flare is lost in the noise of the background. A reasonable estimate seems to put the start somewhere between 450 and 500 s and it is definitely underway in both the NUV and FUV bands by 500 s. Between 500 s and 630 s the flux in both bands increases exponentially, with an e-folding time of ∼60 s. During this time the average color remains constant, at a value of near 0.8. At 650 s the rate of flux increase jumps dramatically, with e-folding times of 23 s in the NUV and 12 s in the FUV, and the FUV/NUV ratio increases from < 1 to more than 6. Both temperature and FUV/NUV ratio reach a peak at 700 s. During this peak period, we have omitted the data because the FUV non-linearity correc- tion is uncertain. After the peak, which only lasts for about 20 s, the flux and FUV/NUV ratio rapidly decrease. Note that the time variations of both the color and the flux for the 100 s centered at the flare peak are symmetrical about the peak, i.e. the rise times and decay times are about the same. By 750 s the color ratio is again near 1 and the flux settles into a slow decline which is interrupted at 865 s by a second flux enhancement. This second en- hancement is somewhat more complicated than the first, but follows the same basic trend of rapid flux increase ac- companied by a rapid increase in FUV/NUV, this time to values in excess of 10. The decay phase for the second enhancement is slower and more complex than was the case for the first enhancement. By the end of the obser- vation the flux and color had returned to the levels seen prior to the second enhancement, though the flux is still well above the 'quiescent' level. 3. ANALYSIS 3.1. Pre-flare Activity On 2004 March 2 GALEX obtained a 110 sec 'qui- escent' observation of GJ 3685A which showed a flux of 1.7 ± 0.5 × 10−16 erg cm−2 s−1 A−1 in the FUV and 3.0 ± 0.3 × 10−16 erg cm−2 s−1 A−1 in the NUV, giv- ing a color ratio of 0.57 ± 0.19. On 20 April, before the onset of the flare event the average 'quiescent' flux was 1.54 × 10−15 erg cm−2 s−1 A−1 in the NUV and 1.52 × 10−15 erg cm−2 s−1 A−1 in the FUV band, result- ing in a flux ratio of 1. Since the FUV/NUV flux ratio is directly related to the effective emission temperature (see section 3.2) we see that preceding the flare the stel- lar emission had not only increased in flux by a factor of 5 in the NUV, but also increased in temperature. One possible explanation is that the pre-flare enhance- ment results from an increased level of microflaring ac- tivity. To test this notion, we performed an analysis on the first 400 s of the data set using the binning tech- nique described by Robinson et al. (1995, 1999). Indi- vidual microflares can be detected by binning the time sequence and searching for times when the counts per bin are significantly higher than those expected from random chance. The maximum visibility occurs when the binning factor is near the lifetime of the event. A distribution of small events which are not individ- ually time resolvable can be determined by comparing the distribution of count rates with a reference distribu- tion from a non-variable source, as described in Robinson (1999). This method reveals that when the binning fac- tor is much smaller than the flare lifetime, then event noise will dominate and both distributions will be Pois- son. However, as the binning factor approaches the flare lifetime, the microflares show up as an enhancement of the high count rate tail of the distribution. Examining the time sequences for binning factors rang- ing from 1 s to 20 s for the GJ 3685A flare data shows no strong evidence for either individual or unresolved flare events, while a regression analysis showed that the level of activity remained statistically constant prior to flare onset. A problem, however, is that the background levels are large compared to the stellar signal. For ex- ample, in the NUV the estimated background is 38.5 counts s−1, while the stellar signal was only 7.5 counts s−1. Thus, noise from the background would swamp all but the largest microflare events. 3.2. Empirical Modeling Since we have no spectroscopic information about this flare it is not possible to perform any type of detailed modeling or radiative transfer calculations on these UV data. However, some good qualitative and semi-quantitative insights can be obtained by assuming black-body emission. The reader should be aware that this assumed model is only one of several possible flare model scenarios. However, this simple theoretical ap- proach has been widely used in the interpretation of op- tical photometry of flare events and yielded some impor- tant insights into flare phenomena (Kahler et al. 1982; van den Oord et al. 1996; Hawley et al. 2003). We have a definite advantage in these UV observations since the photospheric emission from the star is so small that it can be neglected. The first step in our simplified analysis is to determine the relation between the black body temperature and the measured FUV/NUV flux ratio. We have calculated the black body spectrum for a number of different tem- peratures, multiplied by the effective area curves for the GALEX Ultraviolet Flare 3 GALEX FUV and NUV filters and then divided by the total effective area for each filter. This gives the expected flux for a black body, which can be directly related to the average flux deduced from the measured count rates. The relation between temperature and flux ratio is shown in Figure 2. Note that the flux ratio increases dramatically between 4000 and 19000 K (when the peak of the black body distribution falls within the FUV bandpass) and then increases much more slowly toward high tempera- tures. The ratio eventually reaches a value of just over 3 at temperatures of 50,000 K. This immediately shows that our black-body assumption is not totally valid, since we measured ratios during the flare peaks in excess of 6. This is not a surprise, since the FUV wavelength region has numerous strong emission lines (e.g. C II, C IV, Si II, Si IV, etc) which can dominate the continuum, especially during 'quiescent' periods outside of major flare events (van den Oord et al. 1996). Assessing the relative contributions from continuum and/or emission lines at near UV wavelengths is still an open problem for flare research (Hawley & Fisher 1992). We note that the GALEX NUV channel is not signifi- cantly contaminated by the (expected) strong MgII emis- sion lines at 2800A due to the low transmission of the GALEX NUV instrument at this wavelength. However, previous flares have shown an (albeit far less) increase in the NUV emission from the FeII line multiplet at 2600A (Haisch et al. 1987), whereas Butler et al. (1981) have observed a significant rise in the near UV contin- uum during a flare on the star Gl 867A. For the present case of the flare on GJ 3685A, it is difficult to imagine how line emission could be the sole major contributor to the more than 5 UV magnitude increase in NUV flux observed during the flare. Thus, in determining the effec- tive temperature from the FUV/NUV flux ratio, we ar- bitrarily assume that approximately half of the FUV flux was from emission lines and the other half from black- body emission. This is consistent with the time-resolved spectroscopy of Hawley et al. (2003), where the FUV line-to-continuum flux ratio is 0.8, during the brightest flare on AD Leo reported in that paper. Furthermore, Hawley & Fisher (1992) also provide support for the no- tion that the GALEX NUV band is dominated by the star's continuum. A black body of given temperature has a well de- fined emission. Thus, having determined a temper- ature we can define the effective stellar surface cov- erage required to produce the observed NUV flux at the earth, assuming a distance to the star of 14.6 pc (Astronomisches Rechen-Institut 1998). The results of the calculations are presented in Figure 3, where we have converted the flare surface area to an effective source ra- dius. The calculated temperature in the pre-flare phase was about 12,000 K with a deduced scale of about 3500 km. The exact values are uncertain because this phase is dominated by the emission lines in the FUV and the ac- tual structure is most likely composed of small elements spread over a large area. When the flare starts the ef- fective temperature drops to about 10,000 K and the radius increases linearly from 3,500 km at 500 s to about 25,000 km at 650 s, suggesting an initial expansion ve- locity of at least 140 km s−1. The deduced temperature increases dramatically at 660 s, becomes undefined near the flare peak and then rapidly decreases back to a value of about 10,000 K. During this time the deduced radius decreases to 7,000 km and then increases back to 25,000- 28,000 km, about the same value as that seen before the large temperature increase. The temperature and area then remain relatively constant until the start of the sec- ond temperature enhancement at 880 s, where the area again dramatically decreases and then slowly recovers to a value of about 32,000 km. Having estimated a temperature and area for the flare it is possible to determine the total optical and UV en- ergy by integrating under the appropriate black body curve. The result of this calculation is shown in the Fig- ure 4. The overall appearance of this event is two short duration bursts at 670-730 s and 880-970 s superimposed on a more slowly varying event which started at 450 s and which was slowly decaying at the end of the observation. The flare is obviously very energetic, with luminosities reaching more than 1032 erg s−1 and a total integrated energy in excess of 1034 erg. This is comparable to the largest recorded M-dwarf flare outbursts, such as the 1985 flare on the star AD Leonis (Hawley & Petterson 1991). Another way to put this into perspective is to note that the total (optical plus UV) luminosity emit- ted from the photosphere of a M4 star is on the order of 2.0 × 1031 erg s−1. Thus, at the peak the flare would outshine the star by more than a factor of 10. 4. DISCUSSION Assuming that our black body emission model dis- cussed in Section 3.2 provides a reasonable description of the flare event on GJ 3685A (and given the caveat that there may be other models that can provide an al- ternate interpretation of the data that we have chosen not to explore due to the lack of spectroscopic informa- tion on this flare), we can now postulate a schematic model for the event which satisfies all of the observa- tions and is consistent with observations of flares seen on other stars. The flare itself appears to be composed of two distinct types of events, which we will refer to as type A and type B. The flare begins with a type A event, which is characterized by a FUV/NUV flux ratio that is nearly the same as that seen in the pre- flare activity (plage) and remains constant as the flux increases. We conclude that the spectrum has a very plage-like appearance and that the source structure, and possibly the heating processes, are also very plage-like. We postulate that this phase was initiated by a mag- netic reconnection somewhere within a highly stressed active region. This initial energy release triggers recon- nection events in adjacent magnetic structures, leading to an avalanche (Lu & Hamilton 1991; Carbonneau et al. 2001) which propagates throughout the region at a veloc- ity of at least 140 km s−1, which probably represents the local Alfven velocity. The increase in flux is then simply the result of expanding area coverage. At about 670 s (roughly 3 minutes after the start of the flare) the disturbance which is responsible for flare A intersects a highly unstable magnetic structure and triggers an explosive release of energy, resulting in the first type B flare (B1), which occurs between 670 s and about 730 s. The rapid increase in temperature and flux is compatible with a chromospheric evapora- tion model (Haisch, Strong, & Rodono 1991), in which a 4 ROBINSON, et al. large amount of energy (probably in the form of energetic particles) is suddenly released near the top of a magnetic loop complex and propagates down towards the photo- sphere. When it strikes the denser atmospheric layers it impulsively heats the material, which then expands back into the loop. Normally, this type of event will have a long lasting decay as the heated material cools through conduction and radiation. In this case, however, the short duration of the decay suggests that the energy release was sufficient to cause a disruption of the confin- ing loops (Reale, Bocchino, & Peres 2002). The source size for flare B1 is also significantly smaller than that of flare A, since the deduced flare area decreases rapidly when the emission from B1 dominates that from flare A. The structure responsible for flare B1 is probably near the edge of the active region associated with flare A, since the expansion of flare A ends during flare B1. Appar- ently, all of the available structures which can support flare A emission have been activated by that time. It is possible that the disturbance responsible for flare A could continue into regions outside of the initial active region. More likely, the energy release initiating flare B1 also generated an Alfvenic shock, similar to the Morton waves seen on the Sun (Athay & Morton 1961). This dis- turbance propagates to an unstable magnetic structure in a nearby active region and triggers a second B type flare (flare B2) which starts at around 880 s. This flare is both stronger and more complex than B1, indicating more complex field configuration. It also shows a better developed thermal decay, implying that at least some of the hot plasma is confined in the loops. Note also that this is an isolated event, without an accompanying type A event, since there was no increase in source size prior to the onset of the temperature enhancement. By the end of the observation it appears that flare B2 has completely faded and flare A is again visible. Un- fortunately, we do not have much information about the decay phase of flare A. However, from the fact that the integrated flux at the end of the observation at 1250 s is very similar to that at 800 s, after the decay of flare B1, suggests that it may take an hour or more for flare A to decay to pre-flare levels. 5. CONCLUSIONS We report on a large flare from the dM4e star GJ 3685A which was simultaneously observed in the NUV and FUV by the GALEX satellite. Under the as- sumption of a blackbody emission model in which half of the FUV flux arises from line emission and half from continuum, and that the NUV flux arises solely from an increase in the continuum, there is strong evidence for two distinct classes of flares during this event. The first (type A) is characterized by an exponential increase in flux with little of no change in emission temperature, as measured from the FUV/NUV flux ratio. This behavior is compatible with an avalanche model. The fact that the flux ratio during this phase was very similar to that seen in the pre-flare plage emission also supports the idea that plages are heated by microflares, as first proposed by (Parker 1988). The second class (type B) is characterized by an impul- sive increase in both flux and temperature. This class is consistent with a classical explosive event and may be re- lated to the solar two ribbon flare. Two flares of this type were seen during the event. The first (flare B1) probably originated in the same active region as the type A flare. The lack of a well developed decay phase suggests that the confining magnetic structure was disrupted. The sec- ond flare (B2) may have originated outside of that region, being triggered by a disturbance formed during the flare B1 energy release. The 2004 April 20 flare began when GJ 3685A was al- ready in a state of enhanced activity in which both the flux and temperature are substantially higher than in a previous GALEX observation on 2004 March 2. It is un- clear whether this enhancement is the result of coronal variations, or arises from the rotation of an intense active region onto the visible surface of the star, or if it comes from some pre-flare energy release. In order to obtain better models of future dMe star flare events we recom- mend using the low-resolution spectroscopic grism mode of GALEX, instead of the imaging photometric mode. Such future observations would provide better insights into the relative contributions from the UV continuum and/or line emission during these large releases of stellar energy. GALEX is a NASA Small Explorer, launched in 2003 April. We gratefully acknowledge NASA's support for construction, operation, and science analysis for the GALEX mission. Financial support for this research was provided by NASA grant NAS5-98034. REFERENCES Astronomisches Rechen-Institut 1998, ARI Database for Nearby Stars. http://www.ari.uni-heidelberg.de/aricns Athay, R.G., & Morton, G.E. 1961, 133, 935 Butler, C.J., et al. 1981, MNRAS, 197, 815 Carbonneau, P., McIntosh, S.W., Liu, H-L, & Bogdan, T.J. 2001, Solar Physics, 203, 321 Gliese, W., & Jahreiss, H. 1991, Preliminary Version of the Third Catalogue of Nearby Stars. Greenbelt, MD: Goddard Space Flight Center Haisch, B., et al. 1987, A&A, 181, 96 Haisch, B., Strong, K., & Rodono, M. 1991, ARA&A, 29, 275 Hawley, S.L., & Fisher, G.H. 1992, ApJS, 78, 565 Hawley, S.L., & Petterson, B.R. 1991, ApJ, 378, 725 Hawley, S.L., et al. 2003, ApJ, 597, 535 Kahler, S., et al., ApJ, 252, 239 Lacy, C.H., Moffett, T.J., & Evans, D.S. 1976, ApJS, 30, 85 Lu, E.T., & Hamilton, R. 1991, ApJ, 380, L89 Martin, D.C., et al. 2005, ApJ, 619, L1 Morrissey, P., et al. 2005, ApJ, 619, L7 Nakajima, H., Dennis, B.R., Hoyng, P., Nelson, G., Kosugi, T., & Kai, K. 1986, ApJ, 288, 806 Parker, E.N. 1988, ApJ, 330, 474 Reale, E., Bocchino, F., & Peres, G. 2002, A&A,383, 952 Robinson, R.D., Carpenter, K.G., Percival, J.W., & Bookbinder, J.A. 1995, ApJ, 451, 795 Robinson, R.D., Carpenter, K.G., & Percival, J.W., 1999, ApJ, 516, 916 Thomas, J.H., & Weiss, N.O. 2004, ARA&A, 42, 517 van den Oord, G.H.J., et al. 1996, A&A, 310, 908 GALEX Ultraviolet Flare 5 Fig. 1. -- (Top) Ratio of FUV to NUV fluxes. In this and all other figures, data are omitted during the time intervals in which the FUV count rate exceeds the maximum calibrated count rate of 3300 cts s−1. (Bottom) Calibrated fluxes from the GALEX FUV (solid) and NUV (dashed) channels for the entire observation of 2004 April 20. Data have been binned into 25 s intervals for 0-450 s, 5 s intervals between 450 and 630 s and 1 s intervals for times greater than 630 s. 6 ROBINSON, et al. Fig. 2. -- Relation between black body temperature and the FUV/NUV flux ratio determined by convolving the appropriate black body distribution with the GALEX FUV and NUV filter effective area curves. GALEX Ultraviolet Flare 7 Fig. 3. -- (a) Effective black body temperatures derived from the measured FUV/NUV flux ratio using the calibration presented in Figure 2. Preflare activity has been estimated using an integration over the first 400 s. Data between 400 s and 630 s is deduced using 5 s bins, while later data uses 1 s bins. (b) Effective radius of the source deduced from the observed flux, as described in the text. Data for the first 630 s has been binned into 5 s intervals, later data has been binned into 1 s increments. (c) Measured FUV flux, binning as in (b). 8 ROBINSON, et al. Fig. 4. -- Estimated luminosity, assuming black body emission at the temperature and source size presented in Figure 3.
astro-ph/0302521
1
0302
2003-02-25T15:38:33
Angular momentum redistribution and the evolution and morphology of bars
[ "astro-ph" ]
Angular momentum exchange is a driving process for the evolution of barred galaxies. Material at resonance in the bar region loses angular momentum which is taken by material at resonance in the outer disc and/or the halo. By losing angular momentum, the bar grows stronger and slows down. This evolution scenario is backed by both analytical calculations and by $N$-body simulations. The morphology of the bar also depends on the amount of angular momentum exchanged.
astro-ph
astro-ph
Angular momentum redistribution and the evolution and morphology of bars E. Athanassoula1 Observatoire de Marseille, 2 Place Le Verrier, 13248 Marseille cedex 04, France Abstract. Angular momentum exchange is a driving process for the evolution of barred galaxies. Material at resonance in the bar region loses angular momentum which is taken by material at resonance in the outer disc and/or the halo. By losing angular momentum, the bar grows stronger and slows down. This evolution scenario is backed by both analytical calculations and by N -body simulations. The morphology of the bar also depends on the amount of angular momentum exchanged. 1 Introduction Bars are common features of disc galaxies. De Vaucouleurs ([11]), using a classi- fication based on images at optical wavelengths, found that roughly one third of all disc galaxies are barred (family SB), while yet another third have small bars or ovals (family SAB). Observations in the near infrared have shown that galax- ies that had been classified as non-barred from images at optical wavelengths may have a clear bar component when observed in the near infrared. Thus Es- kridge et al. ([17]) classified more than 70% of all disc galaxies as barred, while Grosbøl, Pompei & Patsis ([18]) found that only ∼ 5% of all disc galaxies are definitely non barred. Bars come in a large variety of strengths, lengths, masses, axial ratios and shapes. Great efforts have been made in order to obtain some systematics on bar structure and important advances have been made. Elmegreen & Elmegreen ([16]) have shown that earlier type bars are relatively longer (i.e. relative to the disc diameter) on average than bars in later type galaxies. They also find that early type bars have flat intensity profiles along the bar major axis, while late type bars have exponential-like profiles. A correlation has been found ([5], [27]) be- tween the length of bars and the size of bulges. This is in good agreement with the trend found in [16], since earlier type galaxies have larger bulges than late types. Important differences between early and late type bars are also found with the Fourier decomposition of the surface density. Indeed the relative m = 2 and 4 components are much stronger in early than in late type bars. Moreover, the higher order components (m = 6 and 8), which for the late type bars are negligi- ble, are still important for early types. Finally, the shape of the bar isodensities differ and Athanassoula et al. ([7]), using a sample of strongly barred early type galaxies, showed that their bar isophotes are rectangular-like, particularly in the region near the end of the bar. The first trials of N -body simulations (e.g. [28]) show that bars grow sponta- neously and are long-lived. Yet it is only recently that simulations have achieved 2 E. Athanassoula sufficient quality to provide information on the morphology of N -body bars and on the mechanisms that govern bar formation and bar evolution. I will here dis- cuss some of the latest results of N -body simulations. I will argue that it is the exchange of angular momentum within the galaxy that will determine the bar strength and the rate at which the pattern speed decreases after the bar has formed, as well as the bar morphology. 2 Angular momentum exchange and bar evolution Exchange of energy and angular momentum between stars at resonance with a spiral density wave has been first discussed by Lynden-Bell & Kalnajs ([26]). Using linear perturbation theory, these authors showed that, for a steady forcing, stars emit, or absorb, angular momentum only if they are at resonance. Stars at the inner Lindblad resonance (hereafter ILR) lose angular momentum, while stars at the outer Lindblad resonance (hereafter OLR) gain it. This ground- braking work has to be extended in order to be applied to bars in general and N -body bars in particular. HI observations, basically starting with [10], have now established that, if Newton's law of gravity is valid, then disc galaxies are embedded in a dark matter component, called the halo, whose mass exceeds that of the disc. This component should now be taken into account as an extra partner in the angular momentum exchange process. Furthermore, bars are strongly non-linear features, since they contain a large fraction of the mass in the inner parts of the disc and a considerable part of the total disc mass. Thus any linear theory should be thought of as a guiding line, to be supplemented by and tested against adequate N -body simulations. It is obvious that such simulations should be fully self-consistent, since rigid components can not exchange energy and angular momentum. In [4] I extended the analytical work of [26] to include spheroidal components, like a halo and/or a bulge, and also supplemented it with fully self-consistent N -body simulations. In the analytical part I showed that, if the distribution function of the spheroidal component is a function only of the energy, then at all resonances the halo and bulge particles can only gain angular momentum. Also, since the bar is a strongly nonlinear feature, higher multiplicity resonances should be taken into account. Thus angular momentum is emitted by particles (stars) at the resonances in the inner disc, mainly the ILR, but also the inner -1:m resonances nearer to corotation (hereafter CR). It is absorbed by disc par- ticles (stars) at the OLR, or the outer 1:m resonances, outside corotation, or by the resonant particles in the halo and/or bulge components. Since the bar is a negative angular momentum perturbation ([26]), by losing angular momentum it will grow. This clearly outlines a scenario for the evolution of barred galaxies. 3 The effect of the halo on bar evolution As the bar loses angular momentum, it grows stronger. This, however, can only happen if there are absorbers that can absorb the angular momentum that the Evolution and morphology of bars 3 Fig. 1. Basic information on simulations LH (time t = 600) and RH (time t = 900). The two upper rows give the circular velocity curves at time 0 and t. The dashed and dotted lines give the contributions of the disc and halo respectively, while the thick full lines give the total circular velocity curves. The third row of panels gives the isocontours of the density of the disc particles projected face-on and the fourth and fifth row give the side-on and end-on edge-on views, respectively. The side of the box for the face-on views is 10 length units and the height of the box for the edge-on views is 3.33. The isodensities in the third row of panels have been chosen so as to show best the features in the bar and in the inner disc. No isodensities for the outer disc have been included, although the disc extends beyond the area shown in the figure. The sixth row of panels gives the m = 2, 4, 6, and 8 Fourier components of the mass. bar region emits. Thus the existence of a massive halo component, whose res- onances can absorb considerable amounts of angular momentum, will help the bar grow. At first sight this may be thought to go against old claims that haloes stabilise bars. In fact, a more precise wording is necessary. Indeed, the halo slows down the bar growth in the initial stages of the evolution. At later stages, however, the situation can be reversed, since the halo may absorb the angular 4 E. Athanassoula momentum emitted by the bar, and thus it may allow the latter to grow further. Thus bars that grow in halo-dominated discs can be stronger than bars that grow in disc-dominated surroundings. This effect was not seen till recently, since the older studies were either 2D (e.g. [33], [8]), or 3D but with few particles (e.g. [31]), or had rigid haloes. In all these cases the halo was denied from the onset its destabilising influence. Its effect becomes clear in fully self-consistent N -body simulations, with an adequate particle number and resolution. Thus [6] showed that stronger bars can grow in cases with more important halo components. The influence of the halo is also illustrated in Fig. 1, where I compare the results of two N -body simulations. Initially their disc is exponential, with unit mass and scale-length (Md = 1, Rd = 1) and its Q parameter ([32]) is independent of radius and equal to 1.2. Since G = 1, taking the mass of the disc equal to 5 × 1010 M⊙, and its scale-length equal to 3.5 kpc implies that the unit of velocity is 248 km/sec and the unit of time is 1.4 × 107 yrs. This calibration is reasonable, but is not unique, so in the following I will give all quantities in computer units. The reader can then convert the values to astronomical units according to his/her needs. The halo component is spherical, non-rotating and has an isotropic velocity distribution. It follows a pseudo-isothermal radial density profile ([19]) and has a total mass Mh = 5, a core radius γ = 0.5 and a cutoff radius of rc = 10. Its density is truncated at 15 disc scale-lengths, i.e. at a radius containing more than 96% of its mass. In building the initial conditions I loosely followed [19] and [6], and the simulations were run on the Marseille GRAPE-5 systems (for a description of the GRAPE-5 boards see [22]). The only difference between the initial conditions of the two simulations is that for simulation LH, illustrated in the left panels, the halo is live and represented by roughly 106 particles, while for simulation RH, illustrated in the right panels, it is rigid, i.e. represented by an analytical potential and thus can neither emit nor absorb angular momentum. Although their initial conditions are so similar, the two simulations evolve in a very different way. After some initial multi-spiral episodes, LH forms a bar which grows stronger with time. Its morphology at t = 700 is shown in the left panels. The bar is long and strong and has ansae-type features near the end of its major axis. It is surrounded by a ring, which can be compared to the inner rings often observed in barred galaxies. The bar formation entails considerable redistribution of the disc matter, both radially and azimuthally. On the other hand the disc in simulation RH stays close to axisymmetric, except for some multi-armed spirals which die out with time. Only at the latest stages of the evolution does it form an oval distortion, and even that is weak and short and is confined to the innermost parts of the disc, as can be seen for t = 900 in the right panels of Fig. 1. I show this simulation at a later time than that for simulation LH because at earlier times there is very little structure visible. Seen edge-on with the bar seen side-on (i.e. with the line of sight along the bar minor axis), simulation LH exhibits a very strong peanut, which is totally absent from simulation RH (fourth row of panels). Seen edge-on with the bar seen end-on (i.e. with the line of sight along the bar major axis), the peanut in LH resembles a bulge (left panel on fifth row). This underlines the hazards Evolution and morphology of bars 5 involved in classifying edge-on galaxies, since the classifier may in such cases easily misinterpret the bar for a bulge. A useful way of quantifying the bar strength is with the help of the Fourier components of the mass, or density. These can be defined as Fm(r) = pA2 m(r) + B 2 m(r)/A0(r), m = 0, 1, 2, ..... (1) Am(r) = 1 π Z 2π 0 Σ(r, θ)cos(mθ)dθ, m = 0, 1, 2, ..... (2) where and Bm(r) = 1 π Z 2π 0 Σ(r, θ)sin(mθ)dθ, m = 1, 2, ..... (3) For runs LH and RH, these components for m = 2, 4, 6 and 8 are shown in the lower panels of Fig. 1. For run LH all four components are important, due to the strength of the bar. Their amplitude decreases with increasing m, while the location of the maximum shifts outwards. On the other hand, for model RH only the m = 2 component stands out from the noise, but its amplitude is rather small, smaller than e.g. that of the m = 8 for model LH. Since the only difference between the initial conditions of models LH and RH is that the halo of the one is responsive, while that of the other is rigid, we can conclude that the halo response is crucial for determining the evolution of barred galaxies. 4 Bar slow-down Fig. 2. Bar pattern speed of simulation LH as a function of time. I ran a large number of simulations similar to those described in the previous sections. I noted that, as it loses angular momentum, the bar grows longer, 6 E. Athanassoula and more massive, thus stronger. Angular momentum loss, however, is not only linked to an increase in the bar strength. It is also linked to a slow-down, i.e. to a decrease of the bar pattern speed Ωp with time. Such a slow-down has indeed been seen in a number of simulations and has also been predicted analytically ([34], [36], [24], [25], [20], [1], [12], [13]). It can also be seen in Fig. 2, which shows the run of the bar pattern speed with time for simulation LH, whose morphology at time t = 700 is shown in the left panels of Fig. 1. Note that the bar slows down considerably with time. 5 Resonances In order for haloes to be able to absorb angular momentum, they need to have a considerable fraction of their mass at resonance. This was shown to be true in ([2]). I will illustrate it here for model LH. The procedure is the same as that followed in [2]. I calculate the potential from the mass distribution in the disc and halo component at time t = 800, by freezing all motion except for the bar, to which I assign bulk rotation with a pattern speed equal to that found in the simulation at that time. I then pick at random 100 000 disc and 100 000 halo particles and, using their positions and velocities as initial conditions, I calculate their orbits for 40 bar rotations. Using spectral analysis ([9], [23]), I then find the principal frequencies of these orbits, i.e. the angular velocity Ω, the epicyclic frequency κ and the vertical frequency κz. An orbit is resonant if there are three integers l, m and n, such that Orbits on planar resonances fulfill lκ + mΩ + nκz = −ωR = mΩp lκ + mΩ = −ωR = mΩp (4) (5) The ILR corresponds to l = –1 and m = 2, CR to l = 0, and OLR to l = 1 and m = 2. The upper panels of Fig. 3 show, for time t = 800, the mass per unit frequency ratio MR of particles having a given value of the frequency ratio (Ω − Ωp)/κ as a function of this frequency ratio1. The distribution is not uniform, but has clear peaks at the location of the main resonances, as was already shown in [2] and [4]. The highest peak for the disc component is at the ILR, followed by (−1, 6) and CR. In all simulations with strong bars the ILR peak is strong. The existence of peaks at other resonances as well as their importance varies from one run to another and also during the evolution of a given run. For example the CR peak is, in many other simulations, much stronger than in the example shown here. For the spheroidal component the highest peak is at CR, followed by peaks at the ILR, OLR and (−1, 4). The lower panels show the angular momentum exchanged. For this I calcu- lated the angular momentum of each particle at time 800 and at time 500, as 1 See [4] for more information on MR and on how it is derived. Evolution and morphology of bars 7 Fig. 3. The upper panels give, for time t = 800, the mass per unit frequency ratio, MR, as a function of that ratio. The lower panels give ∆J, the angular momentum gained or lost by the particles between times 800 and 500, plotted as a function of their frequency ratio (Ω − Ωp)/κ, calculated at time t = 800. The left panels correspond to the disc component and the right ones to the halo. The component and the time are written in the upper left corner of each panel. The vertical dot-dashed lines give the positions of the main resonances. described in [4], and plotted the difference as a function of the frequency ratio of the particle at time 800. It is clear from the figure that disc particles at ILR and at the (−1, 6) resonance lose angular momentum, while those at CR gain it. There is a also a general, albeit small, loss of angular momentum from particles with frequencies between CR and ILR. This could be partly due to particles trapped around secondary resonances, and partly due to angular momentum 8 E. Athanassoula taken from particles which are neither resonant, nor near-resonant, but can still lose a small amount of angular momentum because the bar is growing. The cor- responding panel for the spheroid is, as expected, more noisy, but shows that particles at all resonances gain angular momentum. Thus this plot, and similar ones which I did for other simulations, confirm the analytical results of [4], and show that the linear results concerning the angular momentum gain or loss by resonant particles, qualitatively at least, carry over to the strongly nonlinear regime. 6 What determines the strength of bars and their slow-down rate? I have shown in the previous sections that the halo can take angular momentum from the bar, thus making it stronger and slower. However, for this effect to be important, the amount of angular momentum exchanged must be considerable. For the latter to happen the halo must • be sufficiently massive in the regions containing the principal resonances. • not be too hot, i.e. not have too high velocity dispersion. Indeed, hot haloes can not absorb much angular momentum, even if they are massive (e.g. [4]). Thus the length and the slow-down rate of bars are naturally limited by the mass and velocity distribution of the halo. Examples of this can be found in [4]. 7 Trends and correlations In [4] I found trends and correlations between the angular momentum absorbed by the spheroid (i.e. the halo plus, whenever existent, the bulge), the bar strength and the bar pattern speed. They are based on a set of simulations analogous to those described in the previous sections. Such plots are given also in Fig. 4, based on a somewhat larger sample of simulations. About three quarters of them were run with the Marseille GRAPE-5 systems, and roughly one quarter was run on PCs using Dehnen's treecode ([14], [15]). Each point represents one simulation and the trends are the same as those found in [4]. The upper panels show the results for the whole sample, the middle panels contain only simulations where the halo has a small core radius (γ < 2), Mh = 5 and does not extended beyond 15 disc scale-lengths, and the lower ones contain only simulations where the halo has a large core radius (γ > 2), Mh = 5 and again does not extended beyond 15 disc scale-lengths. The right panel shows that there is a correlation between the angular mo- mentum of the spheroid and the bar strength. This correlation holds also when I restrict myself to simulations with large (or small) core radii as seen in the second and third row. A trend also exists between the spheroid angular momentum and the bar pattern speed. In this case, however, simulations with large core radii be- have differently from those with small radii. Indeed, for simulations with a small Evolution and morphology of bars 9 Fig. 4. Relations between the bar strength and the pattern speed (left panels), the spheroid angular momentum and the pattern speed (middle panels) and the spheroid angular momentum and the bar strength (right panels), at times t = 800. The spheroid angular momentum is normalised by the initial disc angular momentum (Lz,d). The simulations under consideration in each panel are marked with a filled circle and the rest by a dot. The upper row includes all simulations, the middle and the lower ones subsamples, as described in the text. In the middle panel simulations with a bulge are marked with a ⊕, simulations with γ = 0.01 with a filled star, simulations with 2 > γ ≥ 1 with a filled triangle and simulations with Qinit ≥ 2 with an open square. In the lower panel simulations with Qinit ≥ 1.4 and z0 ≥ 0.2 are marked with an ⊕. core radius (i.e. centrally concentrated halos) I find a very strong correlation between the spheroid angular momentum and the pattern speed, particularly if I restrict myself to one value of γ. In such simulations the angular momentum is exchanged primarily between the bar region and the spheroid, thus accounting 10 E. Athanassoula for the very tight correlation. Simulations with large cores behave differently (lower middle panel). They show only a rough trend, except for simulations with a hot disc, which show a tight correlation. This is easily explained in the sce- nario of evolution via angular momentum exchange. The outer parts of hot discs absorb only little angular momentum, so that the exchange is basically between the bar region and the spheroid, thus accounting for the tight correlation. On the other hand, if the outer disc is cold, then it can participate more actively in the exchange. Since the angular momentum absorbed by the spheroid (plotted in Fig. 4) is not the total angular momentum exchanged, but only a fraction of it, I find only a trend. 8 Comparing the morphology of N -body and of real bars The correlations discussed in section 7 show clearly that models that have ex- changed more angular momentum have stronger bars than models that have exchanged little. By examining the results of the individual simulations, I could see that, in cases where large amounts of angular momentum have been ex- changed, the bars are long, relatively thin and have rectangular-like isodensities, particularly in their outer parts. A typical example of such a case is given in the left panels of Fig. 5 (see also [6]). Note also the existence of ansae at the ends of the bar, a feature sometimes observed in early type barred galaxies. On the other hand, models that have exchanged little angular momentum have less homogeneous properties. For example they can have either ovals, or short bars. Typical examples of such cases are given in the middle and right panels of Fig. 5, respectively. The model in the left panel has exchanged about 15 percent of the disc angular momentum by the time shown in Fig. 5, while the other two models only of the order of a percent. The edge-on morphology also is strongly influenced by the amount of angu- lar momentum exchanged. The strong bar, when seen edge-on, displays a clear peanut morphology, as often observed. On the other hand the oval has a boxy edge-on appearance, while the small bar has not changed significantly the edge- on morphology of the galaxy. The difference in bar strength is also illustrated in the lower panels of Fig. 5, which show the relative Fourier components of the density for m = 2, 4, 6 and 8 for the three simulations. The simulation that exchanged a lot of angular momentum has a very strong m = 2 component, with a secondary maximum roughly at the position of the ansae. The remaining components, even the m = 6 and 8 ones, are also important. The location of their maximum moves outwards with increasing m. The oval has much lower Fourier components, and only the m = 2 stands out from the noise. For small radii all components are nearly zero, which means that the oval must be nearly axisymmetric in its innermost parts. On the other hand, the m = 2 amplitude drops slowly with radius in the outer parts, thus extending to large radii. The small bar has Fourier components which drop rapidly with radius, i.e. they are noticeable only in the central region, as Evolution and morphology of bars 11 Fig. 5. Comparison of a simulation forming a strong bar (left panels), one forming an oval (middle panels) and one forming a short bar (right panels). The layout is as for Fig. 1. expected since the bar is confined there. Only the m = 2 and 4 components stand out from the noise. The radial rearrangement of the disc material due to the bar can be inferred by comparing the initial with the current circular velocity curves, given in the 12 E. Athanassoula Fig. 6. The upper panels show the run of the ellipticity 1 - b/a as a function of the semi-major axis a. The lower panels show the run of the shape parameter c, also as a function of a. The left panels corresponds to a model with a strong bar, the middle ones to model with an oval and the right ones to a model with a short bar. To improve the signal-to-noise ratio for the model with the oval I took an average over a time interval, namely [620-700]. The dispersion during that time is indicated by the error bars. The times are given in the upper right corner of the upper panels. first and second rows of panels. The strong bar has entailed a substantial radial rearrangement, the final disc mass distribution being considerably more cen- trally concentrated than the initial one. On the other hand, in the other two simulations, and particularly in the one producing the oval, there is very little radial rearrangement of the disc material. Since there is also hardly any radial rearangement of the halo material ([6], [3], [35]), this means that the disc-to-halo mass ratio changes most in the simulations where more angular momentum has been exchanged. Quantitative comparison of the bar form of the three models is given in Fig. 6. The values of the bar semi-major and semi-minor axes (a and b, respectively) and of the shape parameter (c) were obtained by fitting generalised ellipses of the form (x/a)c + (y/b)c = 1, (6) to the bar isodensities. The shape parameter c is 2 for ellipses, larger than 2 for rectangular-like generalised ellipses, and smaller than 2 for diamond-like ones. From this figure one can note that both the strong and the short bar are thin, and in general see how their axial ratios vary with the semi-major axis. The shape parameter is given in the lower panels. We see that both the strong and the short bar have rectangular-like isodensities in the outer regions of the bar, while the oval has a shape very close to elliptical. In fact the strong bar has axial Evolution and morphology of bars 13 ratios and shapes very similar to those found in [7] by applying the same type of analysis to a sample of early type barred galaxies. Plotting the run of the density along the bar major axis ([6]) I find for the strong bar a profile which is rather flat within the bar region, similar to what was found in [16] for early type bars. It is thus clear that the amount of angular momentum exchanged influences the morphology of the bar. In my first example, where a lot of angular momentum was transferred from the bar to the outer halo (mainly), the result is a long, strong bar, with some rectangular-like isophotes and ansae at its ends. The examples where little angular momentum was exchanged have a very different morphology, one forming an oval and the other a short bar. What determines which one of the two it will be? In the examples shown here, and in a rather large sample of similar cases, the oval was formed in an initially hot disc, while the short bar grew in a hot halo. The existing theoretical framework, however, gives no predictions on this point and work is in progress to elucidate this further. 9 Summary In this paper I reviewed evidence that shows that angular momentum is ex- changed between the bar region in the one hand, and the outer disc and the spheroid on the other. This exchange determines the slow-down rate of the bar, as well as its strength and its overall morphology. Acknowledgments. I thank the organisers for inviting me to this interesting conference. I thank M. Tagger, A. Bosma, W. Dehnen, A. Misiriotis, C. Heller, I. Shlosman, F. Masset, J. Sellwood, O. Valenzuela, A. Klypin and P. Teuben for many stimulating discussions. I thank J. C. Lambert for his help with the GRAPE software and with the administration of the simulations and W. Dehnen for making available to me his tree code and related programs. I also thank the INSU/CNRS, the University of Aix-Marseille I, the Region PACA and the IGRAP for funds to develop the GRAPE and Beowulf computing facilities used for the simulations discussed in this paper and for their analysis. References 1. E. Athanassoula: 'Evolution of bars in isolated and in interacting disk galaxies'. In Barred Galaxies, eds. R. Buta, D. Crocker and B. Elmegreen (Astron. Soc. of the Pacific Conference series, 91), pp. 309–321 (1996) 2. E. Athanassoula: Astrophys. J. 569, L83 (2002) 3. E. Athanassoula: 'Formation and Evolution of Bars in Disc Galaxies'. In Disks of Galaxies: Kinematics, Dynamics and Perturbations, eds. E. Athanassoula, A. Bosma, R. Mujica (Astron. Soc. of the Pacific Conference Series, 275) pp. 141–152 (2002) 4. E. Athanassoula: Mon. Not. R. Astron. Soc., in press (2003) 5. E. Athanassoula, L. Martinet: Astron. Astrophys. 87, L10 (1980) 6. E. Athanassoula, A. Misiriotis: Mon. Not. R. Astron. Soc. 330, 35 (2002) 14 E. Athanassoula 7. E. Athanassoula, S. Morin, H. Wozniak, et al.: Mon. Not. R. Astron. Soc. 245, 130 (1990) 8. E. Athanassoula, J. A. Sellwood: Mon. Not. R. Astron. Soc. 221, 213 (1986) 9. J. Binney, D. Spergel: Astrophys. J. 252, 308 (1982) 10. A. Bosma: Astron. J., 86, 1825 (1981) 11. G. de Vaucouleurs: Astrophys. J. Suppl. 8, 31 (1963) 12. V. P. Debattista, J. A. Sellwood: Astrophys. J. 493, L5 (1988) 13. V. P. Debattista, J. A. Sellwood: Astrophys. J. 543, 704 (2000) 14. W. Dehnen: Astrophys. J. 536, L39 (2000) 15. W. Dehnen: J. Comp. Phys. 179, 27 (2002) 16. B. G. Elmegreen, D. M. Elmegreen: Astrophys. J. 288, 438 (1985) 17. P. B. Eskridge, J. A. Frogel, R. W. Pogge, et al.: Astron. J., 119, 536 (2000) 18. P. Grosbøl, E. Pompei, P. Patsis: 'Spiral Structure Observed in Near-Infrared'. In Disks of Galaxies: Kinematics, Dynamics and Perturbations, eds. E. Athanassoula, A. Bosma, R. Mujica (Astron. Soc. of the Pacific Conference Series, 275) pp. 305– 310 (2002) 19. L. Hernquist: Astrophys. J. Suppl. 86, 389 (1993) 20. L. Hernquist, M. D. Weinberg: Astrophys. J. 400, 80 (1992) 21. F. Hohl: Astrophys. J. 168, 343 (1971) 22. A. Kawai, T. Fukushige, J. Makino, M. Taiji: Pub. Astron. Soc. Japan 152, 659 (2000) 23. J. Laskar: Icarus, 88, 266 (1990) 24. B. Little, R. G. Carlberg: Mon. Not. R. Astron. Soc. 250, 161 (1991a) 25. B. Little, R. G. Carlberg: Mon. Not. R. Astron. Soc. 251, 227 (1991b) 26. D. Lynden-Bell, A. J. Kalnajs: Mon. Not. R. Astron. Soc. 250, 161 (1972) 27. P. Martin: Astron. J. 109, 2428 (1995) 28. R. H. Miller, K. H. Prendergast, W. J. Quirk: Astrophys. J. 161, 903 (1979) 29. K. Ohta: 'Global Photometric Properties of Barred Galaxies'. In Barred Galaxies, eds. R. Buta, D. A. Crocker, B. G. Elmegreen (Astron. Soc. of the Pacific Conference Series, 91) pp. 37–43, (1996) 30. K. Ohta, M. Hamabe, K. Wakamatsu: Astrophys. J. 357, 71 (1990) 31. J. P. Ostriker, P. J. E. Peebles: Astrophys. J. 186, 467 (1973) 32. A. Toomre: Astrophys. J. 139, 1217 (1964) 33. A. Toomre: 'What amplifies the spirals' In: The Structure and Evolution of Normal Galaxies, ed. S. M. Fall, D. Lynden-Bell (Cambridge University Press) pp. 111–136 (1981) 34. S. Tremaine, M. D. Weinberg: Mon. Not. R. Astron. Soc. 209, 729 (1984) 35. O. Valenzuela, A. Klypin: in preparation (2003) 36. M. D. Weinberg: Mon. Not. R. Astron. Soc. 213, 451 (1985)
astro-ph/0203033
1
0203
2002-03-04T06:52:55
Maximum mass and radius of strange stars in the linear approximation of the EOS
[ "astro-ph" ]
The Chandrasekhar limit for strange stars described by a linear equation of state (describing quark matter with density-dependent quark masses) is evaluated. The maximum mass and radius of the star depend on the fundamental constants and on the energy density of the quark matter at zero pressure. By comparing the expression for the mass of the star with the limiting mass formula for a relativistic degenerate stellar configuration one can obtain an estimate of the mass of the strange quark.
astro-ph
astro-ph
A&A manuscript no. (will be inserted by hand later) Your thesaurus codes are: missing; you have not inserted them Maximum mass and radius of strange stars in the linear approximation of the EOS T. Harko and K. S. Cheng Department of Physics, The University of Hong Kong, Pok Fu Lam Road, Hong Kong, P. R. China Received 9 August 2001/Accepted 4 February 2002 Abstract. The Chandrasekhar limit for strange stars described by a linear equation of state (describing quark matter with density-dependent quark masses) is evaluated. The maximum mass and radius of the star depend on the fundamental constants and on the energy density of the quark matter at zero pressure. By comparing the expression for the mass of the star with the limiting mass formula for a relativistic degenerate stellar configuration one can obtain an estimate of the mass of the strange quark. Key words: dense matter-equation of state-stars: fundamental parameters 1. Introduction One of the most important characteristics of compact relativistic astrophysical objects is their maximum allowed mass. The maximum mass is crucial for distinguishing be- tween neutron stars and black holes in compact binaries and in determining the outcome of many astrophysical processes, including supernova collapse and the merger of binary neutron stars. The theoretical value of the maximum mass for white dwarfs and neutron 2 0 0 2 r a M 4 1 v 3 3 0 3 0 2 0 / h p - o r t s a : v i X r a stars was found by Chandrasekhar and Landau and is given by Mmax ∼ (cid:16) ¯hc (Shapiro & Teukolsky 1983), where mB is the mass of the baryons (in the case of white G m−4/3 B (cid:17)3/2 dwarfs, even pressure comes from electrons; most of the mass is in baryons). Thus, with the exception of composition-dependent numerical factors, the maximum mass of a de- generate star depends only on fundamental physical constants. The radius Rmax of the degenerate star obeys the condition Rmax ≤ ¯h either electron (white dwarfs) or neutron (neutron stars) (Shapiro & Teukolsky 1983). , with m being the mass of mc(cid:16) ¯hc B(cid:17)1/2 Gm2 White dwarfs are supported against of gravitational collapse by the degeneracy pressure of electrons whereas for neutron stars this pressure comes mainly from the nuclear force between nucleons (Shapiro & Teukolsky 1983). For non-rotating neutron stars with the 2 Harko & Cheng: Maximum Mass and Radius for Linear EOS central pressure at their center tending to the limiting value ρcc2, an upper bound of around 3M⊙ has been found (Rhoades & Ruffini 1974). The quark structure of the nucleons, suggested by quantum cromodynamics, opens the possibility of a hadron-quark phase transition at high densities and/or temper- atures, as suggested by Witten (Witten 1984). In the theories of strong interaction, quark bag models suppose that breaking of physical vacuum takes place inside hadrons. As a result, vacuum energy densities inside and outside a hadron become essen- tially different and the vacuum pressure on the bag wall equilibrates the pressure of quarks, thus stabilizing the system. If the hypothesis of the quark matter is true, then some neutron stars could actually be strange stars, built entirely of strange mat- ter (Alcock, Farhi & Olinto 1986, Haensel, Zdunik & Schaeffer 1986). For a review of strange star properties, see (Cheng, Dai & Lu 1998). Most of the investigations of quark star properties have been done within the frame- work of the so-called MIT bag model. Assuming that interactions of quarks and gluons are sufficiently small, neglecting quark masses and supposing that quarks are confined to the bag volume (in the case of a bare strange star, the boundary of the bag coincides with the stellar surface), the energy density ρc2 and pressure p of a quark-gluon plasma at temperature T and chemical potential µf (the subscript f denotes the various quark flavors u, d, s etc.) are related, in the MIT bag model, by the equation of state (EOS) (Cheng, Dai & Lu 1998) p = (ρ − 4B)c2 3 , (1) where B is the difference between the energy density of the perturbative and non- perturbative QCD vacuum (the bag constant). Equation (1) is essentially the equation of state of a gas of massless particles with corrections due to the QCD trace anomaly and perturbative interactions. These corrections are always negative, reducing the energy density at given temperature by about a factor of two (Farhi & Jaffe 1984). For quark stars obeying the bag model equation of state (1) the Chandrasekhar limit has been evaluated, from simple energy balance relations, in (Bannerjee, Ghosh & Raha 2000). In addition to fundamental constants, the maximum mass also depends on the bag constant. More sophisticated investigations of quark-gluon interactions have shown that Eq. (1) represents a limiting case of more general equations of state. For example, MIT bag models with massive strange quarks and lowest- order QCD interactions lead to some corrections terms in the equation of state of quark matter. Models incorporating restora- tion of chiral quark masses at high densities and giving absolutely stable strange matter can no longer be accurately described by using Eq. (1). On the other hand, in models in which quark interaction is described by an interquark potential originating from gluon exchange and by a density-dependent scalar potential which restores the chiral symmetry Harko & Cheng: Maximum Mass and Radius for Linear EOS 3 at high densities (Dey et al. 1998), the equation of state P = P (ρ) can be well approx- imated by a linear function in the energy density ρ (Gondek-Rosinska et al. 2000). It is interesting to note that Frieman and Olinto (Frieman & Olinto 1989) and Haensel and Zdunik (Haensel & Zdunik 1989) have already mentioned the approximation of the EOS by a linear function (see also Prakash et al. (Prakash, Baron & Prakash 1990), Lattimer et al (Lattimer et al. 1990). Recently Zdunik (Zdunik 2000) has studied the linear ap- proximation of the equation of state, obtaining all parameters of the EOS as polynomial functions of strange quark mass, the QCD coupling constant and bag constant. The scal- ing relations have been applied to the determination of the maximum frequency of a particle in a stable circular orbit around strange stars. It is the purpose of this paper to obtain, by using a simple phenomenological ap- proach (which is thermodynamical in its essence), the maximum mass and radius (the Chandrasekhar limits) for strange stars obeying a linear equation of state. Of course the maximum mass of compact astrophysical objects is a consequence of General Relativity and not of the character of motion of matter constituents. However, the formulae for maximum mass and radius, due to their simple analytical form, give a better insight into the underlying physics of quark stars, also allowing us to obtain some results which cannot be obtained by numerical methods. For example, from the obtained relations one can find the scaling relations for the maximum mass and radius of strange stars in a natural way. The present paper is organized as follows. The maximum mass and radius of quark stars with a general linear equation of state is obtained in Section 2. In Section 3 we discuss our results conclude the paper. 2. Maximum mass and radius for strange stars in the linear approximation of the EOS We assume that the strange star obeys an equation of state that can be obtained by interpolation with a linear function of density in the form: p = a (ρ − ρ0) c2, where a and ρ0 are non-negative constants. ρ0 is the energy density at zero pressure. (2) Such an equation of state has been proposed mainly to describe the strange matter built of u, d and s quarks (Gondek-Rosinska et al. 2000, Zdunik 2000). The physical consistency of the model requires ρ0 > 0. The particle number density and the chemical potential corresponding to EOS (2) are given respectively by (Zdunik 2000) n (p) = n0(cid:18)1 + a + 1 a p ρ0c2(cid:19)1/(a+1) , (3) 4 Harko & Cheng: Maximum Mass and Radius for Linear EOS µ (p) = µ0(cid:18)1 + a + 1 a p ρ0c2(cid:19)a/(a+1) , (4) where n0 is the particle number density at zero pressure and µ0 = ρ0c2/n0. The parameters a and ρ0 can be calculated, for realistic equations of state, by using a least squares fit method (Gondek-Rosinska et al. 2000, Zdunik 2000). For the equations of state incorporating restoration of chiral quark masses at high densities proposed in Dey et al. (Dey et al. 1998) one obtains the values a = 0.463, ρ0 = 1.15 × 1015g/cm3 and a = 0.455, ρ0 = 1.33 × 1015g/cm3, respectively (Gondek-Rosinska et al. 2000). The standard bag model corresponds to a = 0.333 and ρ0 = 4 × 1014g/cm3 (Cheng, Dai & Lu 1998). From Eqs. (3)-(4) it follows that the particle number and chemical potential are related by the equation µ = ρ0c2 na+1 0 na. (5) From the numerical studies of strange star models we know that the density profile of this type of astrophysical object is quite uniform (Glendenning 1996). Therefore we can approximate n ≈ N/V , which leads to µ µ0 R−3a, =(cid:18) 4π 3 (cid:19)−a(cid:18) N n0(cid:19)a (6) where N is the total number of particles in a star of radius R and volume V . With the use of Eqs. (2)-(6) one obtains the energy density of the star in the form ρ0 ρ = a ρ0. a + 1 R−3(a+1) + n0(cid:19)a+1 3 (cid:19)−a−1(cid:18) N a + 1(cid:18) 4π The total mass M of the star is defined according to M = 4πR R a + 1(cid:18) 4π 3 (cid:19)−a(cid:18) N n0(cid:19)a+1 R−3a + ρ0R3, 4π 3 a + 1 a is given by ρ0 M = 0 ρr2dr ≈ 4π 3 ρR3 and (7) (8) where we assumed that the energy density is approximately constant inside the star. Extremizing the mass with respect to the radius R by means of ∂M/∂R = 0 gives the relation ρ0 a + 1(cid:18) 4π 3 (cid:19)−a(cid:18) N n0(cid:19)a+1 R−3a = 4π 3 1 a + 1 ρ0R3. (9) Substituting Eq. (9) into Eq. (8) we obtain the maximum mass of the strange star in the linear approximation of the EOS: M = 4π 3 ρ0R3. (10) This expression is very similar to the expression for the maximum mass of the quark star obtained assuming that the star is composed of three-flavour masslesss quarks, confined in a large bag (Bannerjee, Ghosh & Raha 2000, Cheng & Harko 2000). From Harko & Cheng: Maximum Mass and Radius for Linear EOS 5 a physical point of view, Eq. (10) describes a uniform density zero pressure stellar type configuration. The maximum equilibrium radius corresponds to a minimum total energy of the star (including the gravitational one), for any radius. For ordinary compact stars, the mass is entirely due to baryons, and the corresponding (Newtonian) gravitational potential energy is of the order EG ∼ −αGM 2/R (α = −3/5 for constant density Newtonian stars). For quark stars, assumed to be formed of massless quarks, the total mass can be calculated from the total (thermodynamic as well as confinment) energy in the star. One possibility for the estimation of the gravitational energy per fermion is to define an effective quark mass incorporating all the energy contributions (Bannerjee, Ghosh & Raha 2000). The gravitational energy per particle (the strange star is assumed to be formed from fermions) is EG = − GM mef f R , (11) where mef f is the effective mass of the particles inside the star, incorporating also effects such as quark confinment. For a star with N particles one can write M = N mef f = ρ0V , or mef f = ρ0/n. On the other hand one can assume µ = ρ0/2n (Bannerjee, Ghosh & Raha 2000), leading to mef f = 2µ/c2. Hence, with the use of Eqs. (6), (9) and (10) we can express the gravitational energy per particle as 3 (cid:19)2 EG = −2(cid:18) 4π G ρ2 0 N R5. (12) The energy density per particle of the fermions follows from Eq. (7) and is given by: EF = 4π 3 1 a + 1 ρ0 N R3. The total energy E per particle is E = 4π 3 1 a + 1 ρ0 N 3 (cid:19)2 R3 − 2(cid:18) 4π G ρ2 0 N R5. (13) (14) Extremizing the total energy with respect to the radius (with the total particle number kept constant),(cid:0) ∂E configuration in the linear approximation of the EOS is given by: ∂R(cid:1)N =const. = 0, it follows that the maximum radius of the equilibrium Rmax = R0 pπ (a + 1) Gρ0 The maximum mass of the star can be calculated from Eq. (10) and is: c . Mmax = 4 3 R3 0 (a + 1)3/2 c3 G 1 √πGρ0 . (15) (16) In Eqs. (15) and (16) R0 is a numerical factor of the order R0 ≈ 0.474. The maximum radius of the quark star given by Eq. (15) is the radius corresponding to the maximum mass. On the other hand for the existing models of strange stars, 6 Harko & Cheng: Maximum Mass and Radius for Linear EOS the configuration with maximum mass has a radius which is lower than the maximum radius. For example, for strange stars described by the bag model equation of state, the maximum radius is 11.40km, while the radius corresponding to the maximum mass is 10.93km, which is 4% lower than the maximum radius. This difference is neglected in Eq. (15). The maximum mass and radius of the star are strongly dependent on the numerical value of the coefficient R0 and estimations based on other physical models could lead to different numerical estimates of the limiting values of the basic parameters of the static strange stars. 3. Discussions and final remarks In the present paper we have shown that there is a maximum mass and radius (the Chandrasekhar limits) for quark stars whose equation of state can be approximated by a linear function of the density. We have also obtained the explicit expressions for Mmax and Rmax. With respect to the scaling of the parameter ρ0 of the form ρ0 → kρ0, the maximum mass and radius have the following scaling behaviors: Rmax → k−1/2Rmax, Mmax → k−1/2Mmax. (17) For the maximum mass of the strange stars this scaling relation has also been found from the numerical study of the structure equations in the framework of the bag model (Witten 1984, Haensel, Zdunik & Schaeffer 1986). A rescaling of the parameter a of the form a + 1 → K (a + 1), with ρ0 unscaled, leads to a transformation of the radius and mass of the form Rmax → K −1/2Rmax, Mmax → K −3/2Mmax. (18) A simultaneous rescaling of both a and ρ0, with a + 1 → K (a + 1) , ρ0 → kρ0 gives (19) Rmax → k−1/2K −1/2Rmax, Mmax → k−1/2K −3/2Mmax. The maximum mass and radius of strange stars with linear EOS is strongly dependent on the numerical value of ρ0, the mass decreasing with increasing ρ0. For ρ0 = 4B, with the bag constant B = 1014g/cm3 (56M eV f m−3) we obtain Mmax = 1.83M⊙, a value that must be compared to the value Mmax = 2M⊙ obtained by numerical methods (Witten 1984, Haensel, Zdunik & Schaeffer 1986). The difference between the numerical and theoretical predictions is around 10%. For ρ0 = 1.33 × 1015gcm−3 the maximum mass of the star is about 1M⊙. Generally our formulae (15) and (16) underestimate the maximum values of the mass and radius because we have assumed that the density inside the star is uniform. It Harko & Cheng: Maximum Mass and Radius for Linear EOS 7 is obvious that near the surface the density is much lower than at the center of the compact object. Due to the approximations and simplifications used to derive the basic expressions, reflected mainly in the uncertainties in the exact value of the coefficient R0, Eqs. (15) and (16) cannot provide high precison numerical values of the maximum mass and radius for linear EOS stars, which must be obtained by numerically integrating the gravitational field equations. For the the linear EOS, Mmax and Rmax depend mainly on the fundamental constants c and G and on the zero pressure density ρ0 (the bag constant). The Chandrasekhar expressions for the same physical parameters involve two more fundamental constants, ¯h and the mass of the electron or neutron. For quark stars, usually one assume they are composed of a three-flavour system of massless quarks, confined in a large bag. Hence the mass of the quark cannot play any role in the mass formula. But the linear EOS with arbitrary a can describe quark matter with non-zero quark masses (the mass of the strange quark ms ≈ 200M eV ), forming a degenerate Fermi gas (Gondek-Rosinska et al. 2000, Zdunik 2000). Therefore this system should also be described by the same formulae as white dwarfs or neutron stars, not only by Eqs. (15) - (16). Generally ρ0 is a function of the mass of the strange quark, so this mass implicitly appears in the expression of the maximum mass and radius. But on the other hand we can assume that the Chandrasekhar limit also applies to quark stars with the baryon mass substituted by an effective quark mass mqef f , representing the minimum mass of the quark bubbles composing the star. Hence we must have (cid:18) ¯hc G m−4/3 qef f(cid:19)3/2 c3 G ∼ 1 √πGρ0 . (20) Eq. (20) leads to the following expression of the effective mass of the "elementary" quark bubble: mqef f ∼ ¯hρ1/3 c !3/4 0 . (21) The effective quark mass is determined only by elementary particle physics constants and is independent of G. From its construction mqef f should be relevant when the system is quantum mechanical and involves high velocities and energies. With respect to a scaling of the zero pressure density of the form ρ0 → kρ0, the effective quark mass has the scaling behavior mqef f → mqef f k1/4, similar to the scaling of the strange quark mass (Zdunik 2000). For ρ0 = 4 × 1014gcm−3 we obtain mqef f ∼ 3.63 × 10−25g ≈ 204M eV . For ρ0 = 1.33 × 1015gcm−3 Eq. (21) gives mqef f ∼ 4.9 × 10−25g ≈ 275M eV . The mass given by Eq. (21) can be considered as the minimum mass of the stable quark bubble. It is of the same order of magnitude as the mass ms of the strange quark. Therefore the 8 Harko & Cheng: Maximum Mass and Radius for Linear EOS Chandrasekhar limit applies also for quark stars if we take mqef f for the mass of the elementary constituent of the star . In the present paper we have considered the maximum mass and radius of strange stars in the linear approximation of the equation of state and the dependence of these quantities on the parameter a has been found. We have also pointed out the existence of scaling relations for the maximum radius of strange stars, an aspect that has not been mentioned in previous investigations (Witten 1984, Haensel, Zdunik & Schaeffer 1986, Bannerjee, Ghosh & Raha 2000, Zdunik 2000). Our formulae also lead to the transfor- mation relations for the maximum mass and radius of strange stars with respect to separate and simultaneous scaling of the parameters a and ρ0. On the other hand the possibility of estimation of the mass of the strange quark from general astrophysical considerations can perhaps give a better understanding of the deep connection between micro- and macro-physics. Acknowledgements. This work is partially supported by a RGC grant of Hong Kong Government and T.H. is supported by a studentship of the University of Hong Kong. The authors are very grateful to the anonymous referee whose comments helped to improve an earlier version of the manuscript. References Alcock C., Farhi E. & Olinto A. 1986, Astrophys. J., 310, 261 Bannerjee S., Ghosh S. K. & Raha S. 2000, J. Phys. G: Nucl. Part. Phys., 26, L1 Cheng K. S., Dai Z. G. & Lu T. 1998, Int. J. Mod. Phys. D, 7, 139 Cheng K. S. & Harko T. 2000, Phys. Rev. D, 62, 083001 Dey M., Bombacci I., Dey J., Ray S. & Samanta B. C. 1998, Phys. Lett. B, 438, 123 Farhi E. & Jaffe R. L. 1984, Phys. Rev. D, 30, 2379 Frieman J. A. & Olinto A. 1989, Nature, 341, 633 Glendenning N. K. 1996, Compact stars: nuclear physics, particle physics and general relativity, Springer, New York Gondek-Rosinska D., Bulik T., Zdunik L., Gourgoulhon E., Ray S., Dey J. & Dey M. 2000, Astron. Astrophys., 363, 1005 Haensel P., Zdunik J. L. & Schaeffer R. 1986, Astron. Astrophys., 160, 121 Haensel P. & Zdunik J. L. 1989, Nature, 340, 617 Lattimer J. M., Prakash M., Masaak D. & Yahil A. 1990, Astrophys. J., 355, 241 Prakash M., Baron E. & Prakash M. 1990, Phys. Lett. B, 243, 175 Rhoades C. E. & Ruffini R. 1974, Phys. Rev. Lett., 32, 324 Shapiro S. L. & Teukolsky S. A. 1983, Black Holes, White Dwarfs, and Neutron Stars, John Wiley & Sons, New York Witten E. 1984, Phys. Rev. D, 30, 272 Zdunik J. L. 2000, Astron. Astrophys., 359, 311
0806.1754
1
0806
2008-06-10T21:16:48
Identifying the Low Luminosity Population of Embedded Protostars in the c2d Observations of Clouds and Cores
[ "astro-ph" ]
We present a search for all embedded protostars with internal luminosities < 1 solar luminosity in the sample of nearby, low-mass star-forming regions surveyed by the Spitzer Space Telescope c2d Legacy Project. The internal luminosity (Lint) of a source is the luminosity of the central source and excludes luminosity arising from external heating. On average, the c2d data are sensitive to embedded protostars with Lint > 4E-3 (d/140 pc)^2 solar luminosities, a factor of 25 better than the sensitivity of IRAS to such objects. We present selection criteria used to identify candidates from the Spitzer data and examine complementary data to decide whether each candidate is truly an embedded protostar. We find a tight correlation between the 70 micron flux and internal luminosity of a protostar, an empirical result based on observations and two-dimensional radiative transfer models of protostars. We identify 50 embedded protostars with Lint < 1 solar luminosities; 15 have Lint < 0.1 solar luminosities. The intrinsic distribution of source luminosities increases to lower luminosities. While we find sources down to the above sensitivity limit, indicating that the distribution may extend to luminosities lower than probed by these observations, we are able to rule out a continued rise in the distribution below 0.1 solar luminosities. Between 75-85% of cores classified as starless prior to being observed by Spitzer remain starless to our luminosity sensitivity; the remaining 15-25% harbor low-luminosity, embedded protostars. We compile complete Spectral Energy Distributions for all 50 objects and calculate standard evolutionary signatures, and argue that these objects are inconsistent with the simplest picture of star formation wherein mass accretes from the core onto the protostar at a constant rate.
astro-ph
astro-ph
Accepted for publication in ApJS Identifying the Low Luminosity Population of Embedded Protostars in the c2d Observations of Clouds and Cores Michael M. Dunham1,2, Antonio Crapsi3,4, Neal J. Evans II1, Tyler L. Bourke5, Tracy L. Huard5, Philip C. Myers5, and Jens Kauffmann6,5 ABSTRACT We present the results of a search for all embedded protostars with internal luminosities ≤ 1.0 L⊙ in the full sample of nearby, low-mass star-forming regions surveyed by the Spitzer Space Telescope Legacy Project "From Molecular Cores to Planet Forming Disks" (c2d). The internal luminosity of a source, Lint, is the luminosity of the central source and excludes luminosity arising from external heating. On average, the Spitzer c2d data are sensitive to embedded protostars with Lint ≥ 4 × 10−3 (d/140 pc)2 L⊙, a factor of 25 better than the sensitivity of the Infrared Astronomical Satellite (IRAS) to such objects. We present a set of selection criteria used to identify candidates from the Spitzer data and examine complementary data to decide whether each candidate is truly an embedded pro- tostar. We find a tight correlation between the 70 µm flux and internal luminosity of a protostar, an empirical result based on both observations and detailed two- dimensional radiative transfer models of protostars. We identify 50 embedded protostars with Lint ≤ 1.0 L⊙; 15 have Lint ≤ 0.1 L⊙. The intrinsic distribu- tion of source luminosities increases to lower luminosities. While we find sources down to the above sensitivity limit, indicating that the distribution may extend to luminosities lower than probed by these observations, we are able to rule out a continued rise in the distribution below Lint = 0.1 L⊙. Between 75 −85% of cores 1Department of Astronomy, The University of Texas at Austin, 1 University Station, C1400, Austin, Texas 78712–0259 2E-mail: [email protected] 3Sterrewacht Leiden, Leiden University, P.O. Box 9513, 2300 RA Leiden, the Netherlands 4Observatorio Astron´omico Nacional (IGN), Alfonso XII, 3, E-28014 Madrid, Spain 5Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Cambridge, MA 02138 6Initiative in Innovative Computing at Harvard, 60 Oxford Street, Cambridge, MA 02138 – 2 – classified as starless prior to being observed by Spitzer remain starless to our lu- minosity sensitivity; the remaining 15 − 25% harbor low-luminosity, embedded protostars. We compile complete Spectral Energy Distributions for all 50 ob- jects and calculate standard evolutionary signatures (Lbol, Tbol, and Lbol/Lsmm), and argue that these objects are inconsistent with the simplest picture of star formation wherein mass accretes from the core onto the protostar at a constant rate. Subject headings: stars: formation - stars: low-mass, brown dwarfs 1. Introduction Recently, the Spitzer Space Telescope Legacy Project "From Molecular Cores to Planet Forming Disks" (c2d; Evans et al. 2003) completed an extensive 3.6−160 µm imaging survey of nearby, low-mass star-forming regions. One of the results to come out of this survey is the discovery of very low luminosity objects (VeLLOs; Young et al. 2004). If the internal luminosity of a source, Lint, is the total luminosity of the central protostar and circumstellar disk (if present), a VeLLO is defined to be an object embedded within a dense core with Lint ≤ 0.1 L⊙ (Di Francesco et al. 2007). The bolometric luminosity of an embedded protostar, an observable quantity that can be calculated by integrating over the full Spectral Energy Distribution (SED), is composed of both internal and external luminosity (Lbol = Lint +Lext). The external luminosity is usually that arising from heating of the circumstellar envelope by the Interstellar Radiation Field (ISRF), and will add, on average, a few tenths of a solar luminosity to Lbol (e.g., Evans et al. 2001). Thus, the distinction between Lbol and Lint is most relevant for embedded protostars with Lint . 1.0 L⊙, where the external luminosity can be a significant fraction of the observed Lbol. For VeLLOs, the external luminosity can dominate the observed Lbol. Radiative transfer modeling of the SEDs of embedded protostars, including both the emission from the envelope at submillimeter and millimeter wavelengths and the emission from the central source itself at infrared wavelengths, is required to decouple internal and external luminosities (e.g., Shirley et al. 2002; Young et al. 2004; Dunham et al. 2006). Several VeLLOs have been discovered in cores that were previously classified as starless prior to being observed by Spitzer. In fact, the very first starless core observed by c2d, L1014, was found to harbor a VeLLO with Lint ∼ 0.09 L⊙ (Young et al. 2004). This discovery reinforces that the known sample of embedded protostars is not complete. This sample has been assembled primarily by two methods: (1) searching for IRAS sources that are associated with dense cores and have colors consistent with those expected for embedded – 3 – protostars (e.g., Myers et al. 1987), and (2) identifying molecular outflows and radio point sources associated with dense cores indicating the presence of protostars too deeply embedded to detect with IRAS (e.g., Andr´e et al. 1993). Myers et al. (1987) found that the IRAS data could detect protostars with Lint & 0.1 (d/140 pc)2 L⊙, where d is the distance to the protostar, although this does not include the younger, more deeply embedded protostars that were only identified on a case-by-case basis by the second method. The regions surveyed by c2d with Spitzer are located at distances ranging from 125 − 500 pc. Even in the closest of these regions VeLLOs are likely to fall below the IRAS sensitivity limit. In the more distant regions, no protostars with Lint . 1 L⊙ would be detected. While some of these protostars might have been identified on a case-by-case basis as described above, the full sample of embedded protostars with Lint ≤ 1 L⊙ is clearly incomplete. Constructing a complete sample of embedded protostars with Lint ≤ 1 L⊙ is important for studies of low-mass star formation. Despite substantial progress in recent decades, the details of the physical processes regulating mass accretion from the envelope to the proto- star remain poorly understood (see McKee & Ostriker [2007] for a recent review). Several authors have attempted to constrain evolutionary models of the formation of low-mass stars by determining the observational signatures of these models and comparing them to the properties of known protostars (e.g., Myers et al. 1998; Young & Evans 2005). A result common to all such studies is a substantial population of protostars with luminosities below model predictions. An idea proposed to explain this discrepancy is that the mass accretion is episodic in nature and the protostars with the lowest luminosities are those observed in qui- escent accretion states (e.g., Kenyon & Hartmann 1995; Young & Evans 2005; Enoch 2007; Enoch et al. 2008a, in preparation). Theoretical studies have provided several mechanisms by which such a process may occur, such as material piling up in a circumstellar disk until gravitational instabilities drive angular momentum outward and mass inward in short-lived bursts (Vorobyov & Basu 2005, 2006). Alternatively, quasi-periodic magnetically driven out- flows in the envelope can cause mass accretion onto the protostar to occur in magnetically controlled bursts (Tassis & Mouschovias 2005). Indeed, the evidence for non-steady mass accretion in young protostellar systems still in the embedded phase is steadily growing (e.g., Hartmann & Kenyon 1985; Dunham et al. 2006; Acosta-Pulido et al. 2007; K´osp´al et al. 2007; Fedele et al. 2007). However, as the sample of embedded, low luminosity protostars is incomplete, the true nature of the discrepancy between evolutionary models and obser- vations of protostars is unknown. Future work devoted to assessing the validity of various models by comparing their predictions to the properties of known protostars depends on the existence of a sample that is as complete and unbiased as possible. The VeLLOs are a particularly interesting subset of embedded, low-luminosity proto- stars; in essence, they are an extreme case of the problem discussed above. To date, only – 4 – three VeLLOs have been studied in detail: L1014-IRS (Lint ∼ 0.09 L⊙; Young et al. 2004), L1521F-IRS (Lint ∼ 0.06 L⊙; Bourke et al. 2006), and IRAM 04191-IRS (Lint ∼ 0.08 L⊙; Dunham et al. 2006). Despite the fact that all three have similar internal luminosities, they differ greatly in envelope and outflow properties, as discussed by Bourke et al. (2006). IRAM 04191 drives a large, bipolar molecular outflow, features bright molecular line and dust continuum emission, and shows evidence for infall, depletion, and deuteration (Andr´e et al. 1999; Belloche et al. 2002). L1521F is also bright in molecular line and dust continuum emission and also shows evidence for infall, depletion, and deuteration (Crapsi et al. 2004), but the envelope is not as centrally condensed as IRAM 04191 and the presence of an outflow is uncertain (Crapsi et al. 2004; Bourke et al. 2006). L1014 does not show significant evi- dence for infall, depletion, or deuteration (Crapsi et al. 2005a), but it does drive a compact, weak molecular outflow detected only in interferometer observations (Bourke et al. 2005; Crapsi et al. 2005a). A systematic search for all VeLLOs in the regions surveyed by c2d will allow their properties to be examined in detail both on a case-by-case basis and as a class of objects. Identifying the complete sample of VeLLOs will also allow us to determine how many cores classified as starless prior to being observed by Spitzer truly are starless, a question with important implications for estimates of the lifetime of starless cores (e.g., Kirk et al. 2005). In this paper, we present the results of a search for all embedded protostars with Lint ≤ 1.0 L⊙ in the full c2d imaging dataset. Depending on the distance to each individual source, some will already have been detected by IRAS, while others will be new sources. We consider this work to be complementary to several related studies: A search by Kirk et al. (2007) for embedded protostars in 22 cores classified as starless prior to being observed by Spitzer ; a search by Jørgensen et al. (2007; 2008, in preparation) for all embedded protostars in Perseus and Ophiuchus, regardless of luminosity, conducted by combining Spitzer and SCUBA 850 µm dust continuum emission data; and a search by Enoch (2007) and Enoch et al. (2008a, in preparation) for all embedded objects in the Perseus, Ophiuchus, and Serpens molecular clouds, regardless of luminosity, conducted by combining Spitzer and Bolocam 1.1 mm dust continuum emission data. The key difference between the work presented here and the searches for embedded protostars listed above is that we do not start by identifying dense cores from their millimeter dust continuum emission and then look for associated Spitzer sources embedded within them. Instead, we develop a set of criteria to identify candidate embedded, low-luminosity protostars based on the 3.6 − 70 µm Spitzer data. This way, we are able to identify all of the candidates in the full c2d dataset, regardless of the availability and quality of millimeter wavelength observations for each region. Only after we identify all candidates based on Spitzer data alone do we turn to other observations to distinguish the objects of interest – 5 – from various contaminants masquerading in our sample. Our method identifies candidates for further examination once large-scale surveys of nearby star-forming regions are completed with SCUBA-2 (Ward-Thompson et al. 2007) and Herschel, and the method can easily be extended to search for embedded, low-luminosity protostars in the additional nearby, low- mass star-forming regions being surveyed by the Spitzer Gould Belt Legacy Project (L. Allen et al. 2008, in preparation). The organization of this paper is as follows: In §2, we provide a brief description of the c2d observations and data reduction, emphasizing those aspects relevant to this work. The criteria for identifying candidate embedded, low-luminosity protostars from the 3.6 − 70 µm Spitzer data are discussed in §3, along with the possibilities for estimating source internal luminosities directly from observable quantities. A general proof-of-concept demonstrating the validity of these criteria is given in §4. In §5, we discuss the contamination expected in the list of candidates, both from background extra-galactic sources and from more evolved Young Stellar Objects (YSOs) no longer embedded within their dense cores. We discuss the necessary requirements to prove that a candidate is truly an embedded protostar in §5.1, we apply these requirements to our candidate list in §5.2, and we discuss the difficulties in including regions lacking good quality 70 µm data in §5.3. We discuss several general results of this work in §6. Finally, we present our conclusions in §7. 2. Observations and Data Reduction The Spitzer c2d data have been published in several papers focusing on individual regions in the survey. We summarize here the observation strategy and data reduction method used by c2d; greater detail on individual regions can be found in the references given below. 2.1. Observations 2.1.1. Molecular Clouds The c2d project obtained complete 3.6 − 160 µm Spitzer maps of five nearby, large molecular clouds. The clouds (and approximate areas) surveyed include 1 deg.2 of Serpens (Harvey et al. 2006; 2007a; 2007b), 2.5 deg.2 total of the Lupus I, Lupus III, and Lupus – 6 – IV clouds1 (Chapman et al. 2007; Mer´ın et al. 2008), 1 deg.2 of Chamaeleon II (Young et al. 2005; Porras et al. 2006; Alcal´a et al. 2008), 7 deg.2 of Ophiuchus (Padgett et al. 2007), and 4 deg.2 of Perseus (Jørgensen et al. 2006; Rebull et al. 2007). These clouds were chosen to span a wide range of low-mass star-forming environments (Evans et al. 2003). We adopt distances of 260 ± 10, 150 ± 20, 200 ± 20, 150 ± 20, 178 ± 18, 125 ± 25, and 250 ± 50 pc for Serpens, Lupus I, Lupus III, Lupus IV, Chamaeleon II, Ophiuchus, and Perseus, respectively (see discussion of distances in the references listed above). The observations of Serpens, Lupus, Chamaeleon II, Ophiuchus, and Perseus were obtained in Program IDs (PIDs) 174, 175, 176, 177, and 178, respectively. Observations at 3.6, 4.5, 5.8, and 8.0 µm were obtained with the Infrared Array Cam- era (IRAC; Fazio et al. 2004). Each region was observed in two epochs to allow for the detection and removal of asteroids, and each region was observed twice per epoch with small dithers between the observations. The integration time per pointing was selected to be 12 s, providing an effective integration time per pixel of 10.4 s. This strategy results in four total observations per region (2 epochs of 2 dithers each), resulting in a total integration time of ∼ 42 s per pixel. All observations were obtained in the "High Dynamic Range" mode, which provides an additional short (0.6 s) exposure to enable photometry of bright sources saturated in the longer exposures. Observations at 24, 70, and 160 µm were obtained with the Multiband Imaging Pho- tometer for Spitzer (MIPS; Rieke et al. 2004). The observations consist of two epochs of fast scan maps with a spacing between adjacent scan legs of 240′′. The second epoch was offset by 125′′ from the first in the cross-scan direction to fill in the 70 µm sky coverage that would have otherwise been missed due to detector problems, and by 80′′ in the scan direction to minimize missing 160 µm data. These mapping parameters result in two epochs of observations at 24 µm and one epoch at 70 and 160 µm, with total integration times per pixel of 30, 15, and 3 s, respectively. Small gaps of missing coverage remain in the 160 µm maps. For technical reasons relating to the fact that the length of a MIPS scan map leg must be chosen from a fixed list, the regions observed by MIPS are often significantly larger than those observed by IRAC in order to ensure full coverage. The approximate areas quoted above are for the IRAC+MIPS overlap regions; in general, the areas observed by MIPS are a factor of 2 − 3 larger. 1For simplicity, we refer to the ensemble of these three portions of the Lupus Molecular Cloud Complex as simply "Lupus". – 7 – 2.1.2. Small Dense Cores In addition to the five large molecular clouds, c2d also obtained small maps of 82 regions containing 95 small, dense cores. These cores are selected on the basis of being nearby (within 500 pc), relatively small (generally less than 5′, which corresponds to ∼0.7 pc at a distance of 500 pc), relatively isolated, and showing evidence for dense gas and dust (Evans et al. 2003). The list of all 82 regions is given in Evans et al. (2007). A discussion of the overall results of the Spitzer observations of these cores and estimates of the distance to each core are presented in T. Huard et al. (2008, in preparation). Although "relative isolation" was one of the criteria used to select these cores, we note here that many are not truly isolated in the sense that they are often loosely associated (both in projection on the sky and in velocity) with larger complexes. Approximately 70% of these cores were classified as starless prior to the launch of Spitzer, based primarily on showing no association with IRAS sources (Evans et al. 2003 and references therein). Observations at 3.6, 4.5, 5.8, and 8.0 µm were obtained with IRAC in PID 139 in a manner nearly identical to the clouds. The main differences are that the maps are much smaller (often only a single 5 × 5′ frame or a small map < 30′ on a side), the number of epochs observed (one or two) depends on the location of each core relative to the ecliptic, and short-exposure HDR images were only obtained for cores expected to contain bright sources based on previous surveys. Cores observed in only one epoch featured 4 dithers instead of 2 so that all cores featured a total integration time per pixel identical to the clouds. Observations at 24 and 70 µm were obtained with MIPS in PID 139; 160 µm observa- tions were not obtained. Unlike the clouds, the MIPS observations of the core regions were obtained in the pointed photometry mode. The exposure time per pointing was 3 seconds at 24 µm and either 3 or 10 seconds at 70 µm, and small raster maps were obtained to match the IRAC maps. As a result, the core regions, unlike the clouds, do not feature significantly larger areas observed with MIPS than with IRAC. The exact integration time per pixel varies from region to region due to the differences in field and map size, but is generally 30 − 60 s at 24 µm and 50 − 100 s at 70 µm. 2.2. Data Reduction The IRAC and MIPS images were processed by the Spitzer Science Center (SSC), using their standard pipeline, version S13, to produce Basic Calibrated Data (BCD) images. These images were then improved by the c2d Legacy Project to correct for artifacts remaining in the BCD images. A complete description of the improvements made, as well as the source – 8 – extraction and band-merging process summarized below, is given in Evans et al. (2007). After correcting for artifacts, mosaics were produced using the MOPEX software pro- vided by the SSC. Photometry at 3.6 − 24 µm was obtained using c2dphot, a modified version of DOPHOT (Schechter et al. 1993) that utilizes a digitized rather than analytic point source profile to better match the real Spitzer data, as well as incorporating several other changes (a complete description of which are given in Harvey et al. 2006). Sources that were not detected in all 5 bands from 3.6 − 24 µm were band-filled, a process whereby the c2dphot source extractor was forced to obtain fluxes at the source position (known from the band(s) in which it was detected) in the bands for which the source was not detected in the original source extraction (see Evans et al. (2007) for details). Finally, photometry at 3.6 − 24 µm were band-merged into a final source catalog. Photometry at 70 µm was obtained using the SSC's MOPEX point-source fitting package on filtered BCD data for faint sources (. 2 Jy) and unfiltered BCD data for bright sources (& 2 Jy). Sources extracted at 70 µm were band-merged into the 3.6 − 24 µm catalogs described above, although this process is imperfect due to the significantly worse resolution at 70 µm (∼18′′) compared to 24 µm (∼6′′) and 3.6 − 8.0 µm (∼2′′) (see Evans et al. 2007 and §5.3). All of the above reduction and source extraction is part of the standard c2d pipeline. However, this pipeline does not include source extraction at 160 µm due to problems with diffuse emission and low spatial resolution. Thus, photometry at 160 µm for the five large clouds was obtained using the SSC's MOPEX point-source fitting package. The source extraction lists were then inspected and sources that appeared to be extended dust "clumps" with little or no central condensation were removed (T. Brooke, private communication). As a redundancy check, we also visually inspected all of the 160 µm images at the positions of the candidates identified in this work and performed aperture photometry on those that have associated 160 µm sources. The aperture sizes and sky annuli were chosen to allow us to use one of the standard aperture corrections for 160 µm point sources determined by the SSC. In general, we find very good agreement between the PSF and aperture photometry. The ratio of PSF to aperture flux for sources extracted by both methods has a mean and median of 1.07 and 1.02, respectively, with a standard deviation of 0.2. We thus add an overall 20% systematic uncertainty into our photometry at 160 µm. Both methods assume all 160 µm sources are point sources; fluxes of extended sources may thus be under-estimated. – 9 – 3. Identification of Candidate Embedded, Low-Luminosity Protostars Generally speaking, we identify candidate embedded, low-luminosity protostars from the Spitzer observations by selecting sources with rising fluxes from shorter to longer wavelengths and infrared luminosities LIR, calculated over the 2MASS and Spitzer bands (1.25 − 70 µm), indicating Lint ≤ 1 L⊙ (see below). Throughout this paper we denote flux as Fν (Fν = νSν). When we refer to the shape of the Spectral Energy Distribution (SED) of an object, it is always evaluated in terms of flux (Fν) rather than flux density (Sν). At this stage, we emphasize completeness over reliability; many objects identified as candidates are likely either more evolved Young Stellar Objects (YSOs) no longer embedded in their dense cores, or background galaxies. Separating these from the true embedded protostars is addressed in §5. We have assembled a set of seven criteria that must be met for a source to be identified as a candidate. We list these criteria below, followed by an explanation of each one individually in §3.1. We provide a detailed discussion of the motivation for the third criterion in §3.2, and investigate the results of relaxing the first criterion in §3.3. Finally, we discuss the need to visually inspect each candidate in §3.4 and present a general proof-of-concept in §4. The seven criteria that must be met are: 1. Detected at 24 µm with a signal-to-noise ratio (SNR) ≥ 3. 2. Located at a position observed by a minimum of two Spitzer photometric bands be- tween 3.6 and 70 µm. 3. Detected at 70 µm, unless confused or not observed at this wavelength. 4. Rising (or flat) SED between the longest IRAC wavelength at which the source is detected and 24 µm. 5. Rising (or flat) SED between 24 and 70 µm, unless confused or not observed at 70 µm. 6. Observed luminosity calculated using all detections ranging from 1.25 − 70 µm of LIR ≤ 0.5 L⊙. 7. Not classified as a candidate galaxy in the classification method of Harvey et al. (2007b) unless Log(Pgal) ≤ −1.25, where Pgal is the un-normalized probability of being a galaxy. – 10 – 3.1. Selection Criteria We impose the first criterion simply because 24 µm observations are convenient for determining whether or not a source features a rising SED in the wavelength region covered by our Spitzer observations (3.6 − 70 µm). First of all, 24 µm is located approximately halfway between 3.6 and 70 µm in log space, making it useful for determining the general shape of the SED over the wavelengths observed by Spitzer. Additionally, the diffraction limited resolution of Spitzer, θ = 1.2 λ D , is a factor of 2.92 better at 24 µm than 70 µm. Thus, sources clustered close together (at least in projection on the sky) can often be resolved into individual sources at 24 µm even when they are confused at 70 µm, providing reliable flux information at a wavelength beyond the 3.6 − 8.0 µm region covered by IRAC and better enabling us to evaluate the shape of the SED. We discuss the effects of relaxing this criterion in §3.3. The second criterion ensures that we have flux information available at a minimum of one other Spitzer wavelength besides 24 µm. The MIPS observations of the five large clouds were obtained in the scan map observing mode and often cover significantly larger areas than the IRAC observations of the same regions for technical reasons. Furthermore, there are minor differences between the exact areas observed at 24 and 70 µm, resulting in small regions that are observed at 24 µm but not at any other wavelength. Since it is impossible to determine the shape of the SED of any source observed at only one wavelength, all sources in these "MIPS-24 only" regions are removed from consideration. This issue is not as significant in the regions with dense cores, where the MIPS data were obtained in the pointed observations mode and often cover very similar areas to the IRAC observations. The third criterion is imposed because 70 µm is a crucial wavelength for determining Lint of embedded, low-luminosity protostars. We will return to this point in greater detail below. The fourth and fifth criteria select sources with rising or flat SEDs from shorter to longer wavelengths. Simple, 1-D models of protostars still embedded within their cores show that they should essentially always feature rising SEDs both between the IRAC wavelength regime (3.6 − 8.0 µm) and 24 µm and between 24 and 70 µm (e.g., Young et al. 2005). Two- dimensional models that include the effects of outflow cavities show that the exact shape of the 3.6 − 70 µm SED depends on the inclination of the source (e.g., Whitney et al. 2003a; Whitney et al. 2003b; Robitaille et al. 2006). In particular, embedded protostars observed at nearly pole-on inclinations exhibit much flatter SEDs over these wavelengths since more of the protostellar emission can escape through the outflow cavities instead of being reprocessed to longer-wavelength emission by the envelope. While this flattening of the SED can be very significant, even at the most extreme inclinations the models predict essentially flat rather – 11 – than falling SEDs at these wavelengths. We return to this point in §4.2. We thus require each source to have an increasing or constant flux between the longest IRAC wavelength at which the source is detected and 24 µm, and between 24 and 70 µm. If a source is not detected at any IRAC wavelength, it is considered to pass the fourth criterion. We take the photometric uncertainties into account; a source may actually have a decreasing flux between the two wavelengths considered as long as it is consistent with being at least flat given the uncertainties. Not all sources have 70 µm detections; we will return to these sources at the end of this section, but for now, we keep any source that meets the first rising criterion but has no flux information at 70 µm. We do not place any rising requirements on the IRAC fluxes themselves since the effects of geometry (Whitney et al. 2003a; Whitney et al. 2003b), solid-state ice features (Boogert et al. 2004), and shocked emission from outflows (e.g., Noriega-Crespo et al. 2004) greatly complicate the exact shape of the 3.6 − 8.0 µm SED. The sixth criterion separates out the low-luminosity sources, those of interest to this study, from the more luminous sources. As described in §1, the internal luminosity of an embedded, low-luminosity protostar is not an observable quantity and is usually derived from radiative transfer modeling. Here we investigate whether or not we can estimate Lint without the need to construct detailed models of each source. We have identified 11 embedded protostars from the literature that have been observed with Spitzer by c2d and have accurate internal luminosity determinations from radiative transfer models. We list these 11 objects in Table 1, along with LIR calculated from the observations and Lbol and Lint derived from the published models. Figure 1a plots Lint vs. LIR in log-log space for these 11 objects; a linear correlation is clearly observed. The solid line shows the results of a linear least-squares fit; it has a slope of 0.88 and a y-intercept of 0.32. However, any linear correlation in a luminosity-luminosity plot is suspect since it may simply arise from plotting the square of the distance to the object versus itself. Thus, Figure 1b plots the ratio of Lint/LIR, a quantity that is independent of distance, vs. the log of LIR. An approximately constant ratio is observed over several orders of magnitude of LIR, as would be expected if a true linear correlation exists between these quantities. The solid line shows the average ratio of 1.7, weighted by the uncertainties in Lint. Inverting this quantity, we find that the luminosity calculated between 1.25 − 70 µm gives a result that is approximately half the true value of the internal luminosity. Thus, to select sources with Lint . 1.0 L⊙, we impose the requirement that LIR ≤ 0.5 L⊙. We emphasize that this relationship between Lint and LIR is only an approximation; Figure 1b clearly shows there is some variation from source to source. In fact, the lowest-luminosity sources may have a larger ratio of Lint/LIR than more luminous sources, although the uncertainties – 12 – are large. However, even if the ratio of Lint/LIR does increase for lower-luminosity sources, they will have LIR far below the cut of 0.5 L⊙ and will thus be included in our sample. The seventh criterion is related to the fact that background galaxies are difficult to separate from YSOs when analyzing Spitzer data of star-forming regions. A number of authors have recently shown that such galaxies often exhibit similar mid-infrared colors and SEDs to YSOs (e.g., Harvey et al. 2006; Jørgensen et al. 2006). In an effort to accurately separate out background galaxies from YSOs, Harvey et al. (2007b) developed a classification method that assigns to each object that can not be fitted by an extincted stellar photosphere an un-normalized "probability" of being a galaxy, Pgal. The probabilities are assigned based on the position of each source in various color-color and color-magnitude diagrams, along with information about the nature of the source (point-like vs. extended) given by the c2d source extraction software (see §2 and references therein). Based on its value of Pgal, Harvey et al. classified each source as a candidate YSO or candidate Galaxy. Figure 2 shows a histogram of Pgal for all 851 sources assigned a probability in the ensemble of 82 regions with dense cores observed by c2d. The dotted line shows Log(Pgal) = −1.47 while the dashed line shows Log(Pgal) = −1.25. This figure is similar to Figure 4 of Harvey et al. (2007b), which plots the distribution of Pgal for all sources assigned a probability in the c2d observations of the Serpens Molecular Cloud. As noted by Harvey et al., there are two distinct peaks in probability space: one at Log(Pgal) = −5.00 and one at Log(Pgal) ∼ −0.4, with a tail connecting the two. In a similar analysis on a Spitzer dataset of extragalactic objects, Harvey et al. found only the peak at Log(Pgal) ∼ −0.4 and not the other peak or the tail connecting the two. Based on this, they set the boundary between candidate YSO and candidate galaxy at Log(Pgal) = −1.47, such that objects in the peak at Log(Pgal) = −5.00 and the tail between the two peaks are considered candidate YSOs while objects in the peak at Log(Pgal) ∼ −0.4 are considered candidate Galaxies. However, as is evident both from Figure 2 of this paper and Figure 4 of Harvey et. al, there is likely some overlap in probability space between the tail of objects considered candidate YSOs and the peak of objects considered candidate galaxies. Some of the objects near the boundary of Log(Pgal) = −1.47, but slightly into the regime of galaxies, may in fact still be YSOs. Furthermore, there was no galaxy in the Spitzer dataset of extragalactic objects considered by Harvey et al. with Log(Pgal) ≤ −1.25. Thus, for all the objects that pass the first six criteria, we reject all those classified as candidate galaxies based on this classification method except for those with −1.47 ≤ Log(Pgal) ≤ −1.25. Many of these intermediate objects are likely to be galaxies, but we include them as candidates at this stage for the sake of completeness. Separating out these false candidates is the subject of §5. – 13 – 3.2. Importance of 70 µm Data We now return to the issue of detection at 70 µm. This is a crucial wavelength for determining Lint for embedded, low-luminosity protostars. Radiative transfer models of these objects are strongly constrained by this wavelength since the flux at 70 µm, F70, is mostly unaffected by the details of the source geometry and presence or absence of a circumstellar disk that significantly affect the 3.6 − 24 µm SED. F70 is also quite unaffected by the amount of external heating from the ISRF that is important at wavelengths & 100 µm (e.g., Dunham et al. 2006). To examine this more fully, we again use the sources listed in Table 1 observed with Spitzer by c2d and possessing accurate determinations of Lint through radiative transfer models. Figure 3 plots Fν, normalized to 140 pc, for all six wavelengths observed by Spitzer vs. Lint for these 11 sources, along with the results of linear least-squares fits to the data in log-log space. These fits show that a correlation exists between the flux at each wavelength and Lint; in general terms, more luminous sources emit more energy at all wavelengths. However, the correlation is clearly strongest at 70 µm, weaker at 24 µm, and very poor at 3.6 − 8.0 µm. To quantify this, at each wavelength we calculate χ2 r, a reduced chi-squared for the fit: χ2 r = 1 n − p n Xi=0 (yi − mxi − b)2 σ2 i , (1) where n is the number of datapoints (n = 11), p is the number of free parameters (p = 2), yi and xi are the log of the x and y values for each datapoint, σi is the size of the uncertainty in log space, and m and b are the slope and y-intercept, respectively, derived from the fit in log-log space. The values for m and b, along with their statistical uncertainties from the fit and the results of the χ2 r calculations, are presented in Table 2. We find χ2 r = 3 at 70 µm, χ2 r = 85 at 24 µm, and χ2 r > 100 for all four IRAC wavelengths, in agreement with the qualitative conclusions from above. These results suggest that a direct correlation exists between F70 and Lint. However, this analysis is based on only 11 data points, and furthermore, many of the radiative transfer model determinations of Lint listed in Table 1 are based on the same simple, 1-D modeling setup that, among other things, uses a very simple method of including a disk in a 1-D model (Butner et al. 1994). We do not expect this to be significant, as previous work has concluded that even these simple, 1-D models accurately constrain Lint (e.g., Dunham et al. 2006), but to test the validity of these results, we attempt to reproduce them by running a grid of two-dimensional models that are more physically realistic than those on which many of the values of Lint in Table 1 are based. – 14 – We used the two-dimensional, axisymmetric, Monte Carlo dust radiative transfer code RADMC (Dullemond & Dominik 2004) to compute the radiative transfer and calculate the emergent SEDs of embedded, low-luminosity protostars and look for correlations between Lint and Fν in the Spitzer bands. The density structure assumed for the models consists of a protostellar disk and a rotationally flattened, infalling envelope with an outflow cavity, and is described in greater detail in Crapsi et al. (2008). The details of the models are nearly identical to those of Crapsi et al.; we do not describe them here except to note that they have been tuned to their specific purpose within this work in the following ways: • The protostar has a fixed temperature of 3000 K and a random luminosity in the range of 0.03 − 10 L⊙. • The disk has a fixed size of 100 AU, variable mass between 10−3 and 10−5 M⊙, and fixed flaring parameters (a scale height H of 20 AU at the outer radius and then decreasing as H ∝ r9/7, corresponding to the self-irradiated passive disk of Chiang & Goldreich [1997]). • The envelope has a fixed radius of 14,000 AU, a variable mass in the range of 1 − 10 M⊙, and a variable centrifugal radius in the range of 100 − 900 AU. • The envelope is externally heated by the ISRF, for which we adopt that of Black (1994), modified at the ultraviolet wavelengths to reproduce the ISRF of Draine (1978). This Black-Draine ISRF is then attenuated by dust with properties given by Draine & Lee (1984) to simulate being embedded in a parent cloud and multiplied by a scale factor to account for the fact that the ISRF is not uniform everywhere, where the number of magnitudes of visual extinction and scale factor are chosen randomly. The observed SED was obtained by raytracing the density, temperature, and scattering structures calculated by RADMC at five different inclinations: 20, 35, 50, 65, and 80◦ from the axis of symmetry. Fluxes were integrated inside apertures comparable to the resolution of Spitzer : 2′′ for wavelengths shorter than 10 µm, 6′′ for wavelengths in the range of 10 − 40 µm, and 20′′ in the wavelength range of 40 − 100 µm. Following this description, a grid of 292 models were run and 1460 SEDs were obtained (one at each inclination). The results are shown in Figure 4 and, qualitatively, appear to confirm the correlation between F70 and Lint. The parameters derived from linear least- squares fits in log-log space for Fν vs. Lint for the models are given in Table 2. It is clear from these models that there is in fact a tight correlation between F70 and Lint (most of the scatter seen in the MIPS2 panel of Figure 4 is due to inclination effects). In fact, not only do – 15 – the observations and models both show this correlation, but they also agree remarkably well on the quantitative relationship (within 3% for the slope and within 1% for the y-intercept) considering that, to date, only 11 sources have been modeled by us and are thus available to examine this from an observational standpoint. Since F70 is an observable quantity, the correlation we find between F70 and Lint can be used to estimate Lint for sources which either lack sufficient data to constrain radiative transfer models or have such data but have not yet been modeled. Using the results of the linear least-square fit to the modeled F70 vs. Lint in log-log space, we derive this estimate to be Lint = 3.3 × 108 F70 0.94 L⊙ , (2) where F70 is in cgs units (erg cm−2 s−1) and is normalized to 140 pc. Since a detection at 70 µm is crucial for constraining radiative transfer model determi- nations of Lint and gives an immediate, direct estimate of Lint, we require each source to be detected at 70 µm unless it is not observed at this wavelength or is located too close to another source detected at 70 µm to be resolved into a separate source (see §5.3). Thus, with these two exceptions, we effectively require candidates to be above the detection threshold at 70 µm. This requirement sets the fundamental limit for our luminosity sensitivity. Fig- ure 5 plots a histogram showing the 70 µm 3σ point source sensitivity for each of the 82 regions with dense cores observed by c2d. This sensitivity was derived by first calculating the standard deviation of the background intensity for each region, σsky, in the image units of MJy sr−1. Approximating the 70 µm Spitzer PSF as a Gaussian with θFWHM equal to the diffraction limited resolution of ∼ 18′′, this background intensity is then converted into units of mJy beam −1 as follows: σsky (mJy beam)−1 = 1 × 109 (cid:18) π θ2 FWHM 4 ln2 (cid:19) σsky (MJy sr)−1 (3) This gives the 3σ point source sensitivity for each region, since the total flux density of a point source is equivalent to its flux density per beam. This 70 µm point source sensitivity can be directly translated into a luminosity sensi- tivity using Equation 2. From this relationship, the mean 70 µm 3σ point source sensitivity of 38.6 mJy translates into a luminosity sensitivity of ∼ 4 × 10−3 L⊙ at 140 pc. Figure 6 plots the luminosity sensitivity limit for each of the 82 regions vs. the distance to the region, calculated by translating the 70 µm 3σ point source sensitivity into a luminosity sensitivity using Equation 2 and then scaling from 140 pc to the distance to the region. The solid line shows the relation Lint = 4 × 10−3 (d/140 pc)2 L⊙. There is some scatter between the line and the actual data, caused by the variation in 70 µm point-source sensitivity from one region to the next, but this relation clearly provides an accurate estimate of the luminosity – 16 – sensitivity. On average, the c2d observations of regions with dense cores are sensitive to embedded protostars with internal luminosities Lint ≥ 4 × 10−3 (d/140 pc)2 L⊙. This same argument holds true for the c2d observations of the clouds, as the 70 µm cloud maps reach similar depths as the cores. This is a factor of 25 better than the sensitivity of IRAS quoted in §1. Applying these criteria results in the identification of 673 candidate embedded, low- luminosity protostars: 106 in the ensemble of dense core regions, 196 in Perseus, 112 in Chamaeleon II, 153 in Lupus, 57 in Ophiuchus, and 49 in Serpens. 3.3. Relaxing the First Criterion The first of the seven criteria described above for selecting candidates from the Spitzer c2d observations is a detection at 24 µm. As discussed in §3.1, this is an important wavelength for determining whether or not a source features a rising (or flat) SED from 3.6 − 70 µm. However, it is possible that a source of interest has no detection at 24 µm. For example, it might simply not be covered at 24 µm if it is located close to the edge of an observed region where the coverage between bands is not uniform, or it might be so deeply embedded that it is only detected at 70 µm. The latter could be particularly interesting sources, either very young, deeply embedded protostars or perhaps even first hydrostatic cores, which are short-lived, hydrostatic objects in a phase between the initial collapse of a dense core and the formation of a Class 0 protostar predicted by theory but not yet found by observations (e.g., Boss & Yorke 1995). Thus, to make sure no sources are missed by imposing this first criteron, we search for any source not detected at 24 µm but detected at 70 µm. We then apply the sixth criterion, LIR ≤ 0.5 L⊙, to select the low luminosity sources that are the focus of this work. This gives an additional 53 candidates: 13 in the ensemble of dense core regions, 7 in Perseus, 17 in Chamaeleon II, 10 in Lupus, 2 in Ophiuchus, and 4 in Serpens. We add these to the candidates identified above for a total of 726 candidates. However, only 4 of these 53 additional candidates survive the visual inspection described in §3.4, and none of these surviving 4 are particularly strong candidates (see §3.4). 3.4. Visual Inspection The final step in the identification of candidates is visual inspection. For each of the 726 sources identified above, we examined the images at the six wavelengths observed by – 17 – Spitzer and the source SED. Throughout the course of this inspection, we found five reasons to reject candidates: (1) The source is obviously a resolved galaxy; (2) The 70 µm detection is not real; (3) The source SED is not consistent with that of an embedded protostar; (4) The source is located in a region of nebulosity that produces a false infrared excess at the longer wavelengths; and (5) There is no real source. The first reason is due to the fact that a source must meet certain requirements to be classified as a candidate YSO or candidate galaxy in the method of Harvey et al. (2007b). Specifically a source must be detected in both epochs of observations (or in the single epoch if only one epoch was observed) with SNR ≥ 3 in all 4 IRAC bands and the 24 µm MIPS band. Any source outside the area of overlap between all 5 bands, or simply not detected in any one of these bands, does not get evaluated by the method of Harvey et al. and thus can not be classified as a candidate YSO or galaxy. This results in many galaxies making the candidate list. Visual inspection removes those large and/or close enough to be resolved by IRAC; all others remain candidates at this stage. The second reason relates to the fact that the exact sensitivity of the 70 µm c2d ob- servations is a complicated function of the position within the scan map (clouds) or pointed observation (core regions), since the exact number of frames at a given position depends on the technical details of how Spitzer executes observations. In regions with fewer frames, often near the edges of maps, noise spikes are sometimes falsely identified as sources; we remove these from consideration. Of the 53 candidates identified by searching for sources detected at 70 µm but not at 24 µm, 49 were removed for this reason, and the 4 that remain are all sources with questionable 70 µm detections that simply weren't obvious enough false detections to remove. The third reason relates to the details of how rising sources were selected. Figure 7 shows the SED of SSTc2d J032856.64+311835.6, a source in Perseus representative of those removed for this reason. It is detected at all 6 Spitzer wavelengths, features a rising flux between 8 and 24 µm, features a flat, but slightly rising, flux between 24 and 70 µm, has LIR = 0.37 L⊙, and is classified as a candidate YSO by the classification method of Harvey et al. (2007b). Thus, it fulfills all the criteria for selecting candidates discussed in §3.1. However, it clearly emits more energy in the near-infrared (1.25 − 2.17 µm) than at 24 and 70 µm. Comparison with 2-D models of sources (e.g., Whitney et al. 2003a; Whitney et al. 2003b; Robitaille et al. 2006) shows that, even at the most extreme pole-on inclinations, the effects of outflow cavities are unlikely to result in an observed SED more luminous in the near-infrared than at 24 and 70 µm. A more likely explanation for objects featuring such SEDs is that they are more evolved YSOs surrounded by circumstellar disks but no longer – 18 – embedded within dense cores. Indeed, a search of the SIMBAD2 database shows that this is SSS 108, which Aspin et al. (1994) and Aspin (2003) conclude is a pre-main sequence star with a small thermal excess at 2 µm (i.e., from a disk). The nature of these objects with large infrared excesses at 24 and 70 µm compared to 3.6 − 8.0 µm could potentially be very interesting; indeed, these SEDs appear similar to those of a sample of YSOs with disks featuring large inner holes presented by Brown et al. (2007). Sources that meet our selection criteria but are more luminous in the near-infrared than at 24 and 70 µm may be interesting objects, but they are not relevant to this study and are thus removed from our sample. The final two reasons for rejecting sources upon visual inspection both relate to the bandfilling process described in §2. The first reason is that stars are often detected only in the first two IRAC bands, due both to the decreasing Spitzer sensitivity at longer wavelengths and the fact that Spitzer observes the Rayleigh-Jeans portion of stellar SEDs. Some stars detected only in these two bands are located, at least in projection, in regions of nebulosity. For these objects, the bandfilling process can assign source fluxes in bands beyond IRAC2 that are contaminated by this extended nebulosity, creating false excesses and thus falsely rising SEDs. These are obvious in that there is no real point-source in the longer wavelength images. The second reason is that Spitzer images of star-forming regions often exhibit copious amounts of nebulous, diffuse emission due to scattered light, shocked emisison from outflows, thermal emission from hot dust, PAH emission, etc. In regions where this diffuse emission becomes clumpy, the source extraction may falsely detect a point-source at one wavelength that then gets band-filled at all other wavelengths. Some of these fake sources may pass all 7 criteria for selection, but they are obvious upon visual inspection in that there is no point source present in the images. Both types of sources are removed from consideration. After removing sources for the reasons described above, we are left with 218 candidates: 49 in the ensemble of dense core regions, 70 in Perseus, 16 in Chamaeleon II, 42 in Lupus, 22 in Ophiuchus, and 19 in Serpens. The source number from this work, group number indicating likelihood of being an embedded object (see §5), Spitzer source name, position, LIR, and distance to and name of the nearest source in the IRAS point-source catalog are given for each candidate in Tables 3 − 8. For the candidates in the dense cores, Table 3 also list the name of the core region in which each candidate is located and the assumed distance to this region; these distances have been derived by reddening or association with regions with known distances (T. Bourke, private communication). The full list of distances to core regions will be presented by T. Huard et al. (2008, in preparation). We emphasize 2Available at: http://simbad.u-strasbg.fr/simbad/ – 19 – here that these objects are only candidates. We have chosen to emphasize completeness over reliability, with the result being that many of these 218 candidates are in fact galaxies or more evolved YSOs no longer embedded in their dense cores. Separating the wheat from the chaff is the focus of §5. 4. Proof of Concept In §3.1, we presented a set of criteria to select embedded, low-luminosity protostars from Spitzer c2d observations. We imposed certain restrictions on the shape of the 3.6 − 70 µm SED in order to pick out the embedded objects (the fourth and fifth criteria), and justified their validity by comparing to both 1-D and 2-D models of embedded sources. Here we further examine the ability of these criteria to identify embedded protostars. 4.1. Comparison to Known Class 0 Objects Class 0 objects were added to the original infrared spectral slope classification system of Lada (1987) by Andr´e et al. (1993) as objects younger and more deeply embedded than Class I objects and thus not detected in the infrared. However, with the advent of sensitive infrared facilities such as Spitzer, both Class 0 and I objects are detected in the infrared and are often indistinguishable from one another based on their infrared spectral slope α (e.g., Enoch 2007; Enoch et al. 2008a, in preparation; Kauffmann et al. 2008). Robitaille et al. (2006) discussed the distinction between SED "class" and physical "stage". Physically, the distinction between a Stage 0 and Stage I object is that a Stage 0 object still has more than half of its total mass in the circumstellar envelope (Andr´e et al. 1993). Various evolutionary indicators can be used to separate embedded objects into Class 0 and Class I objects, which should correlate with the true physical stage (see §6.2.2). Froebrich (2005) searched the literature and compiled a database of photometry of all known Class 0 objects. This database is available online3 and is updated as new Class 0 objects are discovered and new photometry is published. If we remove our sixth criterion, which selects only those embedded protostars with Lint . 1 L⊙, our criteria should select all of the objects in this database with available photometry between 3.6 − 70 µm. Since both IRAS and the Infrared Space Observatory (ISO) obtained photometry at 25 and 60 µm, there are many objects in this database with photometry at these two wavelengths 3Available at http://astro.kent.ac.uk/protostars/index.html – 20 – but not at 24 and 70 µm (in other words, observed with IRAS and/or ISO, but not with Spitzer ). Thus, we first selected all sources in the database that had a detection at 60 or 70 µm; 80 of the 135 sources in the database meet this criterion. We then imposed the same criteria as described in §3.1 and 3.3, omitting the sixth criterion as described above, repacing 24 and 70 µm with 25 and 60 µm, respectively, when necessary, and omitting the second, third, and seventh criteria because these require information not available from the database. Of the 80 objects evaluated, 79 pass our criteria and are selected as embedded objects. The one source that does not is IRAS 12500−7658, located in the Chamaeleon II molecular cloud. According to the Class 0 database, this source has F24 = 8.58 × 10−11 ± 0.05 × 10−11 erg s−1 cm−2 and F70 = 6.98 × 10−11 ± 0.43 × 10−11 erg s−1 cm−2, and thus does not have a rising (or flat) flux between 24 or 70 µm, even when the uncertainties are taken into ac- count. This source was observed by c2d, and in fact these values are taken from Young et al. (2005), who presented MIPS data on the Chamaeleon II cloud using a preliminary version of the c2d pipeline. In the final c2d source catalog for Chamaeleon II, which in- cludes, among other things, improved source photometry and more accurate photometric uncertainties (Evans et al. 2007), this source has F24 = 7.64 × 10−11 ± 0.71 × 10−11 erg s−1 cm−2 and F70 = 6.60 × 10−11 ± 0.64 × 10−11 erg s−1 cm−2. Using these new, more accurate values, IRAS 12500−7658 does indeed have a rising (or flat) SED between 24 or 70 µm when the uncertainties are taken into account. Thus, all 80 sources are recovered; we identify all of the Class 0 objects. 4.2. Comparison to Embedded Objects in Perseus, Serpens, and Ophiuchus The above exercise demonstrates that our criteria successfully identify the youngest, most heavily embedded protostars. However, we are interested in the full population of embedded objects, not just the Class 0 objects. To test our method for Class I objects, and to verify the above results, we compare our objects to those identified by Enoch (2007) and Enoch et al. (2008a, in preparation). They searched for Spitzer sources associated with dense cores identified by large-scale, uniform 1.1 mm Bolocam dust continuum emission surveys of Perseus, Serpens, and Ophiuchus. They did not restrict themselves to the low- luminosity population, and since they had good-quality millimeter data available for all three clouds, they started by searching for sources associated with known dense cores. Our work is complementary in that all sources identified here should be identified by Enoch et al., but we can extend our analysis to regions lacking such millimeter data. Comparing the results of both searches shows, in general, good agreement. Enoch et al. – 21 – separated their objects into Class 0 and Class I objects based on the bolometric temperature of each source (see §6.2). After removing sources with LIR ≥ 0.5 L⊙ from their sample, we identify all of their Class 0 sources. This confirms the above result that we identify all Class 0 objects. We also identify most of their Class I sources, but find a population of objects in their sample not identified as candidates in our work. Inspection of this population of objects shows that they are not selected by our criteria because they do not have rising fluxes from 24 to 70 µm. Out of the 37 Class I objects with LIR ≤ 0.5 L⊙ in the Enoch (2007) and Enoch et al. (2008a) sample, 15 (40%) are not selected by our criteria for this reason. These 15 objects range in luminosities from 0.03 ≤ LIR ≤ 0.4 L⊙. Figure 8 shows the average SEDs, weighted by 1/Lbol so that the average is not dominated by the most luminous sources, for all sources identified by Enoch (2007) and Enoch et al. (2008a, in preparation), divided into four bins: Class 0, Class I with rising (or flat) fluxes from 24 to 70 µm, Class I with decreasing fluxes from 24 to 70 µm, and Class II. Following Whitney et al. (2003b), Enoch et al. (2008a, in preparation) use the bolo- metric temperature of each object to go beyond the simple Class 0/I division and divide their sources into Early Class 0, Late Class 0, Early Class I, and Late Class I objects. They present the average SED for each of these four classes of objects; it is the Late Class I ob- jects (300 K ≤ Tbol ≤ 650 K) that feature decreasing fluxes between 24 and 70 µm. Thus, a possible conclusion is that our selection criteria miss the older Class I objects nearing the end of the embedded phase. However, as described in §3.1, we developed our selection criteria based on comparison to models of protostars still embedded within their dense cores. This includes both Class 0 and Class I protostars. Two-dimensional models featuring outflow cavities predict rising or flat fluxes from 24 to 70 µm at even the most extreme pole-on inclinations, (e.g., Whitney et al. 2003a; Whitney et al. 2003b; Robitaille et al. 2006). Indeed, Enoch et al. (2008a, in preparation) compare their average SEDs for Early and Late Class 0 and I objects to the two-dimensional models of each type of object presented by Whitney et al. (2003b) and note that their Late Class I objects differ from the models in that, among other things, they decrease in flux from 24 to 70 µm. The nature of these objects is uncertain. The overall similarities between the average SEDs of the Class I objects rising and falling between 24 and 70 µm (see Figure 8) and the fact that these objects are associated with dense cores suggests they truly are embedded objects. On the other hand, no published models of embedded protostars predict fluxes that decrease between these two wavelengths. Robitaille et al. (2006) noted that the SED class and physical stage of an object do not always agree. For example, a Stage II object located behind a large amount of material (such as a dense core) could feature a Class I SED. It is – 22 – possible that these objects with decreasing fluxes from 24 to 70 µm are not truly embedded protostars but heavily extincted Stage II objects located behind dense cores. A future, detailed study of these objects is needed, but at present, their true physical stage remains unknown. Our criteria will identify the majority of embedded, low-luminosity objects (100% of the Class 0 objects; ≥ 60% of the Class I objects). Even if the objects with decreasing fluxes from 24 to 70 µm are in fact embedded objects, including them in our sample would not significantly change any of our results. 5. Confirming the Candidates Many of the 218 candidates identified in §3 are not truly embedded protostars. The visual inspection described above removed a large number of false candidates, cutting our list down from 726 candidates to 218. However, we only removed sources from consideration if it was obvious that they fell into one of the five types of false candidates discussed in §3.4. Candidates for which there was any ambiguity were kept, and at least some of these are expected to be false candidates. Furthermore, as discussed above, a source is only evaluated by the classification method of Harvey et al. (2007b) for identifying candidate YSOs and candidate galaxies if it is detected in both epochs of observations (or in the single epoch if only one epoch was observed) with SNR ≥ 3 in all 4 IRAC bands and the 24 µm MIPS band. Any source not covered by all 5 of these bands is ineligible for classification as either candidate galaxy or candidate YSO, and instead receives one of several general source types based on the shape of the SED (Evans et al. 2007). Many galaxies thus slip into our candidate list because they pass criterion 7 of §3.1, whereas they would have been classified as candidate galaxies and rejected according to this criterion if they had been observed in all 5 bands. Only those that are large and/or close enough to resolve are removed by the visual inspection described in §3.4. This is particularly relevant for the five large clouds, since the MIPS coverage is substantially larger than the IRAC coverage. To demonstrate the extent of extra-galactic contamination expected, we note that 148 of the 218 candidates have LIR ≤ 0.05 L⊙, which roughly corresponds to Lint . 0.1 L⊙ using the relation between the two derived in §3.1. This suggests a very large number of VeLLOs exist; however, out of these 148 sources, 114 are not considered by the Harvey et al. (2007b) classification method because they do not meet the requirement of SNR ≥ 3 detections in all 5 bands. As Figure 9 demonstrates, the majority of these low-luminosity sources are expected to be galaxies. Out of 851 sources in the 82 regions with dense cores observed by c2d that are classified as either candidate YSOs or candidate galaxies, 604 have LIR ≤ 0.05 L⊙, and 518 of these 604 (∼ 86%) are classified as candidate galaxies. Thus, as many as – 23 – 98 of the 114 candidates not considered by the Harvey et al. (2007b) classification method because they do not meet the requirement of SNR ≥ 3 detections in all 5 bands may be extra-galactic in nature. Clearly, further effort is required in order to determine which of the 218 candidates are in fact embedded, low-luminosity protostars. 5.1. How to Prove a Source is Embedded We examined both the Spitzer data and complementary data for each of the 218 can- didates to search for evidence proving they are truly embedded objects. We divide this evidence into two "levels of certainty": 1. Evidence exists proving that a source is associated, in projection on the sky, with a region of high volume density. 2. Evidence exists proving that this soure is actually embedded within the dense region (to remove the possibility of chance alignment). As discussed below, data on molecular line and dust continuum emission, molecular outflows, extinction, and infrared nebulosity are used to evaluate these conditions. We note here that it is much easier to prove the first level than the second. 5.1.1. Evaluating the First Level The first level is evaluated by observing the region in a tracer of dense material. Such tracers include millimeter dust continuum emission, which typically traces regions with vol- ume densities of n & 104 cm−3 (e.g., Enoch et al. 2007), inversion transition lines of NH3, which also trace regions with volume densities of n & 104 cm−3 (e.g., Ho & Townes 1983; Benson & Myers 1989), and rotational transition lines of N2H+, which trace similar material as the NH3 lines (Caselli et al. 2002). We thus looked for associations with millimeter dust continuum, NH3, or N2H+ sources using a number of different surveys of the regions covered by c2d in these tracers, some of which were specifically designed as complementary surveys to c2d; we list these surveys below. In addition, we searched the SIMBAD database for other such observations besides those listed below. In all cases, a candidate is said to be associated (in projection) with a millimeter dust continuum, NH3, or N2H+ source if it is located within one beam of the peak position of the source. – 24 – The surveys we used for the candidates in the 82 regions with dense cores are as follows: An on-going 350 µm dust continuum survey of cores with the Submillimeter High Angular Resolution Camera II (SHARC-II) at the Caltech Submillimeter Observatory (CSO; Wu et al. 2007; M. Dunham et al. 2008, in preparation), a 450 and 850 µm dust continuum survey of 38 cores with the Submillimeter Common-User Bolometer Array (SCUBA) at the James Clerk Maxwell Telescope (JCMT; Young et al. 2006a), a 1.2 mm dust continuum survey of 37 cores with the Max Planck Millimeter Bolometer (MAMBO) at the IRAM 30 m telescope (Kauffmann et al. 2008), a 1.2 mm dust continuum survey of 151 southern cores with the SEST Imaging Bolometer Array (SIMBA) at the Swedish ESO Submillimeter Telescope (SEST; K. Brede et al. 2008, in preparation), an N2H+ (1-0) survey of 38 cores observed with the Five College Radio Astronomy Observatory (FCRAO; C. De Vries et al. 2008, in preparation), an N2H+ (1-0) survey of 59 cores also observed with the FCRAO (Caselli et al. 2002), and observations of the (J, K) = (1, 1) and (2, 2) lines of NH3 for 264 cores compiled from the literature by Jijina et al. (1999). The surveys used for Perseus are a 1.1 mm dust continuum survey of 7.5 deg2 with Bolocam at the CSO (Enoch et al. 2006), an 850 µm dust continuum survey of 3.5 deg2 with SCUBA at the JCMT (Kirk et al. 2006), a 450 and 850 µm dust continuum survey of 3.0 deg2 with SCUBA at the JCMT (Hatchell et al. 2005; Hatchell et al. 2007a), and the SHARC-II survey described above, which includes several cores in Perseus. For Chamaeleon II, we used a 1.3 mm dust continuum survey of 36 YSOs in Chamaeleon I and II with the 3He- cooled bolometer system at the SEST (Henning et al. 1993), and the SIMBA survey of 151 southern cores described above, which includes several cores in Chamaeleon II. For Lupus, a 1.2 mm dust continuum survey of ∼625 arcmin2 with SIMBA at the SEST (Tachihara et al. 2007) was used. For Ophiuchus, we used a 1.1 mm dust continuum survey of 10.8 deg2 with Bolocam at the CSO (Young et al. 2006b), a 1.2 mm dust continuum survey of 1.0 deg2 with SIMBA at the SEST (Stanke et al. 2006), and an 850 µm dust continuum survey of 700 arcmin2 with SCUBA at the JCMT (Johnstone et al. 2000). Finally, for Serpens we used a 1.1 mm dust continuum survey of 1.5 deg2 with Bolocam at the CSO (Enoch et al. 2007), a 450 and 850 µm dust continuum map of 120 arcmin2 with SCUBA at the JCMT (Davis et al. 1999), and a 1.2 mm dust continuum map of ∼150 arcmin2 with MAMBO at the IRAM 30 m telescope (Djupvik et al. 2006). Observations tracing dense material are not always available at the positions of the candidates. This is especially true for candidates in the southern clouds and cores. As an alternative, observations indicating high column density can be used to evaluate the first level. A positive result does not guarantee this level is satisfied, as high column densities do not necessarily indicate high volume densities, but it does increase the likelihood that the source is an embedded object. – 25 – We utilize two indicators of high column density: extinction maps and absorption against the mid-infrared background. For the extinction maps, we used the maps created by the c2d team. We give a brief description of them here; a more complete description can be found in Evans et al. (2007). A line-of-sight extinction estimate is produced by the c2d pipeline for each source with a 1.25 − 24 µm SED consistent with that of an extincted stellar photosphere, and these line-of-sight estimates were then convolved with uniformly spaced Gaussian beams. Maps of different resolution were produced by using Gaussian beams with different FWHM. We used the highest resolution maps available for each cloud (180′′ for Perseus, 120′′ for Chamaeleon, 120′′ for Lupus I, 90′′ for Lupus III, 90′′ for Lupus IV, 240′′ for Ophiuchus, and 90′′ for Serpens); extinction maps are not yet available for the cores. We did not attempt to derive quantitative estimates of the column density from the extinction maps; instead, we examined the maps at the positions of the candidates and identified any candidate located in a localized region of higher extinction as showing evidence for high column density. For the absorption against the mid-infrared background, we looked for "dark cores" or "shadows" in the 8 and 24 µm images. We examined both the images themselves and radial profiles of the background intensity. Figure 10 shows an example of a dark core at the position of the candidate in L673-7 (source 031). While a quantitative analysis of the extinction profile and column density through the core is possible (e.g., Stutz et al. 2007), it is beyond the scope of this paper. Here we simply identified any candidates that appeared to be located within such a dark core as showing evidence for high column density. An important note is that, in both techniques, negative results do not rule out the possibility of high column density. Very small, localized column density enhancements may not be seen in the extinction maps due to beam dilution, and variations in the strength of both the background and foreground emission will have a significant impact on the presence or absence of absorption against the mid-infrared background. As a result, we consider cases where data are available to evaluate the existence of regions of high column density but return a negative result to be equivalent to cases where no such data are available. 5.1.2. Evaluating the Second Level The second level is primarily evaluated by searching the region around a source for a molecular outflow that is centered both spatially on the source and kinematically at the velocity of the dense core with which the source is associated. Large-scale surveys of 12CO, the primary tracer of molecular outflows, with the necessary sensitivity, spatial resolution, and velocity resolution to detect molecular outflows are not as common as surveys of dust – 26 – continuum. In fact, the only such survey for the regions observed by c2d was a search for outflows in Perseus by Hatchell et al. (2007b), and even this was not an unbiased survey as they targeted known cores from a previous 850 µm dust continuum survey (Hatchell et al. 2005). Thus, we searched the literature for each of the 218 candidates to find those with known molecular outflows. Many candidates had no published 12CO observations, including several candidates in Perseus that were not part of the Hatchell et al. (2007b) survey. Some of these were observed at the CSO as part of a search for molecular outflows, resulting in detections of outflows by us in L673-7 and L1251 (Sources 031, 044, and 045 from this work; M. Dunham et al. 2008, in preparation; J.-E. Lee et al. 2008, in preparation). While a positive result will prove a source is embedded in a dense core, a negative result will not disprove this since the answer can depend on the available data. As an example, L1014 (Source 038 from this work) showed no signs of driving an outflow from single-dish molecular line observations (Crapsi et al. 2005a), but a compact, low-velocity 12CO (2- 1) outflow was discovered by Bourke et al. (2005) with the Submillimeter Array. Thus, in evaluating whether each source shows evidence for being an embedded protostar, we consider cases where observations tracing molecular outflows exist but no outflow is detected to be equivalent to cases where no such observations exist. We also use a 350 µm SHARC-II detection to satisfy the second level. As with all submillimeter and millimeter observations of dust continuum emission, a detection with SHARC-II indicates high volume density. However, unlike other observations of the dust continuum, Wu et al. (2007) concluded that, through a combination of temperature and instrumental effects, a SHARC-II 350 µm detection effectively always indicates that a core has a protostar embedded within it. Many of the 218 candidates have not been observed either in 12CO or with SHARC- II at 350 µm. Both a search for molecular outflows and an extension of the SHARC-II survey of dense cores presented by Wu et al. (2007) are on-going at the CSO (M. Dunham et al. 2008, in preparation), but in the meantime, this leaves us unable to evaluate the second level for many candidates. Thus, we also examine the IRAC 2 (4.5 µm) image for each candidate to look for extended nebulosity or jets, suggesting the presence of molecular outflows and outflow cavities. We specifically focus on 4.5 µm for two reasons: the wavelength is short enough for scattered light off the edges of outflow cavities to often be visible, and the photometric band overlaps with emission lines of molecular hydrogen shocked by outflows (e.g., Noriega-Crespo et al. 2004). Examples of this are shown for Sources 001 and 004 from this work in Dunham et al. (2006) and Bourke et al. (2006), respectively. We do not separately identify "extended nebulosity" and "jets"; instead we consider evidence for either to be the same thing. While this is not strictly correct, it avoids the ambiguity sometimes inherent in assigning extended structure one label or the other. Furthermore, – 27 – both are likely indicators of outflow activity, the actual source property of relevance for this study. A positive result does not guarantee that the second level is satisfied, but it does add to the likelihood that the source is an embedded object. Once again, however, a negative result does not prove that the source is not embedded. To again use L1014 (Source 038) as an example, it does not show any extended nebulosity or jet features in the 4.5 µm image (or any other Spitzer image); only in deep near-infrared images is scattered light from an outflow cavity seen (Huard et al. 2006). Thus, as above, we consider cases where 4.5 µm Spitzer images are available but show no extended nebulosity and/or jets to be equivalent to cases where no such images are available. 5.2. Assessing the Likelihood of Being Embedded Based on the results of evaluating the two levels of certainty described above, we divide the candidates into 6 groups of descending likelihood of being embedded protostars. We list the groups in Table 9 and describe them below: 1. Group 1 consists of candidates that are confirmed as embedded, low-luminosity pro- tostars. Existing observations confirm that sources in this group are embedded within regions of high volume density. 2. Group 2 consists of candidates that have a very high likelihood of being embedded, low-luminosity protostars. Sources in this group are associated with regions of high volume density. They show extended nebulosity and/or jets in 4.5 µm Spitzer images, but they are not confirmed to be embedded within the high-density regions, either because no observations are available to evaluate this condition or such observations are available but give a negative result. 3. Group 3 consists of candidates that have a high likelihood of being embedded, low- luminosity protostars, although not as high as group 2 candidates. Similar to group 2, group 3 candidates are associated with regions of high volume density but are not confirmed to be embedded within these regions. Unlike group 2, however, group 3 candidates are not associated with extended nebulosity and/or jets at 4.5 µm, either because no such extended structure is detected or because 4.5 µm images are not available. 4. Group 4 consists of candidates that show evidence for being embedded, low-luminosity protostars but lack confirmation that either level of certainty is fulfilled. Sources in this group have no available observations to evaluate whether or not they are associated – 28 – with regions of high volume density and either no available observations or observations giving a negative result on whether or not they are embedded within regions of high volume density. These sources either are associated with regions of high column density and show extended nebulosity and/or jets in 4.5 µm Spitzer images (group 4a), are only associated with regions of high column density (group 4b), or only show extended nebulosity and/or jets in 4.5 µm Spitzer images (group 4c). 5. Group 5 consists of candidates that might be embedded, low-luminosity protostars, but show no sign of being such objects in their limited available data. Sources in this group have no available observations to evaluate whether or not they are associated with regions of high volume density and either no available observations or observations returning a negative result for proving they are embedded within dense regions, evalu- ating their association with regions of high column density, and identifying nebulosity or jets at 4.5 µm. 6. Group 6 consists of candidates that are most likely not embedded, low-luminosity protostars. Observations show that sources in this group are not associated with regions of high volume density, with the important caveat that most dust continuum emission surveys used to search for regions of high volume density are only complete to cores with masses ≥ 0.1 − 1.0 M⊙ (e.g., Enoch et al. 2007). The last column of Table 9 lists the total number of the 218 candidates in each of the groups. Table 10 lists, for all 218 candidates, sorted by group, the source number, cloud or core region in which the source is located, LIR, the status of 160 µm observations for this source, and the SIMBAD name of the source, where applicable. If a candidate is not associated with an infrared or radio point source in SIMBAD, we give the name of the nearest extended dense core within a search radius of 1′. Note that this does not guarantee the candidate is embedded within this core. We find that 149 of the 218 candidates are in either group 5 or group 6 and thus show no evidence for being embedded, low-luminosity protostars. A fraction of these 149 candidates are likely real embedded sources that simply lack complementary data of sufficient quality to prove this; these sources should be re-evaluated as additional surveys of these regions become available. Most, however, are likely either galaxies or more evolved sources no longer embedded within dense cores, consistent with our earlier claim that as many as 98 of the candidates may be extra-galactic in nature (§5). This leaves 69/218 candidates that, at minimum, show some evidence for being embedded objects. We are able to obtain complete SEDs for 50 of these 69 candidates; these SEDs are listed in Table 11. We consider only these 50 sources in the rest of the paper. Of the 19 remaining candidates, 16 are located – 29 – close enough to other sources that they cannot be separated into individual sources at 70 µm; these are discussed below. Two sources, 027 and 202, are not covered by the 70 µm c2d observations and have no obvious association with IRAS 60 µm sources. As 70 µm is an essential wavelength for this study (see below), we do not list these objects in Table 11. The last remaining source is Source 132, which is identical to Source 017 as it was observed by both IRAC and MIPS in the map of the dense core DC303.8−14.2 and by MIPS in the map of Chamaeleon II. We thus remove Source 132 from consideration to avoid duplication. 5.3. Effects of Confused 70 µm Data Good quality 70 µm data are essential to this study. Photometry at this wavelength is important for distinguishing between SEDs of embedded protostars and more evolved YSOs no longer embedded in dense cores (§3.1; §4.2), and for deriving accurate protostellar internal luminosities (§3.2). However, the MIPS 70 µm PSF is ∼ 18′′, compared to 6′′ at 24 µm and 2′′ at 3.6 − 8.0 µm. Embedded protostars spaced by more than 6′′ but less than 18′′ will be resolved into individual sources at 24 µm and thus identified by our selection criteria, but confused at 70 µm. Various methods exist for dividing the total 70 µm flux among the confused sources, such as by the ratio of 24 µm fluxes of each source (e.g., Lee et al. 2006), but all such methods introduce large uncertainties. Of the 69 candidates that show at least some evidence for being embedded objects, 16 are located too close to other objects to resolve into individual sources at 70 µm4. Many likely have Lint ≥ 1.0 L⊙ and are thus not relevant to this study, since most have values of LIR, integrated only to 24 µm, greater than 0.1 L⊙. However, a few might be embedded, low-luminosity protostars. We do not list any of these 16 candidates in Table 11 and do not consider them in the analysis discussed below. Higher-resolution far-infrared data (e.g., Herschel) will be necessary to evaluate the nature of these objects. 6. Discussion 6.1. Estimating Internal Luminosities and the Luminosity Distribution The internal luminosity is a key parameter in understanding the evolution of an object from dense core to star, but for low-luminosity, embedded protostars, it is not a directly 4The source numbers are 002, 006, 027, 048, 049, 055, 056, 073, 208, 210, 213, 214, 215, 108, 211, 212. – 30 – observable quantity since a significant component of the bolometric luminosity will arise from external heating by the ISRF. In §3, we presented two methods of estimating Lint based on observable quantities. The first estimate is based on the result that an approximately constant ratio exists between Lint and LIR, where LIR is an observable quantity. We call this estimate of the internal luminosity LIR int. The second estimate is based on the tight correlation observed between the 70 µm flux, scaled to the value it would have if the source were located at 140 pc, and the internal luminosity. This estimate, which we call L70 int, is given by Equation 2. int vs. L70 Figure 11, which plots LIR int, demonstrates that the two estimates agree to within a factor of two of each other for nearly all of the sources. We consider L70 int to be a more accurate estimate of Lint, both because nearly identical relationships between Lint and 70 µm flux normalized to 140 pc were derived separately from observations of protostars and a large grid of models and because LIR int is based on a relationship between Lint and LIR that shows some variation from source to source. Throughout the rest of this paper we use L70 int when we refer to internal luminosities. Detailed radiative transfer modeling of each source is necessary to obtain more accurate internal luminosities. Figure 12 shows the distribution of Log(Lint) for the 50 objects listed in Table 11. The decrease in the number of protostars near Lint ∼ 1 L⊙ is an artifact introduced by requiring that all sources have LIR ≤ 0.5 L⊙, which is roughly equivalent, but not identical, to requiring that all sources have Lint ≤ 1.0 L⊙ given that the ratio of Lint/LIR has a weighted average of 1.7 but varies from source to source. The full distribution of protostellar luminosities will be examined in a future paper (Evans et al. 2008, in preparation) that combines this work with other, complementary searches for embedded protostars in the c2d sample (e.g., Enoch 2007; Enoch et al. 2008a, in preparation; Jørgensen et al. 2007; Jørgensen et al. 2008, in preparation). Here we only focus on the low end of the distribution. We find that 15/50 (30%) of the objects identified here have Lint ≤ 0.1 L⊙ and are thus classified as VeLLOs. Qualitatively speaking, this indicates that there are more embedded protostars at lower luminosities than at higher luminosities (if they were distributed evenly in luminosity, only 10% would have Lint ≤ 0.1 L⊙). To quantify this, Figure 13 shows the distribution of source internal luminosities, cut off at Lint = 0.5 L⊙ since our sample is likely incomplete near 1 L⊙. A visual inspection of this figure suggests that the distribution is not uniform but increases with decreasing luminosity. Applying a K-S test, we find that there is a 94.2% probability that these sources are not drawn from a uniform luminosity distribution. The implications of this increase in number of protostars with decreasing luminosity will be explored in §6.4. Six VeLLOs have internal luminosities lower than our sensitivity limit for the most – 31 – distant regions (∼ 5 × 10−2 L⊙), raising the question of how many such sources we miss in the more distant regions of our sample. A discussion of correction for completeness due to the non-uniform distances to observed regions is presented in Appendix A. Based on this discussion, we conclude that only ∼ 2 sources with luminosities between our sensitivity limit for the closest regions (Lint = 2.5×10−3 L⊙) and the most distant regions (Lint = 5×10−2 L⊙) are missed. Also based on Appendix A, we conclude that the intrinsic luminosity distribution does not continue to increase below Lint = 0.1 L⊙, although there are sources present all the way down to the sensitivity limit. Finally, we note that we can not draw any conclusions about the possible presence of objects with extremely low luminosities below ∼ 10−3 L⊙. Future infrared facilities with sensitivities much greater than that of Spitzer will be required to search for such objects and determine the true lower limit to the intrinsic luminosity distribution. The above conclusions are based on the analysis presented in Appendix A. This analysis assumes that all sources with luminosities above our sensitivity limit of Lint = 4 × 10−3 (d/140 pc)2 L⊙ are in fact identified. To verify that this is indeed the case, we re-examine the criteria described in §3. The only criterion that could potentially filter out sources with very low luminosities is the requirement that a source not be classified as a candidate galaxy in the classification method of Harvey et al. (2007b) unless Log(Pgal) ≤ −1.25, since several of the steps in this classification method are based on the general fact that galaxies are faint (see Harvey et al. for details). Using Perseus as an example, we re-applied our selection criteria, leaving out the step that filters out candidate galaxies with Log(Pgal) ≥ −1.25. This results in only an additional 36 sources, since most of the additional sources that would otherwise be selected are rejected because they are not detected at 70 µm. We visually inspected these sources and rejected 14 based on either showing obvious galaxy morpologies or not being real sources (see §3.4 for details). This left us with 22 sources. The positions of all 22 sources were covered by at least one of the three submillimeter/millimeter dust continuum emission surveys of Perseus listed in §5.1.1, but none were associated with dense cores. All 22 sources would thus be placed into group 6 in the terminology of §5.2. We conclude that we do indeed identify all sources above our sensitivity limit. 6.2. Evolutionary Indicators For the 50 objects with complete SEDs listed in Table 11, we calculate the following quantities: the bolometric luminosity (Lbol), the ratio of bolometric to submillimeter lumi- nosity (Lbol/Lsmm), and the bolometric temperature (Tbol). Lbol is calculated by intergrating – 32 – over the full observed SED, Lbol = 4πd2Z ∞ 0 Sνdν , (4) while the submillimeter luminosity is calculated by integrating over the observed SED for λ ≥ 350 µm, Lsmm = 4πd2Z ν=c/350 µm 0 Sνdν . (5) The bolometric temperature is defined to be the temperature of a blackbody with the same flux-weighted mean frequency as the source (Myers & Ladd 1993). Following Myers & Ladd, Tbol is calculated as Tbol = 1.25 × 10−11 R ∞ R ∞ 0 νSνdν 0 Sνdν K . (6) We list Lbol, Lbol/Lsmm, and Tbol for these 50 objects in the last column of Table 11. The stated uncertainties reflect only those arising from measurement error. Since we are inte- grating over finitely sampled SEDs, the values calculated for all three quantities will depend on how well each SED is sampled. We discuss the errors introduced by incomplete sampling in Appendix B. 6.2.1. Bolometric Luminosity-Temperature (BLT) Diagram Figure 14 places the 50 objects listed in Table 11 on Figure 19 from Young & Evans (2005), which plots Lbol vs. Tbol. As discussed by Myers et al. (1998), who refer to this as a Bolometric Luminosity-Temperature (BLT) diagram, this is effectively a Hertzsprung- Russell diagram for protostars. The lines show the evolutionary tracks followed by various models from Young & Evans (2005) and Myers et al. (1998). In general, models predict that protostars form at low values of Lbol and Tbol and move up and to the left (increasing Lbol, increasing Tbol) as the protostar grows. Whether or not Lbol steadily increases or begins to decrease depends on the details of the mass accretion process. Young & Evans considered the collapse of a singular isothermal sphere in the "standard model" of star formation (Shu, Adams, & Lizano 1987), which features a constant mass accretion rate over the entire protostellar phase. Myers et al. (1998), on the other hand, assume that the accretion rate decreases exponentially with time and mass is lost from the system (from outflows, winds, etc.), explaining the different trends shown by their evolutionary tracks. The colors/symbols of the points in Figure 14 for the 50 objects from this work reflect how well the source SED is sampled. All four colors/symbols have SEDs that are sampled at 1.25 − 70 µm by 2MASS and Spitzer and at millimeter wavelengths. Black circles are also – 33 – sampled at, at minimum, one wavelength in both the 100−200 and 350−450 µm wavelength ranges (category 1 SEDs in the notation of Appendix B). Blue triangles are sampled between 350−450 µm but not 100−200 µm (category 2 SEDs). Green diamonds are sampled between 100−200 µm but not 350−450 µm (category 3 SEDs). Red squares are not sampled in either wavelength range (category 4 SEDs). A detailed analysis of the errors introduced by the incomplete sampling in these various categories is presented in Appendix B; here we simply note that, depending on the category, the error in Lbol and Tbol is between approximately 20 − 60%. There are many Class I sources (Tbol ≥ 70 K; Chen et al. [1995]) with Lbol below any evolutionary track by up to an order of magnitude. The most extreme cases are all sources with category 4 SEDs, where Lbol is underestimated by ∼ 50%, on average. However, even accounting for this, many are still substantially below any of the evolutionary tracks. A similar result was found by Enoch (2007) and Enoch et al. (2008a, in preparation), who suggest episodic accretion as a potential solution. Another potential explanation for this discrepancy is that the sources with the lowest Lbol are simply those with envelope masses below those considered in the models. However, this is ruled out by Enoch et al., who find good agreement between observations and models when comparing envelope mass vs. bolometric temperature despite the scatter seen in BLT diagrams, as well as several individual studies of VeLLOs that conclude at least some of the lowest-luminosity protostars are embedded in envelopes with masses similar to those of more luminous protostars (a few M⊙; Young et al. 2004; Bourke et al. 2006; Dunham et al. 2006). A similar population of Class 0 objects with luminosities below model predictions is not as apparent from Figure 14, and is also not seen by Enoch et al. However, on average, external heating from the ISRF will add a larger contribution to Lbol for objects with more massive envelopes, thus such heating is likely to be more significant for Class 0 objects than Class I objects and can add up to several tenths of a solar luminosity to Lbol (e.g., Evans et al. 2001). Since, as described in §1, this distinction between Lbol and Lint is most relevant for very low-luminosity sources where external heating from the ISRF can dominate the total observed Lbol, a similar population of Class 0 objects with very low luminosities may be present but not readily apparent when examining a plot of Lbol vs. Tbol. We return to this point in §6.4. 6.2.2. Comparison of two Indicators of Evolutionary Status Both Tbol and Lbol/Lsmm are indicators of the evolutionary status of a protostar. As the envelope mass decreases through a combination of accretion and dissipation processes, – 34 – both quantities should increase. Eventually, once the star has fully formed, Tbol will become equal to Tef f and Lbol/Lsmm will become very large. We show a plot of Lbol/Lsmm vs. Tbol for the objects considered here in Figure 15. The division between Class 0/I objects in Tbol is from Chen et al. (1995), while the division between Class 0/I and Class I/II objects in Lbol/Lsmm is from Young & Evans (2005). The colors/symbols of the data points hold the same meaning as in Figure 14. As stated in §4.1, the distinction between a Stage 0 and Stage I object is that a Stage 0 object still has more than half of the total mass of the system in the envelope, whereas less than half the total mass remains in the envelope of a Stage I object. Both Tbol and Lbol/Lsmm are used to separate embedded objects into Class 0 and Class I objects; in principle, these observational classes should correspond to Stage 0 and Stage I objects. Young & Evans (2005) concluded that Lbol/Lsmm is a better measure of physical stage than Tbol, since their evolutionary models of different initial mass cores featured similar values of Lbol/Lsmm but different values of Tbol when they reached the point at which half of the total mass had accreted onto the protostar. They also found that embedded objects cross the Class 0/I boundary in Tbol (Tbol = 70K; Chen et al. 1995) when they are still Stage 0 objects. We see a similar trend in Figure 15 in that several objects are Class I objects according to Tbol but Class 0 objects according to Lbol/Lsmm. However, we caution that the details of where the observed SED is sampled affect the calculated value of Lbol/Lsmm significantly more than the calculated value of Tbol; this explains the objects seen in Figure 15 with extremely high values of Lbol/Lsmm (see Appendix B). Great care must be taken to consider the errors introduced by incomplete sampling before using Lbol/Lsmm as an evolutionary indicator for any particular object. 6.3. Fraction of Starless Cores that Remain Starless Out of the 95 individual dense cores observed by c2d5, 67 (∼70%) were classified as starless prior to being observed by Spitzer. The very first starless core observed, L1014, was found to actually harbor a very low-luminosity, embedded protostar (Lint ∼ 0.09 L⊙; Young et al. 2004; Bourke et al. 2005; Huard et al. 2005). This naturally gives rise to the question of how many cores classified as starless prior to the launch of Spitzer remain so after being 5Evans et al. (2007) list the 82 core regions surveyed by c2d. Those regions that covered multiple cores are as follows (see also Evans et al. 2003 and references therein): CB130-3 (CB130-1, CB130-2, and CB130-3), CG30-31 (CG30, CG31A, CG31B, CG31C), IRAM04191+1522 (IRAM04191+1522 and IRAS04191+1523), L1251 (L1251A, L1251C, and L1251E), L43 (L43-East and L43-RNO91), L673 (L673-SMM1 and L673- SMM2), and TMC1 (TMC1-1, TMC1-2, TMC1-1C-1, and TMC1-1C-2). – 35 – observed with Spitzer ? A substantial number of protostellar cores misclassified as starless cores would result in an incorrect estimate of the lifetime of the starless core phase6, as such estimates are derived by comparing the numbers of starless and protostellar cores (e.g., Kirk et al. 2005; Enoch et al. 2008b). Kirk et al. (2007) present the results of a search for embedded, low-luminosity protostars in Spitzer observations of 22 starless cores previously studied using submillimeter data (Kirk et al. 2005); they find only one such source, L1521F- IRS (Bourke et al. 2006). This suggests that ∼95% of starless cores remain starless, leading them to conclude that the errors in starless core lifetime estimates introduced by previously undetected, low-luminosity protostars are smaller than the uncertainties in such estimates. Our sample of 67 starless cores allows us to test these conclusions with a dataset approx- imately three times larger than that of Kirk et al. (2007). Searching all 95 cores considered here, we have identified a total of 49 candidate embedded, low-luminosity protostars in 33 different cores (Table 3). Seventeen of these cores were classified as starless prior to being observed by Spitzer. If we restrict ourselves only to those sources in Groups 1 − 4, those that, at minimum, show some additional evidence of being embedded protostars besides the shapes of their SEDs, we identify 29 candidates in 21 different cores, 9 of which were classified as starless prior to being observed by Spitzer. Thus, between 9 and 17 out of the 67 starless cores actually harbor embedded, low-luminosity protostars. Approximately 75−85% of star- less cores remain starless down to our luminosity sensitivity of Lint ≥ 4 × 10−3 (d/140 pc)2. Most starless cores do indeed remain starless after being observed by Spitzer, in agreement with Kirk et al. (2007), although the fraction is smaller than they found with their smaller sample. 6.4. Non-Steady Mass Accretion In this work, we have identified a sample of 50 embedded protostars with Lint ≤ 1.0 L⊙; 15 of these objects have Lint ≤ 0.1 L⊙ and are thus classified as VeLLOs. Assuming spherical mass accretion at the rate predicted by the standard model ( Macc ∼ 2 × 10−6 M⊙ yr−1; Shu, Adams, & Lizano 1987) onto an object with a typical protostellar radius of R ∼ 3 R⊙, a protostar located on the stellar/substellar boundary (M = 0.08 M⊙) would have an accretion luminosity, Lacc = GM Macc , of L ∼ 1.6 L⊙. The objects identified here are difficult to understand in the context of the standard model of star formation. VeLLOs, with luminosities more than an order of magnitude lower than the above calculation, are R 6In actuality, this is an estimate of the lifetime of the portion of the starless core phase detectable by (sub)millimeter observations. This is often referred to as the pre-stellar core phase (Kirk et al. 2005). – 36 – an extreme example of this problem. If the objects identified here were observed edge-on through circumstellar disks their luminosities could possibly be underestimated, but, for at least some of them, this possibility can be eliminated (e.g., IRAM04191, Dunham et al. 2006). Thus, they must either feature mass accretion rates far below that predicted by the standard model, masses far below the stellar/substellar boundary, or some combination of the two. Several authors have invoked episodic accretion to explain the existence of these low-luminosity objects (e.g., Kenyon & Hartmann 1995; Young & Evans 2005; Enoch 2007; Enoch et al. 2008a, in preparation). Strong evidence for non-steady mass accretion exists for the VeLLO IRAM 04191-IRS. This object drives a well-defined, bipolar molecular outflow (Andr´e et al. 1999). Taking the average mass accretion rate implied by the outflow and the dynamical time of this outflow (h Macci ∼ 5 × 10−6 M⊙ yr−1 and td ∼ 104 years), Andr´e et al. (1999) calculate that a protostellar mass of 0.05 M⊙ will have accreted over the lifetime of the outflow. Accretion at this rate onto an object with a mass of 0.05 M⊙ and a radius of 3 R⊙ would give rise to an accretion luminosity of Lacc ∼ 2 L⊙. This is clearly inconsistent with the internal luminosity of Lint ∼ 0.08 L⊙ derived by Dunham et al. (2006). Dunham et al. suggested non-steady mass accretion to resolve this inconsistency: the current mass accretion rate must be much lower than the average rate over the lifetime of the outflow. We show the location of the 15 VeLLOs in our sample on the BLT diagram in Figure 16. Most are located at Tbol < 70 K and Lbol > 0.1 L⊙ and are thus Class 0 objects with very low internal luminosities but bolometric luminosities that generally appear consistent with evolutionary tracks. If there is a population of Class 0 objects with luminosities lower than predicted by evolutionary tracks, as is the case for Class I objects (see §6.2.1), the inconsis- tency between the luminosities of these Class 0 objects and the predictions of evolutionary tracks will likely be masked on a BLT diagram since, as described in §6.2.1, the contribution to Lbol from external heating will be larger for younger objects with more massive envelopes. In addition to Lbol and Tbol, considering time as a third dimension relevant to protostellar evolution can help illuminate whether or not these Class 0 objects with Lint ≤ 0.1 L⊙ but Lbol > 0.1 L⊙ are truly as consistent with constant mass-accretion evolutionary tracks as they appear to be on a BLT diagram. Assuming a constant rate of mass accretion, the protostellar mass M will grow linearly with time. If we assume the protostellar radius remains fixed, Lacc ∝ M and thus also grows linearly with time. Adding in the assumption that star formation has been occuring continuously over a time much longer than the lifetime of the embedded phase, the number of protostars as a function of luminosity should be constant. However, as discussed in §6.1, the intrinsic luminosity distribution is not uniform and instead increases with decreasing – 37 – luminosity. Even though many of the lowest luminosity objects appear consistent with the location of evolutionary tracks calculated assuming constant mass accretion, there are too many relative to higher luminosity sources to explain by constant mass accretion. We have presented evidence for non-steady mass accretion in low-mass protostars, both from the above statistical argument and from a detailed study of one individual source (Dun- ham et al. 2006). While we note here that non-steady mass accretion does not necessarily imply episodic mass accretion, these results, combined with other studies that find clear evi- dence for episodicity in outflow activity (e.g., HH211; Lee et al. 2007), strongly suggest that protostellar mass accretion is episodic in nature. Future work is needed on both observational and theoretical fronts to better understand the implications of the objects identified in this study on the mass accretion process in low-mass star formation. On the observational front, the detailed study of the outflow driven by IRAM04191 must be repeated for other sources known to drive outflows, and dedicated outflow searches are needed towards objects lacking such data. Additionally, detailed chemical and physical studies of the objects identified in this work are needed to explore the nature of these sources beyond the simple evolutionary indicators calculated here. Such studies are of particular relevance to understanding the evolution of starless cores and identifying those cores on the verge of collapse since VeLLOs have been discovered in cores that were classified as starless prior to Spitzer observations but were not always those cores believed to be the most evolved and nearest to the onset of collapse (Crapsi et al. 2005b; Bourke et al. 2006). Chemical studies can also be used to distinguish between low-luminosity protostars that have featured higher past luminosities (presumably through higher past mass accretion rates) and those that have always featured such low luminosities (e.g., Lee 2008). Candidates identified in this study that show no evi- dence for being embedded in dense cores should be re-examined as additional data on each source becomes available (e.g., through large-scale surveys with Herschel, SCUBA-2, etc.) to ensure as complete a sample as possible. Finally, the full sample of embedded, low-mass protostars must be assembled from this work and the various other searches for such objects outlined in §1. On the theoretical front, evolutionary models predicting the observational signatures of dense cores forming stars must be revisited. These models must be able to explain both the existence of embedded protostars with luminosities much lower than predicted by current models and the very large scatter in Lbol observed at each value of Tbol in Figure 14. Models that incorporate episodic mass accretion and the effects of source geometry may provide such explanations, but will require higher dimensions than the 1-D models considered by us to date. – 38 – 7. Conclusions We have conducted a search for all embedded protostars with Lint ≤ 1.0 L⊙ in the c2d dataset of nearby, low-mass star-forming regions. We identify 218 candidates from the Spitzer data alone; examining all available complementary data for each candidate results in a sample of 50 objects that show at least some evidence that they are indeed embedded within dense cores. A summary of our major results is as follows: • On average, the Spitzer c2d data are sensitive to embedded protostars with Lint ≥ 4 × 10−3 (d/140 pc)2 L⊙, a factor of 25 better than the sensitivity of the Infrared Astronomical Satellite (IRAS) to such objects. • The 70 µm flux and internal luminosity of a protostar are tightly correlated. As the former is a directly observable quantity but the latter is not, this correlation gives a powerful method for estimating protostellar internal luminosities when detailed radia- tive transfer models for each source are lacking. • Of the 50 objects in our sample, 15 (30%) have Lint ≤ 0.1 L⊙ and are thus classified as VeLLOs. The distribution of source luminosities is not uniform and instead increases with decreasing luminosity. Accounting for incompleteness arising from non-uniform distances to the observed regions, we find sources down to the above sensitivity limit, indicating that the intrinsic luminosity distribution may extend to lower luminosities than probed by these observations. Despite this, we are able to rule out a continued rise in the distribution below Lint = 0.1 L⊙. • Between 75 − 85% of cores classified as starless prior to being observed by Spitzer remain starless down to the above luminosity sensitivity; the remaining 15 − 25% harbor low-luminosity, embedded protostars. This is in general agreement with Kirk et al. (2007), who examined archival Spitzer data of 22 starless cores and found only one to be harboring a low-luminosity protostar. However, with our larger sample size, we are able to better constrain the fraction of cores previously classified as starless that in fact harbor such objects. We confirm that recent estimates of starless core lifetimes (e.g., Kirk et al. 2005; Enoch et al. 2008b) do not feature large errors introduced by previously undetected, low-luminosity protostars. • The observed luminosity distribution for embedded objects with Lint ≤ 1.0 L⊙ is incon- sistent with the simplest picture of star formation wherein mass accretes from the core onto the protostar at a constant rate. Combining this result with other studies that find clear indications of episodic outflow activity strongly suggests that protostellar mass accretion is episodic in nature. – 39 – We have outlined several avenues of future work that must be pursued now that rela- tively complete and unbiased samples of embedded, low-mass protostars are being compiled. Only with such future studies can we begin to build a coherent picture of low-mass star formation consistent with the growing observational database provided by systematic, large- scale surveys of low-mass star forming regions. The authors thank M. Enoch for providing the data used in Figure 8, as well as for several helpful discussions that have improved the quality of this paper. We express our gratitude to the anonymous referee for several comments that have improved the quality of this publication. We thank Jes Jørgensen and Paul Harvey for providing helpfuul comments, and we also thank Jes Jørgensen for his IDL scripts to display three-color images. This work is based primarily on observations obtained with the Spitzer Space Telescope, operated by the Jet Propulsion Laboratory, California Institute of Technology. We thank the Lorentz Center in Leiden for hosting several meetings that contributed to this paper. This publication makes use of the Protostars Webpage hosted by the University of Kent, as well as 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. These data were provided by the NASA/IPAC Infrared Science Archive, which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with NASA. This research has made use of NASA's Astrophysics Data System (ADS) Abstract Service and of the SIMBAD database, operated at CDS, Strasbourg, France. Support for this work, part of the Spitzer Legacy Science Program, was provided by NASA through contracts 1224608 and 1288658 issued by the Jet Propulsion Laboratory, California Institute of Technology, under NASA contract 1407. Support was also provided by NASA Origins grants NNG04GG24G, NNX07AJ72G, and NAG5-13050. A. Internal Luminosity Completeness and the Intrinsic Luminosity Distribution The regions observed by c2d are located at distances ranging from 125 − 500 pc. Using the result from §3.1 that our observations are sensitive to embedded protostars with Lint ≥ 4 × 10−3 (d/140 pc)2 L⊙, the effect of these non-uniform distances is that our sensitivity to embedded protostars varies between 2.5 × 10−3 and 5 × 10−2 L⊙. Two important questions thus arise: (1) How many objects are lost because of these non-uniform sensitivities (stated another way, how many additional objects would we expect to identify if all observed regions were located at the distance of the closest regions)?, and (2) What conclusions can be drawn – 40 – about the intrinsic luminosity distribution? The latter is of particular significance to future work aimed at explaining the observed distribution of sources in Lbol-Tbol space through evolutionary modeling. A.1. Internal Luminosity Completeness The regions observed consist of both targeted observations of 82 regions with 95 dense cores and unbiased surveys of 5 large, molecular clouds (§2). Thus, an analysis of complete- ness that uses the area observed at different distances is incorrect since the sample consists both of observations biased towards dense cores and observations with no such bias. Instead, we start from the total number of dense cores observed at different distances since all em- bedded protostars are, by definition, within these cores. The sample consists of 122 cores in Perseus (Enoch et al. 2006), 35 cores in Serpens (Enoch et al. 2007), and 48 cores in Ophi- uchus (Young et al. 2006b), along with the targeted observations of 95 cores. This yields a total of 300 cores; the distribution of the sensitivity to protostars embedded within these cores is shown in Figure 17. We have no information on the number of dense cores in the Lupus and Chamaeleon II clouds; any such cores have comparable luminosity sensitivities to the rest of the sample (Figure 17). To start with a simple analysis, we ask what percentage of the cores are located at distances close enough such that sources in each of the lowest three bins of Figure 12 could be detected. Statistically, this should be equal to the percentage of sources in that bin we are able to detect. All of the cores are located at distances such that sources in the bin -1.2 ≤ Log(Lint/L⊙) ≤ -0.9 could be detected, thus we are complete to sources in this bin. 96% of the cores are located at distances such that sources in the bin -1.5 ≤ Log(Lint/L⊙) ≤ -1.2 could be detected; correcting for completeness raises the total number of sources in this bin from 5 to 5.2. 90% of the cores are located at distances such that sources in the bin -1.8 ≤ Log(Lint/L⊙) ≤ -1.5 could be detected; correcting for completeness raises the total number of sources in this bin from 4 to 4.4. We conclude that we miss, at most, 1 source in the range -1.8 ≤ Log(Lint/L⊙) ≤ -0.9; this is actually a strong upper limit since the percentage of cores with distances close enough to detect sources in these bins would increase if the dense cores in the Lupus and Chamaeleon II clouds were included. However, this says nothing about the bins that are above our sensitivity limit for the nearest regions (2.5 × 10−3 L⊙) in which we detect no sources. To examine the completeness in these bins, and confirm the above results, we turn to a Monte Carlo simulation. For the Monte Carlo simulation, we create a sample of 10,000 embedded protostars. We assign an internal luminosity to each protostar from a distribution specified at the start; in – 41 – all cases, this distribution has a maximum of Lint = 0.1 L⊙. We then randomly place each of the 10,000 sources in one of the 300 dense cores. If a source has Lint less than the sensitivity limit for the core in which it is placed, it is considered "missed" by our observations and rejected; otherwise, the source is kept. We run the simulation 1000 times, average the results, and compare the observed distribution of Lint for the 15 VeLLOs with the average simulated observed distribution. Figure 18 shows the results of a simulation where the 10,000 sources have internal luminosities evenly distributed (in linear space) between 10−5 and 10−1 L⊙. The normalized observed distributions of Lint for the 15 VeLLOs identified by this work and for the simulated sources agree given the uncertainties due to small number statistics. Furthermore, since we detect no sources with Log(Lint) ≤ −1.8, we can set an upper limit of ≤ 1/15 sources (or ≤ 7% of all sources detected) have such luminosities. Out of all the sources detected in the Monte Carlo simulation, 6.1% have such luminosities, consistent with the observed upper limit of 7%. Out of the 10,000 sources considered by this simulation, 1158 (12%) have luminosities below the sensitivity limit of the core in which they are placed and are thus "missed" by our observations. This means that the 15 VeLLOs identified are only 88% of the total number of objects, indicating that approximately 2 sources are missed. The exact details of the input luminosity distribution do not affect these results, as long as the simulation is consistent with the upper limit on the number of detected sources with Log(Lint) ≤ −1.8 described above. Input luminosity distributions that have a much larger fraction of sources at lower luminosities, such as a distribution flat in log space, would lead to an increase in the number of sources missed. However, such distributions are inconsistent with the above upper limit (for example, a simulation with an input luminosity distribution flat in log space between 10−3 and 10−1 L⊙ results in 24% of the total sources detected having Log(Lint) ≤ −1.8, clearly inconsistent with the observed upper limit of 7%). Thus, we conclude that ∼ 2 sources are missed due to the non-uniform distances of the observed regions. A.2. Intrinsic Luminosity Distribution We now consider what constraints we can place on the low end of the intrinsic luminosity distribution. Additional constraints will be provided by a future paper aimed at analyzing the shape of the full luminosity distribution, including higher luminosity sources (Evans et al. 2008, in preparation). These constraints will serve as guides to future work aimed at explaining the full distribution of sources in Lbol-Tbol space. – 42 – As stated in §6.1, the intrinsic luminosity distribution for the objects identified here increases to lower luminosities. To examine whether or not this increase continues to our sensitivity limit, Figure 19 shows the distribution of internal luminosities, in linear bins of 0.01 L⊙, for the 15 objects with Lint ≤ 0.1 L⊙. Considering the small sample size, this distribution appears uniform down to Lint = 0.02 L⊙. From the discussion above, approximately two sources are missing from this figure because of incompleteness arising from the non-uniform distances to the observed regions. Additionally, a strong upper limit of 1 source is lost due to incompleteness for Lint ≥ 10−1.8 L⊙ (denoted by the dashed line in Figure 19), suggesting that the sources missing due to incompleteness likely fall in the lowest two bins. Taking into account the small sample size, adding ∼ 2 sources to these two bins results in a distribution that appears approximately uniform all the way down to the sensitivity limit (denoted by the dotted line in Figure 19). This suggests that the intrinsic luminosity distribution may extend to lower luminosities than probed by these observations. However, even if all missed sources fall in the lowest bin of Figure 19, this portion of the intrinsic luminosity distribution (Lint ≤ 0.1 L⊙) would appear to be uniform rather than increasing to lower luminosities. We thus conclude that the instrinsic luminosity distribution does not continue to increase to our sensitivity limit, although K-S tests should be used to quantitatively confirm this result once a larger sample becomes available. This result provides an important constraint for future attempts at understanding these very low luminosity objects. However, we caution that we can draw no conclusions about the possible presence of objects with luminosities below ∼ 10−3 L⊙. To illustrate this with an example, Figure 20 shows the results of a Monte Carlo simulation where half of the 10,000 sources have luminosities evenly distributed between 10−4 and 10−3 L⊙, while the other half have luminosities evenly distributed between 10−3 and 10−1 L⊙. The simulated observed distribution matches the observed distribution quite well; the population of objects with Lint between 10−4 and 10−3 L⊙ have no effect on the final result since none of them are detected. A large population of embedded objects with such low luminosities would be very difficult to explain physically (see discussion of VeLLOs in §6.4), but we cannot rule out the presence of such objects based on these observations. B. Errors in Evolutionary Indicators From Incomplete, Finitely Sampled Spectral Energy Distributions The integrals defined in Equations 4 − 6 are evaluated using the trapezoid rule to integrate over the finitely sampled source SEDs. The contribution δi to the total integral – 43 – from measurements yi and yi+1, measured at xi and xi+1, is δi = (xi+1 − xi) (cid:18)yi+1 + yi 2 (cid:19) . (B1) The total value of the integral is then the sum of each individual δi. Clearly, the result will depend on how well the source SED is sampled, especially near the peak. Most of the material in the envelope of embedded objects is cold (10 − 15 K; e.g., Evans et al. 2001; Shirley et al. 2002). The emission peaks from blackbodies with temperatures of 10 and 20 K are located at approximately 500 and 250 µm, respectively, overlapping with the ∼ 50 − 300 µm far-infrared regime that cannot be observed from the ground. Incomplete sampling near the peak may thus introduce significant errors in the calculations of Lbol, Tbol, and Lbol/Lsmm. To quantify these errors, we consider the 13 group 1 candidates in Perseus. Most of these sources have well-sampled SEDs, with 2MASS and Spitzer detections from 1.25 − 70 µm, Spitzer detections at 160 µm and possibly IRAS detections at 100 µm, SHARC-II detections at 350 µm, SCUBA detections at 450 and 850 µm, and Bolocam detections at 1.1 mm. We construct four SEDs for each source: 1. We include all detections. These "category 1" SEDs are those that are the most completely sampled. 2. We only include detections in the wavelength ranges 1.25 ≤ λ ≤ 70 µm and λ ≥ 350 µm. These "category 2" SEDs are meant to simulate sources for which 2MASS and Spitzer 1 − 70 µm data, 350 or 450 µm submillimeter data, and millimeter wavelength data are available, but neither Spitzer 160 µm nor IRAS 100 µm data are available. 3. We only include detections in the wavelength ranges 1.25 ≤ λ ≤ 160 µm and λ ≥ 850 µm. These "category 3" SEDs are meant to simulate sources for which 2MASS and Spitzer 1.25 − 70 µm data, Spitzer 160 µm or IRAS 100 µm data, and millimeter wavelength data are available, but no 350 or 450 µm submillimeter data are available. 4. We only include detections in the wavelength ranges 1.25 ≤ λ ≤ 70 µm and λ ≥ 850 µm. These "category 4" SEDs are meant to simulate sources for which only 2MASS and Spitzer 1.25 − 70 µm and millimeter wavelength data are available. For each of the 13 sources, we calculate Lbol, Tbol, and Lbol/Lsmm for each of the four versions of the SED. Comparing the results will give an estimate of the errors introduced by incompletely sampling the source SED. – 44 – B.1. Bolometric Luminosity The three panels of figure 21 shows the percent error for each source between Lbol calculated from the SEDs in categories 2 − 4 and Lbol calculated from the SED in category 1. For category n, where n = 2, 3, 4, this percent error (P E) is defined as P E = 100 × Lbol,n − Lbol,1 Lbol,1 (B2) For all three cateogories, the calculated Lbol is an underestimate of the Lbol calculated from the more completely sampled Category 1 SEDs. The average of the absolute values of P E, ignoring values of 0 which simply indicate that the complete SED for the source is not a Category 1 SED, are 31%, 25%, and 54% for categories 2, 3, and 4, respectively. As expected, including detections either between 100 − 160 or 350 − 450 µm significantly improves the accuracy of Lbol; adding detections in both ranges improves it further. B.2. Bolometric Temperature The results of the same analysis for Tbol as that performed above for Lbol are shown in Figure 22. For category 3 and category 4 SEDs, the calculated Tbol is always higher than that calculated from a well-sampled SED. This is expected since both categories leave out detections near the peak of the SED. For category 2, however, the direction of the error in Tbol depends on the true value of Tbol. Category 2 SEDs of the coldest sources (Tbol . 50 K) will lead to underestimates of Tbol, whereas they will lead to overestimates for warmer sources (Tbol & 50 K). We calculate the percent error in Tbol in the same manner as above for Lbol. The average of the absolute values of these percent errors are 21%, 23%, and 64% for categories 2, 3, and 4, respectively. As for Lbol, category 2 and 3 SEDs yield significantly more accurate values of Tbol than category 4 SEDs, but are not as accurate as Category 1 SEDs. We note here that our results are in good agreement with Enoch (2007), who found that overall uncertainties for measured Lbol and Tbol values are approximately 20−50%, depending on whether or not 160 µm data are available. We also note that distance uncertainties will add additional uncertainties to Lbol beyond those considered here. – 45 – B.3. Bolometric to Submillimeter Luminosity We performed a similar anaysis for Lbol/Lsmm as above for Lbol and Tbol. The average values of the percent errors are 31%, 1282%, and 605% for categories 2, 3, and 4, respectively, significantly larger than the errors in Lbol and Tbol. The value of Lbol/Lsmm calculated for category 4 SEDs overestimates the actual value; both Lbol and Lsmm are underestimated by not sampling the SED at all between 70 and 850 µm, but Lbol is underestimated less than Lsmm. The same is true for category 3 SEDs, except the error is even larger because the calculation of Lbol is now less of an underestimate (see above), but the calculation of Lsmm remains the same. On the other hand, Lbol/Lsmm calculated from category 2 SEDs underes- timates the actual value, since Lbol is underestimated but Lsmm is accurately calculated. The errors here are more than an order of magnitude smaller than Categories 3 and 4. Clearly, great care must be taken to obtain the most well-sampled SEDs before attempting to draw conclusions from the value of Lbol/Lsmm for a given source. REFERENCES Acosta-Pulido, J. A., et al. 2007, AJ, 133, 2020 Alcal´a, J.M., et al. 2008, ApJ, 676, 427 Andr´e, P., Ward-Thompson, D., & Barsony, M. 1993, ApJ, 406, 122 Andr´e, P. & Montmerle, T. 1994, ApJ, 420, 837 Andr´e, P., Motte, F., & Bacmann, A. 1999. ApJ, 513, L57 Aspin, C., Sandell, G., & Russell, A.P.G. 1994, A&AS, 106, 165 Aspin, C. 2003, AJ, 125, 1480 Belloche, A., Andr´e, P., Despois, D., & Blinder, S. 2002, A&A, 393, 927 Benson, P.J., & Myers, P.C. 1989, ApJS, 71, 89 Black, J.H. 1994, ASP Conf. Ser. 58, The First Symposium on the Infrared Cirrus and Diffuse Enterstellar Clouds, ed. R.M. Cutri & W.B. Latter (San Francisco: ASP), 355 Boogert, A.C.A., et al. 2004, ApJS, 154, 359 Boss, A.P. & Yorke, H.W. 1995, ApJ, 439, L55 Bourke, T.L., Crapsi, A., Myers, P.C., Evans, N.J. II, Wilner, D.J., Huard, T.L., Jørgensen, J.K., & Young, C.H. 2005, ApJ, 633, L129 Bourke, T.L., et al. 2006, ApJ, 649, L37 – 46 – Brown, J.M., et al. 2007, ApJ, 664, L107 Butner, H.M., Natta, A., & Evans, N.J. II 1994, ApJ, 420, 326 Caselli, P., Benson, P.J., Myers, P.C., & Tafalla, M. 2002, ApJ, 572, 238 Chapman, N., et al. 2007, ApJ, 667, 288 Chen, H., Myers, P.C., Ladd, E.F., & Wood, D.O.S. 1995, ApJ, 445, 377 Chiang, E.I. & Goldreich, P. 1997, ApJ, 490, 368 Clark, F. O. 1991, ApJS, 75, 611 Crapsi, A., Caselli, P., Walmsley, C.M., Tafalla, M., Lee, C.W., Bourke, T.L., & Myers, P.C. 2004, A&A, 420, 957 Crapsi, A., et al. 2005a, A&A, 439, 1023 Crapsi, A., et al. 2005b, ApJ, 619, 379 Crapsi, A., et al. 2008, A&A, in press Davis, C.J., Matthews, H.E., Ray, T.P., Dent, W.R.F, & Richer, J.S. 1999, MNRAS, 309, 141 Di Francesco, J., Evans. N.J., II, Caselli, P., Myers, P.C., Shirley, Y., Aikawa, Y., & Tafalla, M. 2007, in Protostars and Planets V, ed. B. Reipurth, D. Jewitt, & K. Keil (Tucson: Univ. Arizona Press) Djupvik, A.A., Andr´e, P., Bontemps, S., Motte, F., Olofsson, G., G alfalk, M., & Flor´en, H.-G. 2006, A&A, 458, 789 Draine, B.T. 1978, ApJS, 36, 595 Draine, B.T. & Lee, H.M. 1984, ApJ, 285, 89 Dullemond, C.P., & Dominik, C. 2004, A&A, 417, 159 Dunham, M.M., et al. 2006, ApJ, 651, 945 Enoch, M.L., et al. 2006, ApJ, 638, 293 Enoch, M.L, Glenn, J., Evans, N.J. II, Sargent, A.I., Young, K.E., & Huard, T.L. 2007, ApJ, 666, 982 Enoch, M.L. 2007, Ph.D. thesis, California Institute of Technology Enoch, M.L., et al. 2008a, in preparation Enoch, M.L., Evans, N.J. II, Sargent, A.I., Glenn, J., Rosolowsky, E., & Myers, P. 2008b, ApJ, in press Evans, N.J. II, Rawlings, J.M.C., Shirley, Y.L., & Mundy, L.G. 2001, ApJ, 557, 193 – 47 – Evans, N.J. II., et al. 2003, PASP, 115, 965 Evans, N.J. II, et al. 2007, Final Delivery of Data from the c2d Legacy Project: IRAC and MIPS (Pasadena: SSC), http://ssc.spitzer.caltech.edu/legacy/all.html Fazio, G.G., et al. 2004, ApJS, 154, 10 Fedele, D., van den Ancker, M.E., Petr-Gotzens, M.G., & Rafanelli, P. 2007, A&A, 472, 207 Froebrich, D. 2005, ApJS, 156, 169 Harvey, P.M., et al. 2006, ApJ, 644, 307 Harvey, P.M., et al. 2007, ApJ, 663, 1139 Harvey, P.M., Merin, B., Huard, T.L., Rebull, L.M., Chapman, N., Evans, N.J. II, & Myers, P.C. 2007b, ApJ, 663, 1149 Hatchell, J., Richer, J.S., Fuller, G.A., Qualtrough, C.J., Ladd, E.F., & Chandler, C.J. 2005, A&A, 440, 151 Hatchell, J., Fuller, G.A., Richer, J.S., Harries, T.J., & Ladd, E.F. 2007a, A&A, 468, 1009 Hatchell, J., Fuller, G.A., & Richer, J.S. 2007b, A&A, 472, 187 Hartmann, L., & Kenyon, S.J. 1985, ApJ, 299, 462 Henning, T., Pfau, W., Zinnecker, H., & Prusti, T. 1993, A&A, 276, 129 Ho, P.T.P., & Townes, C.H. 1983, ARA&A, 21, 239 Hodapp, K.W., Bally, J., Eisloffel, J., & Davis, C.J. 2005, AJ, 129, 1580 Huard, T.L., et al. 2006, ApJ, 640, 391 Jijina, J., Myers, P.C., & Adams, F.C. 1999, ApJS, 125, 161 Johnstone, D., Wilson, C.D., Moriarty-Schieven, G., Joncas, G., Smith, G., Gregersen, E., & Fich, M. 2000, ApJ, 545, 327 Jørgensen, J.K., et al. 2006, ApJ, 645, 1246 Jørgensen, J.K., Johnstone, D., Kirk, H., & Myers, P.C. 2007, ApJ, 656, 293 Kauffmann, J.,Bertoldi, F., Evans, N.J., & the c2d Collaboration 2005, Astron. Nachr., 326, 878 Kauffmann, J., Bertoldi, F., Bourke, T.L., Evans, N.J. II., & Lee, C.W. 2008, A&A, sub- mitted Kenyon, S.J., & Hartmann, L. 1995, ApJS, 101, 117 Kirk, J.M., Ward-Thompson, D., & Andr´e, P. 2005, MNRAS, 360, 1506 Kirk, H., Johnstone, D., & Di Francesco, J. 2006, ApJ, 646, 1009 – 48 – Kirk, J.M., Ward-Thompson, D., & Andr´e, P. 2007, MNRAS, 375, 843 K´osp´al, ´A, ´Abrah´am, P., Prusti, T., Acosta-Pulido, J., Hony, S., Mo´or, A., & Siebenmorgen, R. 2007, A&A, 470, 211 Kun, M. 1998, ApJS, 115, 59 Lada, C.J. 1987, Star Forming Regions, M. Peimbert & J. Jugaku (eds.), IAU, 1 Lee, C.-F., Ho, P.T.P., Palau, A., Hirano, N., Bourke, T.L., Shang, H., & Zhang, Q. 2007, ApJ, 670, 1188 Lee, J.-E., et al. 2006, ApJ, 648, 491 Lee, J.-E. 2008, JKAS, in press, arXiv:0712.1866 Lehtinen, K., Mattila, K., & Lemke, D. 2005, A&A, 437, 159 McKee, C.F. & Ostriker, E.C. 2007, ARA&A, in press Mer´ın, B., et al. 2008, ApJ, in press Motte, F., & Andr´e P. 2001, A&A, 365, 440 Myers, P.C., Fuller, G.A., Mathieu, R.D., Beichman, C.A., Benson, P.J., Schild, R.E., & Emerson, J.P. 1987, ApJ, 319, 340 Myers, P.C. & Ladd, E.F. 1993, ApJ, 413, 47 Myers, P.C., Adams, F.C., Chen, H., & Schaff, E. 1998, ApJ, 492, 703 Noriega-Crespo, A., et al. 2004, ApJS, 154, 352 Padgett, D., et al. 2008, ApJ, 672, 1013 Porras, A., et al. 2007, ApJ, 656, 493 Rebull, L.M., et al. 2007, ApJS, 171, 447 Rieke, G.H., et al. 2004, ApJS, 154, 25 Robitaille, T.P., Whitney, B.A., Indebetouw, R., Wood, K., & Denzmore, P. 2006, ApJS, 167, 256 Schechter, P.S., Mateo, M., & Saha, A. 1993, PASP, 105, 1342 Shirley, Y.L., Evans, N.J. II, & Rawlings, J.M.C. 2002, ApJ, 575, 337 Shu, F.H., Adams, F.C., & Lizano, S. 1987, ARA&A, 25, 23 Stanke, T., Smith, M.D., Gredel, R., & Khanzadyan, T. 2006, A&A, 447, 609 Stutz, A.M., et al. 2007, ApJ, 665, 466 Tachihara, K, et al. 2007, ApJ, 659, 1382 – 49 – Tassis, K., & Mouschovias, T.C. 2005, ApJ, 618, 783 Vorobyov, E.I., & Basu, S. 2005, ApJ, 633, L137 Vorobyov, E.I., & Basu, S. 2006, ApJ, 650, 956 Ward-Thompson, D., et al. 2007, PASP, 119, 855 Whitney, B.A., Wood, K., Bjorkman, J.E., & Wolff, M.J. 2003a, ApJ, 591, 1049 Whitney, B.A., Wood, K., Bjorkman, J.E., & Cohen, M. 2003b, ApJ, 598, 1079 Wu, J., Dunham, M.M., Evans, N.J. II, Bourke, T.L., & Young, C.H. 2007, AJ, 133, 1560 Young, C.H., Shirley, Y.L., & Evans, N.J. 2003, ApJS, 145, 111 Young, C.H., et al. 2004, ApJS, 154, 396 Young, C.H., & Evans, N.J. 2005, ApJ, 627, 293 Young, C.H., et al. 2006a, AJ, 132, 1998 Young, K.E., et al. 2005, ApJ, 628, 283 Young, K.E., et al. 2006b, ApJ, 644, 326 This preprint was prepared with the AAS LATEX macros v5.2. – 50 – Table 1. Embedded Protostars Observed by Spitzer with Published Models Protostar LIR (L⊙) Lint (L⊙) Lbol (L⊙) Referencea CB130-3-IRS IRAM 04191+1522 IRAS 04191+1523 IRAS 18148−0440 IRAS 19156+1906 L1014-IRS L1148-IRS L1221-IRS1 L1221-IRS3 L1521F-IRS L673-7-IRS 0.05 0.02 0.25 8.19 1.13 0.06 0.08 1.12 0.36 0.01 0.02 0.15 ± 0.05 0.08 ± 0.04 0.4 ± 0.1 13.0 ± 0.5 2.6 ± 0.5 0.09 ± 0.03 0.1 ± 0.05 2.5 ± 0.25 1.0 ± 0.15 0.05 ± 0.02 0.04 ± 0.02 0.96 0.28 0.64 12.9 3.4 0.3 0.15 2.7 1.9 0.36 0.13 1 2 3 4 4 5 6 7 7 8 3 aReference for the radiative transfer model and determination of Lint and Lbol: (1) H.J. Kim et al. (2008, in preparation); (2) Dunham et al. (2006); (3) M. Dunham et al. in preparation); (4) Shirley et al. (2002); (5) Young et al. (2004); (6) Kauffmann et al. (2005); (7) C. Young et al. (2008, in preparation); (8) Bourke et al. (2006.). (2008, Table 2. Linear Least-Squares Fits to Plots of Log Fν vs. Log Lint. Wavelength (µm) Observations m b 3.6 4.5 5.8 8.0 24 70 0.57 ± 0.42 −11.07 ± 0.32 0.88 ± 0.26 −10.54 ± 0.20 0.97 ± 0.29 −10.54 ± 0.22 0.83 ± 0.27 −10.72 ± 0.22 0.87 ± 0.20 −10.05 ± 0.17 −9.26 ± 0.06 1.06 ± 0.06 m Models b 0.86 ± 0.05 −11.87 ± 0.03 0.94 ± 0.04 −11.63 ± 0.03 0.99 ± 0.04 −11.43 ± 0.03 1.02 ± 0.04 −11.38 ± 0.03 1.38 ± 0.02 −10.12 ± 0.01 −9.02 ± 0.01 1.06 ± 0.01 χ2 r 296 148 179 160 85 3 χ2 r 3309 2846 2305 2044 433 15 Table 3. Candidate Low-Luminosity Embedded Objects in the Cores Source Number Groupa Source Nameb Core Regionc Spitzer c2d Dist. (pc) RA (J2000) Dec (J2000) LIR (L⊙) d DIRAS 001 002 003 004 005 006 007 008 009 010 011 012 013 014 015 016 017 018 019 020 021 022 023 024 025 026 027 028 029 030 031 032 033 034 1 1 1 1 1 1 6 5 5 5 3 5 5 6 5 2 1 3 5 5 1 6 3 1 1 6 1 5 1 3 1 3 6 6 IRAM04191+1522 IRAM04191+1522 IRAM04191+1522 L1521F TMC1 B35A B35A BHR22 BHR16 CG30−31 J042156.88+152946.0 J042200.07+153024.8 J042200.41+153021.2 J042838.90+265135.6 J044112.65+254635.4 J054429.26+090856.8 J054443.94+090307.2 J071357.60−482711.9 J080533.05−390924.8 J080850.47−360652.2 J081705.26−395415.8 DC2573−25 J081706.34−394635.8 DC2573−25 J081710.92−395244.4 DC2573−25 J092851.50−513558.6 DC2742−04 J123040.90−710007.6 Mu8 J124539.96−552522.4 DC302.1+7.4 J130736.89−770009.7 DC303.8−14.2 J154216.99−524802.2 DC3272+18 J155856.14−373354.0 DC3391+117−2 J163705.11−353219.7 DC3460+78 J165719.63−160923.4 J171111.83−272655.0 J171122.18−272602.0 J181616.39−023237.7 J181659.47−180230.5 J191744.28+191523.7 J192025.32+112217.4 J192025.92+112221.0 J192026.16+111949.1 J192026.54+112025.4 J192134.82+112123.4 J204056.66+672304.9 J204105.95+671820.9 J204355.51+673850.3 CB68 B59 B59 CB130−3 L328 L723 L673 L673 L673 L673 L673−7 L1148 L1148 L1155E 140 ± 10 140 ± 10 140 ± 10 140 ± 10 140 ± 10 400 ± 40 400 ± 40 450 ± 50 440 ± 100 450 ± 50 440 ± 100 440 ± 100 440 ± 100 130 ± 10 150 ± 30 300 ± 50 178 ± 18 250 ± 50 150 ± 20 150 ± 20 125 ± 25 125 ± 25 125 ± 25 270 ± 50 270 ± 50 300 ± 100 300 ± 100 300 ± 100 300 ± 100 300 ± 100 300 ± 100 325 ± 25 325 ± 25 325 ± 25 04 21 56.88 +15 29 46.0 04 22 00.07 +15 30 24.8 04 22 00.41 +15 30 21.2 04 28 38.90 +26 51 35.6 04 41 12.65 +25 46 35.4 05 44 29.26 +09 08 56.8 05 44 43.94 +09 03 07.2 07 13 57.60 −48 27 11.9 08 05 33.05 −39 09 24.8 08 08 50.47 −36 06 52.2 08 17 05.26 −39 54 15.8 08 17 06.34 −39 46 35.8 08 17 10.92 −39 52 44.4 09 28 51.50 −51 35 58.6 12 30 40.90 −71 00 07.6 12 45 39.96 −55 25 22.4 13 07 36.89 −77 00 09.7 15 42 16.99 −52 48 02.2 15 58 56.14 −37 33 54.0 16 37 05.11 −35 32 19.7 16 57 19.63 −16 09 23.4 17 11 11.83 −27 26 55.0 17 11 22.18 −27 26 02.0 18 16 16.39 −02 32 37.7 18 16 59.47 −18 02 30.5 19 17 44.28 +19 15 23.7 19 20 25.32 +11 22 17.4 19 20 25.92 +11 22 21.0 19 20 26.16 +11 19 49.1 19 20 26.54 +11 20 25.4 19 21 34.82 +11 21 23.4 20 40 56.66 +67 23 04.9 20 41 05.95 +67 18 20.9 20 43 55.51 +67 38 50.3 0.023 0.033 0.250 0.015 0.383 0.141 0.006 0.006 0.081 0.064 0.193 0.013 0.018 0.005 0.001 0.232 0.387 0.034 0.005 0.003 0.299 0.024 0.343 0.054 0.065 0.048 0.201 0.168 0.037 0.138 0.017 0.081 0.003 0.004 (′′) 43.6 17.4 20.8 380.9 15.7 8.1 405.4 170.7 241.9 525.5 51.1 194.3 161.0 712.4 413.5 44.5 5.7 337.1 331.8 221.3 2.0 125.9 86.3 905.7 445.0 225.4 42.9 52.5 5.3 35.4 427.6 222.3 250.9 642.1 Nearest IRAS Source IRAS 04191+1523 IRAS 04191+1523 IRAS 04191+1523 IRAS 04260+2642 IRAS 04381+2540 IRAS 05417+0907 IRAS 05417+0907 IRAS 07128−4820 IRAS 08036−3904 IRAS 08076−3556 IRAS 08152−3945 IRAS 08151−3934 IRAS 08152−3945 IRAS 09274−5134 IRAS 12288−7047 IRAS 12427−5508 IRAS 13036−7644 IRAS 15387−5233 IRAS 15552−3727 IRAS 16335−3528 IRAS 16544−1604 IRAS 17082−2724 IRAS 17081−2721 IRAS 18144−0242 IRAS 18145−1801 IRAS 19156+1906 IRAS 19180+1116 IRAS 19180+1116 IRAS 19180+1114 IRAS 19180+1114 IRAS 19189+1109 IRAS 20410+6710 IRAS 20410+6710 IRAS 20423+6736 – 5 1 – Source Number Groupa Source Nameb Core Regionc Spitzer c2d Dist. (pc) RA (J2000) Dec (J2000) Table 3-Continued d DIRAS LIR (L⊙) 035 036 037 038 039 040 041 042 043 044 045 046 047 048 049 6 6 6 1 6 6 1 6 3 1 1 6 6 1 1 J204427.17+673835.9 J205706.72+773656.2 J205707.85+773659.8 J212407.58+495908.9 J220633.22+590232.6 J220637.27+590315.8 J222807.42+690038.9 J222933.39+751316.0 J222959.52+751403.1 J223031.94+751408.9 J223105.59+751337.2 J223514.06+751502.5 J223731.13+751041.5 J223846.15+751132.3 J223846.44+751128.0 L1155E L1228 L1228 L1014 L1165 L1165 L1221 L1251 L1251 L1251 L1251 L1251 L1251 L1251 L1251 325 ± 25 200 ± 50 200 ± 50 250 ± 50 300 ± 50 300 ± 50 250 ± 50 300 ± 50 300 ± 50 300 ± 50 300 ± 50 300 ± 50 300 ± 50 300 ± 50 300 ± 50 20 44 27.17 +67 38 35.9 20 57 06.72 +77 36 56.2 20 57 07.85 +77 36 59.8 21 24 07.58 +49 59 08.9 22 06 33.22 +59 02 32.6 22 06 37.27 +59 03 15.8 22 28 07.42 +69 00 38.9 22 29 33.39 +75 13 16.0 22 29 59.52 +75 14 03.1 22 30 31.94 +75 14 08.9 22 31 05.59 +75 13 37.2 22 35 14.06 +75 15 02.5 22 37 31.13 +75 10 41.5 22 38 46.15 +75 11 32.3 22 38 46.44 +75 11 28.0 0.004 0.199 0.038 0.087 0.030 0.008 0.404 0.066 0.272 0.156 0.077 0.030 0.004 0.274 0.081 aGroup indicating likelihood of being an embedded protostar (see §5.2). bAll source names are preceded by the prefix "SSTc2d ". cName of the core region from Evans et al. (2007) in which this source is located. (′′) 768.6 72.7 75.4 246.8 135.6 107.5 41.9 119.7 13.1 120.5 252.0 129.3 295.5 5.2 2.7 Nearest IRAS Source IRAS 20423+6736 IRAS 20582+7724 IRAS 20582+7724 IRAS 21220+4944 IRAS 22051+5848 IRAS 22051+5848 IRAS 22266+6845 IRAS 22290+7458 IRAS 22290+7458 IRAS 22290+7458 IRAS 22290+7458 IRAS 22343+7501 IRAS 22376+7455 IRAS 22376+7455 IRAS 22376+7455 – 5 2 – dIntegrated luminosity using all available detections between 1.25 (2MASS J-band) and 70 µm. Entries in italics denote that the calculation only extends to 24 µm due to no flux information available at 70 µm. – 53 – Table 4. Candidate Low-Luminosity Embedded Objects in Perseus Source Number Groupa Source Nameb Spitzer RA (J2000) Dec (J2000) 050 051 052 053 054 055 056 057 058 059 060 061 062 063 064 065 066 067 068 069 070 071 072 073 074 075 076 077 078 079 080 081 082 083 084 085 086 087 088 089 090 091 092 093 094 5 5 5 5 5 1 1 6 5 5 3 5 5 1 1 1 6 5 1 6 6 1 6 1 2 2 6 6 2 5 3 1 6 6 1 6 6 5 1 5 1 6 1 1 5 J032353.74+310244.5 J032430.79+310651.8 J032433.14+311433.0 J032503.14+310322.7 J032515.26+312511.3 J032536.22+304515.8 J032539.12+304358.1 J032621.65+305726.3 J032627.62+293002.2 J032629.33+312154.4 J032637.46+301528.1 J032655.27+312831.4 J032705.62+294601.9 J032738.26+301358.8 J032832.57+311105.3 J032839.10+310601.8 J032843.58+311736.2 J032843.70+291050.2 J032845.29+310542.0 J032856.33+312227.8 J032856.59+310736.8 J032900.55+311200.7 J032906.05+303039.2 J032912.07+311301.6 J032913.54+311358.1 J032917.16+312746.4 J032919.75+311256.9 J032920.35+311250.4 J032923.47+313329.5 J032950.42+300328.8 J032951.82+313906.1 J033032.69+302626.5 J033047.81+313906.8 J033116.90+302958.2 J033120.98+304530.2 J033146.63+302440.0 J033147.33+302438.5 J033149.90+295423.8 J033217.95+304947.6 J033225.70+300631.3 J033229.18+310240.9 J033257.84+310608.3 J033314.38+310710.9 J033316.44+310652.6 J033357.17+314330.0 03 23 53.74 +31 02 44.5 03 24 30.79 +31 06 51.8 03 24 33.14 +31 14 33.0 03 25 03.14 +31 03 22.7 03 25 15.26 +31 25 11.3 03 25 36.22 +30 45 15.8 03 25 39.12 +30 43 58.1 03 26 21.65 +30 57 26.3 03 26 27.62 +29 30 02.2 03 26 29.33 +31 21 54.4 03 26 37.46 +30 15 28.1 03 26 55.27 +31 28 31.4 03 27 05.62 +29 46 01.9 03 27 38.26 +30 13 58.8 03 28 32.57 +31 11 05.3 03 28 39.10 +31 06 01.8 03 28 43.58 +31 17 36.2 03 28 43.70 +29 10 50.2 03 28 45.29 +31 05 42.0 03 28 56.33 +31 22 27.8 03 28 56.59 +31 07 36.8 03 29 00.55 +31 12 00.7 03 29 06.05 +30 30 39.2 03 29 12.07 +31 13 01.6 03 29 13.54 +31 13 58.1 03 29 17.16 +31 27 46.4 03 29 19.75 +31 12 56.9 03 29 20.35 +31 12 50.4 03 29 23.47 +31 33 29.5 03 29 50.42 +30 03 28.8 03 29 51.82 +31 39 06.1 03 30 32.69 +30 26 26.5 03 30 47.81 +31 39 06.8 03 31 16.90 +30 29 58.2 03 31 20.98 +30 45 30.2 03 31 46.63 +30 24 40.0 03 31 47.33 +30 24 38.5 03 31 49.90 +29 54 23.8 03 32 17.95 +30 49 47.6 03 32 25.70 +30 06 31.3 03 32 29.18 +31 02 40.9 03 32 57.84 +31 06 08.3 03 33 14.38 +31 07 10.9 03 33 16.44 +31 06 52.6 03 33 57.17 +31 43 30.0 c LIR (L⊙) 0.010 0.017 0.008 0.015 0.008 0.061 0.432 0.006 0.016 0.018 0.462 0.008 0.008 0.347 0.036 0.015 0.414 0.010 0.185 0.119 0.014 0.105 0.019 0.018 0.305 0.183 0.002 0.020 0.149 0.010 0.166 0.033 0.009 0.004 0.299 0.015 0.002 0.014 0.135 0.007 0.180 0.012 0.098 0.144 0.050 DIRAS (′′) Nearest IRAS Source 440.4 419.3 407.6 178.7 689.4 4.3 90.4 281.5 932.8 434.6 5.8 227.9 133.6 60.0 163.6 72.7 15.8 757.9 10.5 182.5 194.6 170.3 212.5 217.2 183.8 184.3 283.7 293.9 9.3 922.0 3.3 147.0 402.2 383.7 9.6 209.2 211.3 642.7 6.2 409.5 18.4 258.9 47.3 58.4 461.0 IRAS 03203+3048 IRAS 03219+3056 IRAS 03219+3103 IRAS 03222+3053 IRAS 03219+3103 IRAS 03225+3034 IRAS 03225+3034 IRAS 03232+3051 IRAS 03239+2905 IRAS 03235+3118 IRAS 03235+3004 IRAS 03235+3118 IRAS 03239+2933 IRAS 03245+3002 IRAS 03255+3103 IRAS 03256+3055 IRAS 03256+3107 IRAS 03253+2848 IRAS 03256+3055 IRAS 03260+3111 IRAS 03256+3055 IRAS 03258+3104 IRAS 03258+3023 IRAS 03259+3105 IRAS 03259+3105 IRAS 03262+3114 IRAS 03259+3105 IRAS 03259+3105 IRAS 03262+3123 IRAS 03279+2950 IRAS 03267+3128 IRAS 03273+3018 IRAS 03279+3123 IRAS 03286+3017 IRAS 03282+3035 IRAS 03286+3017 IRAS 03286+3017 IRAS 03295+2947 IRAS 03292+3039 IRAS 03297+2951 IRAS 03293+3052 IRAS 03301+3057 IRAS 03301+3057 IRAS 03301+3057 IRAS 03302+3131 – 54 – Table 4-Continued Source Number Groupa Source Nameb Spitzer RA (J2000) Dec (J2000) 095 096 097 098 099 100 101 102 103 104 105 106 107 108 109 110 111 112 113 114 115 116 117 118 119 5 6 6 6 6 6 5 5 5 3 1 1 3 3 3 6 6 6 5 6 5 6 6 5 6 J033544.18+302357.5 J033554.41+304501.1 J033626.54+313634.9 J033745.17+305445.7 J033812.17+314020.6 J033925.54+321708.2 J034010.20+304446.7 J034015.79+305504.8 J034227.67+310145.1 J034351.02+320307.9 J034356.52+320052.9 J034356.83+320304.7 J034402.40+320204.9 J034402.64+320159.5 J034421.36+315932.6 J034511.57+322753.3 J034513.82+321210.1 J034527.98+323635.6 J034530.41+324801.8 J034632.45+313443.7 J034635.38+330746.6 J034656.74+325248.4 J034734.34+324640.1 J034754.26+331512.2 J034910.08+323249.2 03 35 44.18 +30 23 57.5 03 35 54.41 +30 45 01.1 03 36 26.54 +31 36 34.9 03 37 45.17 +30 54 45.7 03 38 12.17 +31 40 20.6 03 39 25.54 +32 17 08.2 03 40 10.20 +30 44 46.7 03 40 15.79 +30 55 04.8 03 42 27.67 +31 01 45.1 03 43 51.02 +32 03 07.9 03 43 56.52 +32 00 52.9 03 43 56.83 +32 03 04.7 03 44 02.40 +32 02 04.9 03 44 02.64 +32 01 59.5 03 44 21.36 +31 59 32.6 03 45 11.57 +32 27 53.3 03 45 13.82 +32 12 10.1 03 45 27.98 +32 36 35.6 03 45 30.41 +32 48 01.8 03 46 32.45 +31 34 43.7 03 46 35.38 +33 07 46.6 03 46 56.74 +32 52 48.4 03 47 34.34 +32 46 40.1 03 47 54.26 +33 15 12.2 03 49 10.08 +32 32 49.2 c LIR (L⊙) 0.057 0.067 0.026 0.029 0.019 0.292 0.008 0.013 0.019 0.172 0.310 0.253 0.101 0.021 0.200 0.026 0.051 0.011 0.010 0.018 0.012 0.015 0.006 0.013 0.010 DIRAS (′′) Nearest IRAS Source 432.3 13.1 381.9 694.5 683.5 28.8 652.3 326.0 641.4 99.9 59.0 105.2 124.0 122.0 171.1 512.3 311.9 185.7 247.3 223.8 198.8 549.0 314.5 229.8 594.5 IRAS 03324+3020 IRAS 03328+3035 IRAS 03335+3120 IRAS 03337+3043 IRAS 03356+3121 IRAS 03363+3207 IRAS 03375+3043 IRAS 03375+3043 IRAS 03399+3059 IRAS 03407+3152 IRAS 03407+3152 IRAS 03407+3152 IRAS 03407+3152 IRAS 03410+3152 IRAS 03410+3152 IRAS 03426+3214 IRAS 03424+3203 IRAS 03421+3229 IRAS 03424+3234 IRAS 03436+3123 IRAS 03436+3300 IRAS 03434+3235 IRAS 03445+3242 IRAS 03448+3302 IRAS 03454+3230 aGroup indicating likelihood of being an embedded protostar (see §5.2). bAll source names are preceded by the prefix "SSTc2d ". cIntegrated luminosity using all available detections between 1.25 (2MASS J-band) and 70 µm. Entries in italics denote that the calculation only extends to 24 µm due to no flux information available at 70 µm. – 55 – Table 5. Candidate Low-Luminosity Embedded Objects in Chamaeleon II Source Number Groupa Source Nameb Spitzer RA (J2000) Dec (J2000) LIR (L⊙) c DIRAS 120 121 122 123 124 125 126 127 128 129 130 131 132 133 134 135 5 5 5 5 3 5 5 5 5 5 5 5 1 5 5 5 J124657.72−771916.0 J124816.42−781511.5 J124940.03−775657.1 J125254.34−780913.7 J125342.86−771511.5 J125507.66−765755.4 J125701.58−764834.9 J130100.05−770453.0 J130242.22−764402.0 J130326.59−764502.9 J130509.53−770716.3 J130515.96−771115.7 J130737.22−770009.4 J130832.38−773226.2 J131031.92−772423.8 J131331.97−773438.3 12 46 57.72 −77 19 16.0 12 48 16.42 −78 15 11.5 12 49 40.03 −77 56 57.1 12 52 54.34 −78 09 13.7 12 53 42.86 −77 15 11.5 12 55 07.66 −76 57 55.4 12 57 01.58 −76 48 34.9 13 01 00.05 −77 04 53.0 13 02 42.22 −76 44 02.0 13 03 26.59 −76 45 02.9 13 05 09.53 −77 07 16.3 13 05 15.96 −77 11 15.7 13 07 37.22 −77 00 09.4 13 08 32.38 −77 32 26.2 13 10 31.92 −77 24 23.8 13 13 31.97 −77 34 38.3 0.004 0.012 0.006 0.010 0.184 0.006 0.050 0.003 0.003 0.020 0.005 0.003 0.361 0.003 0.008 0.018 aGroup indicating likelihood of being an embedded protostar (see §5.2). bAll source names are preceded by the prefix "SSTc2d ". (′′) 372.4 896.9 897.8 483.2 5.7 182.1 10.5 347.6 396.9 355.1 203.0 336.8 6.2 308.5 694.4 1271.6 Nearest IRAS Source IRAS 12416−7703 IRAS 12440−7813 IRAS 12504−7745 IRAS 12504−7745 IRAS 12500−7658 IRAS 12522−7640 IRAS 12533−7632 IRAS 12571−7654 IRAS 12584−7621 IRAS 13005−7633 IRAS 13022−7650 IRAS 13022−7650 IRAS 13036−7644 IRAS 13031−7714 IRAS 13030−7707 IRAS 13031−7714 cIntegrated luminosity using all available detections between 1.25 (2MASS J-band) and 70 µm. Entries in italics denote that the calculation only extends to 24 µm due to no flux information available at 70 µm. – 56 – Table 6. Candidate Low-Luminosity Embedded Objects in Lupus Source Number Groupa Source Nameb Regionc Spitzer RA (J2000) Dec (J2000) LIR (L⊙) d DIRAS 136 137 138 139 140 141 142 143 144 145 146 147 148 149 150 151 152 153 154 155 156 157 158 159 160 161 162 163 164 165 166 167 168 169 170 171 172 173 174 175 176 177 5 5 5 5 5 5 5 5 5 4c 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 4a 5 5 5 5 5 5 5 6 1 5 5 6 5 5 5 5 J153706.36−334102.4 J153751.12−335806.6 J153800.05−340524.4 J153826.95−325628.7 J153833.84−334030.0 J153834.66−344819.4 J153914.59−341105.3 J153953.90−334821.6 J154017.64−324649.4 J154051.62−342104.7 J154125.92−344445.6 J154126.95−325204.1 J154138.54−325005.3 J154141.93−324025.7 J154148.05−333536.6 J154302.28−344406.4 J154339.98−335554.8 J154421.91−332303.8 J154507.97−334254.7 J154513.15−335524.2 J154548.24−340510.7 J154601.18−335816.7 J154612.58−340250.3 J160003.79−421024.2 J160020.16−410333.8 J160115.55−415235.4 J160132.74−423628.4 J160709.79−392817.0 J160727.19−383842.0 J160729.66−391634.0 J160740.42−384602.3 J160741.50−383614.0 J160754.74−391544.6 J160851.43−390530.5 J160918.07−390453.4 J160934.85−384654.1 J160950.09−380756.3 J160954.82−391122.6 J161002.52−383025.9 J161022.30−373618.4 J161110.30−372008.9 J161316.27−384827.7 15 37 06.36 −33 41 02.4 Lupus I 15 37 51.12 −33 58 06.6 Lupus I 15 38 00.05 −34 05 24.4 Lupus I 15 38 26.95 −32 56 28.7 Lupus I 15 38 33.84 −33 40 30.0 Lupus I 15 38 34.66 −34 48 19.4 Lupus I 15 39 14.59 −34 11 05.3 Lupus I 15 39 53.90 −33 48 21.6 Lupus I 15 40 17.64 −32 46 49.4 Lupus I 15 40 51.62 −34 21 04.7 Lupus I 15 41 25.92 −34 44 45.6 Lupus I 15 41 26.95 −32 52 04.1 Lupus I 15 41 38.54 −32 50 05.3 Lupus I 15 41 41.93 −32 40 25.7 Lupus I 15 41 48.05 −33 35 36.6 Lupus I 15 43 02.28 −34 44 06.4 Lupus I 15 43 39.98 −33 55 54.8 Lupus I 15 44 21.91 −33 23 03.8 Lupus I 15 45 07.97 −33 42 54.7 Lupus I 15 45 13.15 −33 55 24.2 Lupus I 15 45 48.24 −34 05 10.7 Lupus I 15 46 01.18 −33 58 16.7 Lupus I Lupus I 15 46 12.58 −34 02 50.3 Lupus IV 16 00 03.79 −42 10 24.2 Lupus IV 16 00 20.16 −41 03 33.8 Lupus IV 16 01 15.55 −41 52 35.4 Lupus IV 16 01 32.74 −42 36 28.4 16 07 09.79 −39 28 17.0 Lupus III Lupus III 16 07 27.19 −38 38 42.0 16 07 29.66 −39 16 34.0 Lupus III 16 07 40.42 −38 46 02.3 Lupus III 16 07 41.50 −38 36 14.0 Lupus III 16 07 54.74 −39 15 44.6 Lupus III Lupus III 16 08 51.43 −39 05 30.5 16 09 18.07 −39 04 53.4 Lupus III 16 09 34.85 −38 46 54.1 Lupus III 16 09 50.09 −38 07 56.3 Lupus III 16 09 54.82 −39 11 22.6 Lupus III Lupus III 16 10 02.52 −38 30 25.9 16 10 22.30 −37 36 18.4 Lupus III 16 11 10.30 −37 20 08.9 Lupus III Lupus III 16 13 16.27 −38 48 27.7 0.004 0.004 0.004 0.003 0.004 0.003 0.003 0.005 0.006 0.011 0.001 0.004 0.005 0.008 0.003 0.007 0.003 0.008 0.006 0.006 0.006 0.007 0.004 0.006 0.003 0.049 0.006 0.008 0.015 0.009 0.009 0.018 0.012 0.035 0.134 0.004 0.014 0.009 0.021 0.009 0.008 0.006 aGroup indicating likelihood of being an embedded protostar (see §5.2). bAll source names are preceded by the prefix "SSTc2d ". cName of the Lupus region (I, III, or IV) in which this source is located. (′′) 682.1 680.6 709.8 178.5 511.7 486.1 532.3 186.7 926.8 11.8 659.8 845.4 733.6 192.0 203.4 559.0 603.3 196.2 375.2 1110.8 862.0 1302.4 1152.9 487.1 225.4 390.2 637.5 894.9 960.9 401.1 1081.8 737.8 411.1 52.1 19.3 715.3 312.3 63.2 679.7 611.7 897.5 531.3 Nearest IRAS Source IRAS 15334−3340 IRAS 15337−3350 IRAS 15349−3407 IRAS 15350−3246 IRAS 15349−3337 IRAS 15358−3444 IRAS 15367−3405 IRAS 15364−3339 IRAS 15364−3250 IRAS 15376−3411 IRAS 15379−3445 IRAS 15375−3253 IRAS 15387−3228 IRAS 15387−3228 IRAS 15384−3327 IRAS 15399−3425 IRAS 15397−3350 IRAS 15409−3312 IRAS 15417−3327 IRAS 15417−3327 IRAS 15420−3408 IRAS 15420−3408 IRAS 15420−3408 IRAS 15570−4155 IRAS 15571−4052 IRAS 15573−4147 IRAS 15571−4230 IRAS 16043−3933 IRAS 16051−3820 IRAS 16042−3901 IRAS 16036−3854 IRAS 16051−3820 IRAS 16042−3901 IRAS 16054−3857 IRAS 16059−3857 IRAS 16059−3850 IRAS 16063−3755 IRAS 16064−3903 IRAS 16069−3811 IRAS 16072−3738 IRAS 16084−3725 IRAS 16106−3842 – 57 – dIntegrated luminosity using all available detections between 1.25 (2MASS J-band) and 70 µm. Entries in italics denote that the calculation only extends to 24 µm due to no flux information available at 70 µm. – 58 – Table 7. Candidate Low-Luminosity Embedded Objects in Ophiuchus Source Number Groupa Source Nameb Spitzer RA (J2000) Dec (J2000) 178 179 180 181 182 183 184 185 186 187 188 189 190 191 192 193 194 195 196 197 198 199 5 5 5 3 1 3 3 1 5 5 6 6 5 5 5 5 5 5 5 5 5 5 J162135.57−224351.6 J162145.12−234231.7 J162458.99−252119.1 J162648.48−242838.6 J162705.23−243629.5 J162715.89−243843.1 J162717.57−242856.3 J162821.60−243623.4 J162843.68−231918.1 J163122.87−231655.2 J163131.25−242628.0 J163136.77−240419.9 J163307.27−225041.6 J163502.45−235100.7 J164001.80−244429.0 J164114.35−243758.1 J164147.64−243030.2 J164206.74−233859.3 J164301.87−245420.9 J164305.88−242357.5 J164433.26−233348.2 J164512.50−233848.1 16 21 35.57 −22 43 51.6 16 21 45.12 −23 42 31.7 16 24 58.99 −25 21 19.1 16 26 48.48 −24 28 38.6 16 27 05.23 −24 36 29.5 16 27 15.89 −24 38 43.1 16 27 17.57 −24 28 56.3 16 28 21.60 −24 36 23.4 16 28 43.68 −23 19 18.1 16 31 22.87 −23 16 55.2 16 31 31.25 −24 26 28.0 16 31 36.77 −24 04 19.9 16 33 07.27 −22 50 41.6 16 35 02.45 −23 51 00.7 16 40 01.80 −24 44 29.0 16 41 14.35 −24 37 58.1 16 41 47.64 −24 30 30.2 16 42 06.74 −23 38 59.3 16 43 01.87 −24 54 20.9 16 43 05.88 −24 23 57.5 16 44 33.26 −23 33 48.2 16 45 12.50 −23 38 48.1 c LIR (L⊙) 0.004 0.033 0.012 0.103 0.107 0.440 0.374 0.046 0.000 0.003 0.013 0.036 0.034 0.006 0.004 0.007 0.005 0.005 0.002 0.005 0.005 0.016 DIRAS (′′) Nearest IRAS Source 1467.3 246.9 6.6 366.6 55.0 163.5 10.4 11.6 516.2 1354.8 70.9 51.1 15.0 1074.3 447.9 539.7 591.5 538.5 295.6 345.0 329.6 375.3 IRAS 16200−2251 IRAS 16187−2339 IRAS 16219−2514 IRAS 16235−2416 IRAS 16240−2430 IRAS 16240−2430 IRAS 16242−2422 IRAS 16253−2429 IRAS 16262−2317 IRAS 16275−2251 IRAS 16285−2421 IRAS 16285−2358 IRAS 16301−2244 IRAS 16318−2402 IRAS 16375−2439 IRAS 16377−2426 IRAS 16381−2419 IRAS 16384−2334 IRAS 16403−2447 IRAS 16396−2419 IRAS 16411−2329 IRAS 16420−2327 aGroup indicating likelihood of being an embedded protostar (see §5.2). bAll source names are preceded by the prefix "SSTc2d ". cIntegrated luminosity using all available detections between 1.25 (2MASS J-band) and 70 µm. Entries in italics denote that the calculation only extends to 24 µm due to no flux information available at 70 µm. – 59 – Table 8. Candidate Low-Luminosity Embedded Objects in Serpens Source Number Groupa Source Nameb colhead(J2000) (J2000) Spitzer RA Dec 200 201 202 203 204 205 206 207 208 209 210 211 212 213 214 215 216 217 218 6 6 3 6 6 6 6 6 1 6 2 3 3 2 2 2 6 6 6 J182841.86−000321.2 J182844.02+005338.0 J182844.78+005125.9 J182844.95+005203.4 J182854.91+001832.8 J182902.11+003120.6 J182902.83+003009.7 J182902.95+003008.3 J182909.05+003127.8 J182923.18+013454.8 J182949.63+011522.0 J182951.98+011538.2 J182952.51+003611.9 J182953.04+003606.8 J182954.31+003601.4 J182957.67+011304.4 J183000.55+010304.0 J183004.03+003424.2 J183005.26+004104.6 18 28 41.86 18 28 44.02 18 28 44.78 18 28 44.95 18 28 54.91 18 29 02.11 18 29 02.83 18 29 02.95 18 29 09.05 18 29 23.18 18 29 49.63 18 29 51.98 18 29 52.51 18 29 53.04 18 29 54.31 18 29 57.67 18 30 00.55 18 30 04.03 18 30 05.26 −00 03 21.2 +00 53 38.0 +00 51 25.9 +00 52 03.4 +00 18 32.8 +00 31 20.6 +00 30 09.7 +00 30 08.3 +00 31 27.8 +01 34 54.8 +01 15 22.0 +01 15 38.2 +00 36 11.9 +00 36 06.8 +00 36 01.4 +01 13 04.4 +01 03 04.0 +00 34 24.2 +00 41 04.6 c LIR (L⊙) 0.153 0.146 0.016 0.209 0.033 0.032 0.134 0.035 0.009 0.040 0.395 0.028 0.198 0.169 0.119 0.166 0.025 0.151 0.056 DIRAS (′′) Nearest IRAS Source 240.0 99.1 40.9 14.2 321.8 70.9 52.0 51.3 70.9 891.6 11.3 31.2 18.8 10.0 10.0 64.9 157.8 144.5 92.5 IRAS 18261−0009 IRAS 18262+0050 IRAS 18262+0050 IRAS 18262+0050 IRAS 18267+0016 IRAS 18265+0028 IRAS 18265+0028 IRAS 18265+0028 IRAS 18265+0028 IRAS 18276+0124 IRAS 18273+0113 IRAS 18273+0113 IRAS 18273+0034 IRAS 18273+0034 IRAS 18273+0034 IRAS 18274+0112 IRAS 18273+0059 IRAS 18273+0030 IRAS 18275+0040 aGroup indicating likelihood of being an embedded protostar (see §5.2). bAll source names are preceded by the prefix "SSTc2d ". cIntegrated luminosity using all available detections between 1.25 (2MASS J-band) and 70 µm. Entries in italics denote that the calculation only extends to 24 µm due to no flux information available at 70 µm. Table 9. Groups Showing Likelihood Each Candidate is an Embedded Protostar Associated with Region of High Volume Densitya Volume Densitya Column Densitya Associated with Region of High Embedded in Region of High 4.5 µm Number Jets/Nebulositya of Candidatesb Yes Yes Yes Unknown Unknown Unknown Unknown No Yes No/Unknown No/Unknown No/Unknown No/Unknown No/Unknown No/Unknown N/A N/A N/A N/A Yes Yes No/Unknown No/Unknown N/A N/A Yes No/Unknown Yes No/Unknown Yes No/Unknown N/A 40 8 19 1 0 1 95 54 Group 1 2 3 4a 4b 4c 5 6 a"Yes" indicates data exists and satisfies this condition. "No" indicates data exists and shows that this condition is not satisfied. "Unknown" indicates no data exists to evaluate this condition. As discussed in the text, "No" and "Unknown" are considered equivalent for all but "Associated with Region of High Volume Density" since negative results do not prove a source is not an embedded protostar. bTotal number of the 218 candidates in this group. – 60 – Table 10. Candidates Divided Into Groups Core/Cloud IRAM04191+1522 IRAM04191+1522 IRAM04191+1522 L1521F TMC1 B35A DC303.8−14.2 CB68 CB130−3 L328 L673 L673 L673−7 L1014 L1221 L1251 L1251 L1251 L1251 Perseus Perseus Perseus Perseus Perseus Perseus Perseus Perseus Perseus Perseus Perseus Perseus Perseus Perseus Perseus Perseus Chamaeleon II Lupus III Ophiuchus Ophiuchus Serpens DC302.1+7.4 Perseus Perseus Perseus Serpens a LIR (L⊙) MIPS3b SIMBAD 0.023 0.033 0.250 0.015 0.383 0.141 0.387 0.299 0.054 0.065 0.201 0.037 0.017 0.087 0.404 0.156 0.077 0.274 0.081 0.061 0.432 0.347 0.036 0.015 0.185 0.105 0.018 0.033 0.299 0.135 0.180 0.098 0.144 0.310 0.253 0.361 0.134 0.107 0.046 0.009 0.232 0.305 0.183 0.149 0.395 N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A C/S C/S Y Y Y Y C/S C/S Y Y Y Y C/S Y C/S C/S Y Y C/S Y C/S N/A C/S C/S Y C/S NAME IRAM 04191−IRS IRAS F04191+1523 IRAS F04191+1523 NAME LDN 1521F−IRS IRAS F04381+2540 Barnard 35A BHR 86 [CB88] 68 [CB88] 130 [VRC2001] L328 SMM 1 [VRC2001] L673 SMM 1 [VRC2001] L673 SMM 2 NAME LDN 673 7 NAME LDN 1014−IRS [LH2005] MM 2 LDN 1251A [KP93] 3−6 LDN 1251B LDN 1251B [BC86b] LDN 1448 IRS 3B NAME LDN 1448* 2MASS J03273825+3013585 [JJK2007] 9 [JJK2007] 12 IRAS 03256+3055 [JCC87] IRAS 4B1 [JCC87] IRAS 4B SSTc2d J033032.7+302626 [WBK2005] SMM J033135+30455 IRAS 03292+3039 IRAS 03293+3052 [WBK2005] SMM J033328+31078 [JJK2007] 35 HH 211 NAME IC 348 MMS BHR 86 NAME LUPUS 3 MMS CRBR 2403.7−2948 [SSG2006] MMS126 [DAB2006] VLA−7 CG 19 [JCC87] IRAS 4B2 [JJK2007] 27 IRAS 03262+3123 [SB86] S68 1b Group Source Number 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 2 2 2 2 2 001 002 003 004 005 006 017 021 024 025 027 029 031 038 041 044 045 048 049 055 056 063 064 065 068 071 073 081 084 088 090 092 093 105 106 132c 170 182 185 208 016 074 075 078 210 – 61 – Table 10-Continued Core/Cloud Serpens Serpens Serpens DC2573−25 DC3272+18 B59 L673 L1148 L1251 Perseus Perseus Perseus Perseus Perseus Perseus Chamaeleon II Ophiuchus Ophiuchus Ophiuchus Serpens Serpens Serpens Lupus IV Lupus I BHR22 BHR16 CG30−31 DC2573−25 DC2573−25 Mu8 DC3391+117−2 DC3460+78 L673 Perseus Perseus Perseus Perseus Perseus Perseus Perseus Perseus Perseus Perseus Perseus Perseus a LIR (L⊙) MIPS3b SIMBAD 0.169 0.119 0.166 0.193 0.034 0.343 0.138 0.081 0.272 0.462 0.166 0.172 0.101 0.021 0.200 0.184 0.103 0.440 0.374 0.016 0.028 0.198 0.049 0.011 0.006 0.081 0.064 0.013 0.018 0.001 0.005 0.003 0.168 0.010 0.017 0.008 0.015 0.008 0.016 0.018 0.008 0.008 0.010 0.010 0.014 C/S Y C/S N/A N/A N/A N/A N/A N/A Y Y C/S C/S C/S C/S Y C/S C/S C/S N C/S C/S Y Y N/A N/A N/A N/A N/A N/A N/A N/A N/A N N N N N N N N N U N N IRAS 18273+0034 IRAS 18273+0034 GCNM 104 NAME DCLD 257.3−2.5 1 [LM99] 135 [BHB2007] 10 [VRC2001] L673 SMM 2 IRAS F20404+6712 IRAS 22290+7458 [JJK2007] 4 IRAS 03267+3218 [JJK2007] 42 [JJK2007] 47 Cl* IC 348 LRL 54460 Cl* IC 348 LRL 1889 DENIS−P J125338.9−771553 BBRCG 7 NAME WL 20W BBRCG 36 None NAME Serpens SMM 10 IR IRAS 18273+0034 2MASS J16011549-4152351 IRAS 15376−3411 None None None None Dcld 257.3−02.5 None None None None None None None None None None None None None None None None Group Source Number 2 2 2 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 4a 4c 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 213 214 215 011 018 023 030 032 043 060 080 104 107 108 109 124 181 183 184 202 211 212 161 145 008 009 010 012 013 015 019 020 028 050 051 052 053 054 058 059 061 062 067 079 087 – 62 – Table 10-Continued Core/Cloud Perseus Perseus Perseus Perseus Perseus Perseus Perseus Perseus Perseus Chamaeleon II Chamaeleon II Chamaeleon II Chamaeleon II Chamaeleon II Chamaeleon II Chamaeleon II Chamaeleon II Chamaeleon II Chamaeleon II Chamaeleon II Chamaeleon II Chamaeleon II Lupus I Lupus I Lupus I Lupus I Lupus I Lupus I Lupus I Lupus I Lupus I Lupus I Lupus I Lupus I Lupus I Lupus I Lupus I Lupus I Lupus I Lupus I Lupus I Lupus I Lupus I Lupus I Lupus IV a LIR (L⊙) MIPS3b SIMBAD 0.007 0.050 0.057 0.008 0.013 0.019 0.010 0.012 0.013 0.004 0.012 0.006 0.010 0.006 0.050 0.003 0.020 0.005 0.003 0.003 0.008 0.018 0.004 0.004 0.004 0.003 0.004 0.003 0.003 0.005 0.006 0.001 0.004 0.005 0.008 0.003 0.007 0.003 0.008 0.006 0.006 0.006 0.007 0.004 0.006 N N U U N U N N N U U U N N N N N N N N N N N N N N N N N N U N N U U N N N U U U N N N N None None None None None None None None None None None None None None IRAS 12533−7632 None None None None None None None None None None None None None None None None None None None None None None None None None None None None None None Group Source Number 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 089 094 095 101 102 103 113 115 118 120 121 122 123 125 126 128 129 130 131 133 134 135 136 137 138 139 140 141 142 143 144 146 147 148 149 150 151 152 153 154 155 156 157 158 159 – 63 – Table 10-Continued Core/Cloud Lupus IV Lupus IV Lupus III Lupus III Lupus III Lupus III Lupus III Lupus III Lupus III Lupus III Lupus III Lupus III Lupus III Lupus III Ophiuchus Ophiuchus Ophiuchus Ophiuchus Ophiuchus Ophiuchus Ophiuchus Ophiuchus Ophiuchus Ophiuchus Ophiuchus Ophiuchus Ophiuchus Ophiuchus Ophiuchus B35A DC2742−04 B59 L723 L1148 L1155E L1155E L1228 L1228 L1165 L1165 L1251 L1251 L1251 Perseus Perseus a LIR (L⊙) MIPS3b SIMBAD 0.003 0.006 0.008 0.015 0.009 0.009 0.018 0.012 0.004 0.014 0.021 0.009 0.008 0.006 0.004 0.033 0.012 0.000 0.003 0.034 0.006 0.004 0.007 0.005 0.005 0.002 0.005 0.005 0.016 0.006 0.005 0.024 0.048 0.003 0.004 0.004 0.199 0.038 0.030 0.008 0.066 0.030 0.004 0.006 0.414 U N N N N N N N N N N N N N N Y N N N N N N U N Y N N N Y N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N Y None None None None None None None 2MASS J16075475−3915446 None None None None None None None [L89] R7 IRAS Z16219−2514 None None IRAS 16301−2244 None None None None None None None None None None None 2MASS J17111182−2726547 None None None None 2MASX J20570656+7736557 None None None None None None None HBC 341 Group Source Number 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 160 162 163 164 165 166 167 168 171 172 174 175 176 177 178 179 180 186 187 190 191 192 193 194 195 196 197 198 199 007 014 022 026 033 034 035 036 037 039 040 042 046 047 057 066 – 64 – Table 10-Continued Core/Cloud Perseus Perseus Perseus Perseus Perseus Perseus Perseus Perseus Perseus Perseus Perseus Perseus Perseus Perseus Perseus Perseus Perseus Perseus Perseus Perseus Perseus Perseus Chamaeleon II Lupus III Lupus III Ophiuchus Ophiuchus Serpens Serpens Serpens Serpens Serpens Serpens Serpens Serpens Serpens Serpens Serpens a LIR (L⊙) MIPS3b SIMBAD 0.119 0.014 0.019 0.002 0.020 0.009 0.004 0.015 0.002 0.012 0.067 0.026 0.029 0.019 0.292 0.026 0.051 0.011 0.018 0.015 0.006 0.010 0.003 0.035 0.009 0.013 0.036 0.153 0.146 0.209 0.033 0.032 0.134 0.035 0.040 0.025 0.151 0.056 C/S N N C/S C/S N N N N N N N N N Y N N N N N N N N C/S N N Y N N Y Y C/S C/S C/S N C/S N N 2MASS J03285630+3122279 None None None HH5 None None None None None IRAS 03328+3035 None None None None None Cl* IC 348 LRL 904 None None None None None None V* V1192 Sco None None None None None IRAS 18262+0050 None SSTc2d J192902.1+003120 2MASS J18290283+0030092 None None None None None Group Source Number 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 6 6 6 6 6 6 6 6 069 070 072 076 077 082 083 085 086 091 096 097 098 099 100 110 111 112 114 116 117 119 127 169 173 188 189 200 201 203 204 205 206 207 209 216 217 218 aIntegrated luminosity using all available detections between 1.25 and 70 µm. Italics denote that the calculation only extends to 24 µm due to no flux information available at 70 µm. bIndicates whether each candidate is associated with a 160 µm source. Y: Yes; N: No; U: Unknown - located off the edge of the 160 µm map; C/S: Unknown - located in a region that is confused and/or saturated at 160 µm; N/A: Not applicable for the dense cores. cSource 132 is the same as Source 017. – 65 – Source Number 001 003 004 005 Table 11. SEDs of Groups 1-3 Wavelength Flux Density σ Aperturea (µm) (mJy) (mJy) (′′) Referenceb Evolutionary Indicators 3.6 4.5 5.8 8.0 24 60 70 90 160 200 350 450 850 1300 1.25 1.65 2.17 3.6 4.5 5.8 8.0 24 25 60 70 100 450 850 1300 3.6 4.5 5.8 8.0 24 70 350 450 850 1200 1.25 1.65 2.17 4.5 8.0 12 0.22 0.78 0.94 0.63 16.4 500 875 800 6000 10000 5000 10000 2500 650 0.3 1.7 8.3 48.5 93.0 135 179 985 1300 5910 4320 14400 1400 600 400 0.42 0.24 0.35 0.62 24.0 509 8900 7000 1500 600 0.22 2.27 16.1 139 346 396 0.02 0.06 0.08 0.06 1.73 100 120 160 1200 2000 800 2000 500 65 0.03 0.2 0.9 3.42 6.60 9.36 12.5 105 117 591 591 1730 800 100 ... 0.06 0.03 0.05 0.06 2.52 86.5 1400 1800 500 150 0.05 0.17 0.88 10.3 24.7 32 ... ... ... ... ... 60 ... 60 60 60 40 60 60 60 ... ... ... ... ... ... ... ... ... ... ... ... 40 40 60 ... ... ... ... ... ... 40 40 40 40 ... ... ... ... ... ... 1 1 1 1 1 2 1 2 2 2 3 2 2 2 4 4 4 1 1 1 1 1 5 5 1 5 6 6 7 8 1 1 1 1 1 9 10 6 11 4 4 4 1 1 5 LIR L70 int = 0.04 L⊙ int = 0.05 L⊙ Lbol = 0.12 ± 0.02 L⊙ Lbol/Lsmm = 5 ± 1 Tbol = 27 ± 3 K LIR L70 int = 0.42 L⊙ int = 0.23 L⊙ Lbol = 0.50 ± 0.06 L⊙ Lbol/Lsmm = 197 ± 23 Tbol = 149 ± 8 K LIR L70 int = 0.03 L⊙ int = 0.03 L⊙ Lbol = 0.13 ± 0.02 L⊙ Lbol/Lsmm = 7 ± 1 Tbol = 20 ± 3 K LIR L70 int = 0.64 L⊙ int = 0.36 L⊙ Lbol = 0.73 ± 0.07 L⊙ Lbol/Lsmm = 89 ± 9 Tbol = 165 ± 10 K – 66 – Table 11-Continued Source Number 017 021 024 025 Wavelength Flux Density σ Aperturea (µm) (mJy) (mJy) 24 25 60 70 100 450 850 1200 3.6 4.5 5.8 8.0 24 25 60 70 100 200 1200 1300 3.6 4.5 5.8 8.0 24 25 60 70 100 350 450 850 3.6 4.5 5.8 8.0 24 70 350 450 850 3.6 4.5 5.8 8.0 1720 2520 9810 6740 12600 6410 1350 353 0.98 12.8 40.1 82.7 995 1050 4100 5450 8100 29000 1260 225 0.36 1.38 1.90 1.12 659 1480 19500 13100 33700 19100 27800 9400 1.44 4.23 5.85 6.76 54.1 312 2800 1100 1480 1.55 3.79 3.34 1.75 182 151 490 919 1260 2800 110 71 0.07 0.89 2.76 5.75 105 83.9 200 574 1100 3400 140 49.0 0.03 0.12 0.15 0.09 69.3 133 1950 1780 8080 2900 14900 2300 0.11 0.30 0.41 0.47 5.68 35.0 700 400 296 0.14 0.30 0.28 0.27 (′′) ... ... ... ... ... 120 120 40 ... ... ... ... ... ... 43.5 ... 43.5 89.4 120 ... ... ... ... ... ... ... ... ... ... 40 40 120 ... ... ... ... ... ... 40 40 120 ... ... ... ... Referenceb Evolutionary Indicators 1 5 5 1 5 12 12 11 1 1 1 1 1 5 13 1 13 13 14 15 1 1 1 1 1 5 5 1 5 3 6 6 1 1 1 1 1 1 3 3 16 1 1 1 1 LIR L70 int = 0.65 L⊙ int = 0.46 L⊙ Lbol = 0.91 ± 0.09 L⊙ Lbol/Lsmm = 2502 ± 257 Tbol = 69 ± 5 K LIR L70 int = 0.50 L⊙ int = 0.54 L⊙ Lbol = 0.94 ± 0.16 L⊙ Lbol/Lsmm = 16 ± 3 Tbol = 52 ± 4 K LIR L70 int = 0.09 L⊙ int = 0.07 L⊙ Lbol = 0.20 ± 0.04 L⊙ Lbol/Lsmm = 9 ± 2 Tbol = 55 ± 10 K LIR L70 int = 0.11 L⊙ int = 0.07 L⊙ Lbol = 0.18 ± 0.04 L⊙ Lbol/Lsmm = 9 ± 2 – 67 – Table 11-Continued Source Number Wavelength Flux Density σ Aperturea (µm) (mJy) (mJy) (′′) Referenceb Evolutionary Indicators 24 70 350 4.5 24 70 450 1200 3.6 4.5 5.8 8.0 24 70 350 450 850 1200 3.6 4.5 5.8 8.0 24 70 350 450 850 3.6 4.5 5.8 8.0 24 70 350 1200 3.6 4.5 5.8 8.0 24 70 350 1200 1.25 1.65 87.0 337 2000 0.11 1.94 148 1100 358 4.02 12.2 21.2 26.2 89.5 503 3200 21500 1800 630 0.57 3.10 4.96 3.84 47.5 5080 6400 11300 7900 0.13 0.42 0.32 0.16 4.97 1400 9400 1010 0.43 1.07 0.75 0.37 1.90 688 7500 906 0.24 7.99 9.14 109 500 0.01 0.28 23.8 400 72 0.34 0.91 1.53 1.85 9.39 70.4 500 16100 400 126 0.04 0.24 0.25 0.35 5.00 691 1000 6100 1900 0.02 0.05 0.04 0.03 0.54 192 1400 202 0.04 0.10 0.06 0.04 0.23 94.5 1100 181 0.05 0.32 ... ... 20 ... ... ... 40 80 ... ... ... ... ... ... 40 120 120 120 ... ... ... ... ... ... 40 40 120 ... ... ... ... ... ... 40 80 ... ... ... ... ... ... 40 80 ... ... 1 1 3 1 1 1 3 11 1 1 1 1 1 1 3 17 17 17 1 1 1 1 1 1 3 6 6 1 1 1 1 1 1 3 11 1 1 1 1 1 1 3 11 4 4 Tbol = 62 ± 9 K LIR L70 int = 0.03 L⊙ int = 0.04 L⊙ Lbol = 0.09 ± 0.03 L⊙ Lbol/Lsmm = 10 ± 3 Tbol = 24 ± 6 K LIR int = 0.15 L⊙ L70 int = 0.09 L⊙ Lbol = 0.34 ± 0.11 L⊙ Lbol/Lsmm = 3 ± 1 Tbol = 66 ± 21 K LIR L70 int = 0.67 L⊙ int = 0.81 L⊙ Lbol = 0.93 ± 0.16 L⊙ Lbol/Lsmm = 8 ± 1 Tbol = 42 ± 4 K LIR L70 int = 0.26 L⊙ int = 0.34 L⊙ Lbol = 0.78 ± 0.11 L⊙ Lbol/Lsmm = 8 ± 1 Tbol = 24 ± 3 K LIR L70 int = 0.13 L⊙ int = 0.17 L⊙ Lbol = 0.55 ± 0.08 L⊙ Lbol/Lsmm = 7 ± 1 Tbol = 21 ± 2 K LIR L70 int = 0.58 L⊙ int = 0.20 L⊙ 031 038 041 044 045 063 – 68 – Table 11-Continued Source Number Wavelength Flux Density σ Aperturea (µm) (mJy) (mJy) (′′) Referenceb Evolutionary Indicators 2.17 3.6 4.5 5.8 8.0 24 70 160 350 450 850 850 1100 1.25 1.65 2.17 3.6 4.5 5.8 8.0 24 70 160 350 850 1100 3.6 4.5 5.8 8.0 24 70 160 350 850 850 1100 1.25 1.65 2.17 3.6 4.5 5.8 8.0 12 36.4 62.5 72.0 79.1 93.1 343 1120 16200 9800 1600 720 260 270 0.17 0.69 1.07 1.01 2.28 2.57 4.01 55.3 179 3340 3100 240 210 0.13 0.27 0.43 0.94 24.4 94.0 4930 1900 750 830 290 0.06 0.46 0.72 1.22 2.73 2.67 2.69 20 0.84 4.43 5.13 5.46 6.57 36.2 160 4600 1500 ... ... ... 30 0.01 0.03 0.11 0.09 0.17 0.18 0.28 5.83 26.5 996 500 ... 70 0.01 0.02 0.04 0.07 2.58 14.5 1420 300 ... ... 20 0.01 0.02 0.10 0.09 0.19 0.19 0.19 ... ... ... ... ... ... ... ... ... 40 T T T T ... ... ... ... ... ... ... ... ... ... 40 T 120 ... ... ... ... ... ... ... 40 T T T ... ... ... ... ... ... ... ... Lbol = 1.1 ± 0.2 L⊙ Lbol/Lsmm = 36 ± 7 Tbol = 199 ± 33 K LIR L70 int = 0.06 L⊙ int = 0.03 L⊙ Lbol = 0.20 ± 0.05 L⊙ Lbol/Lsmm = 12 ± 3 Tbol = 65 ± 12 K LIR L70 int = 0.03 L⊙ int = 0.02 L⊙ Lbol = 0.22 ± 0.06 L⊙ Lbol/Lsmm = 15 ± 4 Tbol = 29 ± 3 K LIR L70 int = 0.31 L⊙ int = 0.26 L⊙ Lbol = 0.45 ± 0.09 L⊙ Lbol/Lsmm = 193 ± 38 Tbol = 64 ± 8 K 4 1 1 1 1 1 1 1 3 18, 19 18, 19 20 21 4 4 4 1 1 1 1 1 1 1 9 18, 19 21 1 1 1 1 1 1 1 9 18, 19 20 21 4 4 4 1 1 1 1 22, 23 064 065 068 – 69 – Table 11-Continued Source Number Wavelength Flux Density σ Aperturea (µm) (mJy) (mJy) (′′) Referenceb Evolutionary Indicators 24 25 60 70 100 160 450 850 4.5 5.8 24 70 450 850 850 1100 3.6 4.5 5.8 8.0 24 70 160 350 1100 3.6 4.5 5.8 8.0 24 60 70 100 160 350 450 850 850 1100 3.6 4.5 8.0 24 60 70 213 150 1430 1520 6000 4960 552 102 0.09 0.11 15.4 1320 2100 720 130 370 0.06 0.20 0.19 0.16 30.5 310 2560 3200 800 0.13 0.85 0.83 0.62 14.7 2332 3880 13900 16400 12300 3300 2260 910 760 0.39 0.42 1.29 13.0 1590 1720 22.6 ... 129 214 ... 1620 140 12 0.01 0.03 1.63 183 ... ... ... 50 0.01 0.01 0.04 0.03 3.21 43.1 865 500 70 0.01 0.07 0.07 0.06 1.55 210 533 1520 5680 2100 ... ... ... 30 0.08 0.14 0.11 1.38 127 183 ... ... ... ... ... ... T T ... ... ... ... T T T 80 ... ... ... ... ... ... ... 40 120 ... ... ... ... ... ... ... ... ... 40 T T T T ... ... ... ... ... ... 1 22, 23 5 1 22, 23 1 23 23 1 1 1 1 18, 19 18, 19 20 21 1 1 1 1 1 1 1 9 21 1 1 1 1 1 5 1 5 1 3 18, 19 18, 19 20 21 1 1 1 1 5 1 LIR L70 int = 0.18 L⊙ int = 0.23 L⊙ Lbol = 0.24 ± 0.09 L⊙ Lbol/Lsmm = 24 ± 4 Tbol = 41 ± 4 K LIR L70 int = 0.06 L⊙ int = 0.06 L⊙ Lbol = 0.18 ± 0.04 L⊙ Lbol/Lsmm = 7 ± 2 Tbol = 33 ± 4 K LIR L70 int = 0.50 L⊙ int = 0.63 L⊙ Lbol = 1.1 ± 0.2 L⊙ Lbol/Lsmm = 22 ± 4 Tbol = 35 ± 3 K LIR L70 int = 0.23 L⊙ int = 0.29 L⊙ Lbol = 1.2 ± 0.3 L⊙ Lbol/Lsmm = 7 ± 2 Tbol = 28 ± 2 K 071 081 084 088 – 70 – Table 11-Continued Source Number Wavelength Flux Density σ Aperturea (µm) (mJy) (mJy) (′′) Referenceb Evolutionary Indicators 090 092 093 105 100 160 350 450 850 850 1100 1.25 1.65 2.17 3.6 4.5 5.8 8.0 24 60 70 160 350 1100 3.6 4.5 5.8 8.0 24 70 350 1100 4.5 5.8 24 70 160 350 450 850 850 1100 3.6 5.8 24 70 350 450 850 9087 24200 17500 21500 5600 2130 2310 0.71 2.69 5.81 11.1 17.2 23.3 32.8 181 1350 1130 8250 2900 290 0.16 0.94 2.05 4.21 111 809 3700 1840 0.26 0.44 19.5 1820 13600 13300 50400 5370 4960 1840 0.06 0.28 2.20 4080 25200 83300 8190 1270 8380 2800 ... ... ... 60 0.06 0.16 0.21 0.86 1.25 1.62 2.28 19.0 122 157 2380 400 50 0.01 0.07 0.15 0.30 11.8 116 600 40 0.04 0.05 2.08 256 3930 500 ... ... ... 40 0.02 0.06 0.34 565 4100 ... ... ... ... 40 T T T 120 ... ... ... ... ... ... ... ... ... ... ... 40 80 ... ... ... ... ... ... 20 T ... ... ... ... ... 40 T T T T ... ... ... ... 40 T T 1 1 3 18, 19 18, 19 20 21 4 4 4 1 1 1 1 1 5 1 1 3 21 1 1 1 1 1 1 9 21 1 1 1 1 1 9 18, 19 18, 19 20 21 1 1 1 1 3 18, 19 18, 19 LIR L70 int = 0.30 L⊙ int = 0.20 L⊙ Lbol = 0.57 ± 0.11 L⊙ Lbol/Lsmm = 30 ± 6 Tbol = 114 ± 17 K LIR L70 int = 0.16 L⊙ int = 0.14 L⊙ Lbol = 0.29 ± 0.04 L⊙ Lbol/Lsmm = 8 ± 1 Tbol = 47 ± 5 K LIR L70 int = 0.24 L⊙ int = 0.31 L⊙ Lbol = 1.1 ± 0.2 L⊙ Lbol/Lsmm = 4 ± 1 Tbol = 24 ± 2 K LIR L70 int = 0.52 L⊙ int = 0.66 L⊙ Lbol = 1.8 ± 0.3 L⊙ Lbol/Lsmm = 4 ± 1 Tbol = 21 ± 3 K – 71 – Table 11-Continued Source Number Wavelength Flux Density σ Aperturea (µm) (mJy) (mJy) (′′) Referenceb Evolutionary Indicators 850 1100 3.6 4.5 5.8 8.0 24 70 450 850 850 1100 3.6 4.5 5.8 8.0 24 60 70 160 1200 1.25 1.65 2.17 3.6 4.5 5.8 8.0 24 70 850 1200 3.6 4.5 5.8 8.0 24 60 70 100 160 350 1100 1200 3.6 3140 2070 0.05 0.20 0.52 0.60 10.8 3280 17600 3590 2270 1420 0.26 1.00 0.99 0.55 32.4 1670 2610 8710 990 0.06 0.28 1.11 6.11 10.7 16.6 29.6 459 3490 440 433 0.59 2.55 4.49 3.13 90.1 2911 2000 16300 18700 45400 1140 411 0.13 ... 50 0.01 0.06 0.06 0.12 1.17 454 ... ... ... 30 0.03 0.08 0.09 0.06 3.44 217 363 2510 130 0.01 0.01 0.09 0.46 0.80 0.82 1.17 48.3 526 ... ... 0.05 0.20 0.37 0.22 9.47 378 281 2780 5560 6800 100 ... 0.01 T 80 ... ... ... ... ... ... T T T T ... ... ... ... ... ... ... ... 56 ... ... ... ... ... ... ... ... ... T T ... ... ... ... ... ... ... ... ... 40 120 T ... 20 21 1 1 1 1 1 1 18, 19 18, 19 20 21 1 1 1 1 1 5 1 1 24 4 4 4 1 1 1 1 1 1 25 26 1 1 1 1 1 5 1 5 1 9 27 26 1 LIR L70 int = 0.42 L⊙ int = 0.54 L⊙ Lbol = 1.1 ± 0.2 L⊙ Lbol/Lsmm = 15 ± 3 Tbol = 25 ± 4 K LIR L70 int = 0.22 L⊙ int = 0.29 L⊙ Lbol = 0.37 ± 0.08 L⊙ Lbol/Lsmm = 275 ± 62 Tbol = 39 ± 4 K LIR L70 int = 0.18 L⊙ int = 0.15 L⊙ Lbol = 0.15 ± 0.02 L⊙ Lbol/Lsmm = 335 ± 46 Tbol = 105 ± 2 K LIR L70 int = 0.08 L⊙ int = 0.09 L⊙ Lbol = 0.45 ± 0.08 L⊙ Lbol/Lsmm = 7 ± 1 Tbol = 30 ± 2 K LIR int = 0.39 L⊙ 106 170 182 185 016 – 72 – Table 11-Continued Source Number Wavelength Flux Density σ Aperturea (µm) (mJy) (mJy) (′′) Referenceb Evolutionary Indicators 4.5 5.8 8.0 24 70 1200 3.6 5.8 24 70 450 850 3.6 4.5 5.8 8.0 24 70 450 850 850 1100 3.6 4.5 5.8 8.0 24 60 70 160 850 850 1100 1.25 1.65 2.17 3.6 4.5 5.8 8.0 24 70 1200 3.6 4.5 1.10 2.04 2.39 87.7 1750 370 0.03 0.25 28.4 3910 20900 2920 0.42 1.22 1.24 0.82 111 1950 2500 1420 470 550 0.76 2.26 2.47 3.03 191 1070 1150 4380 1560 390 510 0.75 1.87 2.42 2.17 2.12 1.81 2.43 56.4 492 100 0.47 2.03 0.08 0.15 0.16 9.20 240 50 0.01 0.04 3.00 538 ... ... 0.04 0.12 0.10 0.11 11.7 292 ... ... ... 50 0.08 0.17 0.24 0.21 20.1 96.3 159 1270 ... ... 40 0.07 0.14 0.14 0.15 0.15 0.13 0.17 5.92 69.5 20 0.04 0.16 ... ... ... ... ... 120 ... ... ... ... T T ... ... ... ... ... ... T T T 80 ... ... ... ... ... ... ... ... T T 80 ... ... ... ... ... ... ... ... ... 40 ... ... 1 1 1 1 1 14 1 1 1 1 18, 19 18, 19 1 1 1 1 1 1 18, 19 18, 19 20 21 1 1 1 1 1 5 1 1 18, 19 20 21 4 4 4 1 1 1 1 1 1 14 1 1 L70 int = 0.42 L⊙ Lbol = 0.37 ± 0.05 L⊙ Lbol/Lsmm = 321 ± 43 Tbol = 66 ± 1 K LIR L70 int = 0.51 L⊙ int = 0.63 L⊙ Lbol = 1.3 ± 0.2 L⊙ Lbol/Lsmm = 16 ± 3 Tbol = 26 ± 4 K LIR L70 int = 0.31 L⊙ int = 0.33 L⊙ Lbol = 0.37 ± 0.06 L⊙ Lbol/Lsmm = 25 ± 4 Tbol = 53 ± 4 K LIR L70 int = 0.25 L⊙ int = 0.20 L⊙ Lbol = 0.38 ± 0.07 L⊙ Lbol/Lsmm = 134 ± 26 Tbol = 59 ± 7 K LIR L70 int = 0.32 L⊙ int = 0.26 L⊙ Lbol = 0.28 ± 0.04 L⊙ Lbol/Lsmm = 416 ± 54 Tbol = 206 ± 5 K LIR L70 int = 0.06 L⊙ int = 0.04 L⊙ 074 075 078 011 018 – 73 – Table 11-Continued Source Number Wavelength Flux Density σ Aperturea (µm) (mJy) (mJy) (′′) Referenceb Evolutionary Indicators 023 030 032 043 060 5.8 8.0 24 70 1200 3.6 4.5 5.8 8.0 24 70 3.6 4.5 5.8 24 70 1.25 1.65 2.17 3.6 4.5 5.8 8.0 24 25 60 70 1200 1.25 1.65 2.17 3.6 4.5 5.8 8.0 24 25 60 70 1200 1.25 1.65 2.17 3.6 4.5 4.16 4.94 42.9 219 290 3.18 11.9 22.1 42.2 1870 10200 0.11 0.48 0.51 7.61 1230 0.09 0.59 0.87 1.21 1.39 0.88 1.25 70.5 126 321 306 170 0.33 1.76 6.43 15.8 21.8 25.2 27.3 272 375 734 848 705 0.18 0.50 1.19 3.62 9.27 0.30 0.43 4.50 32.9 40 0.23 0.84 1.53 2.91 199 1390 0.03 0.05 0.07 0.85 170 0.01 0.03 0.08 0.09 0.10 0.07 0.09 7.41 15.1 48.2 41.8 35 0.02 0.13 0.23 1.22 1.64 1.82 1.93 28.5 48.8 58.7 117 141 0.01 0.03 0.08 0.29 0.65 ... ... ... ... 80 ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... 80 ... ... ... ... ... ... ... ... ... ... ... 80 ... ... ... ... ... 1 1 1 1 14 1 1 1 1 1 1 1 1 1 1 1 4 4 4 1 1 1 1 1 5, 28 5, 28 1 11 4 4 4 1 1 1 1 1 5 5 1 11 4 4 4 1 1 Lbol = 0.06 ± 0.01 L⊙ Lbol/Lsmm = 90 ± 12 Tbol = 105 ± 3 K LIR int = 0.57 L⊙ L70 int = 0.42 L⊙ Lbol ≥ 0.34 L⊙ Lbol/Lsmm = ... Tbol ≤ 106 K LIR int = 0.23 L⊙ L70 int = 0.30 L⊙ Lbol ≥ 0.14 L⊙ Lbol/Lsmm = ... Tbol ≤ 60 K int = 0.14 L⊙ int = 0.09 L⊙ Lbol = 0.13 ± 0.02 L⊙ Lbol/Lsmm = 207 ± 28 Tbol = 145 ± 1 K LIR L70 LIR L70 int = 0.45 L⊙ int = 0.21 L⊙ Lbol = 0.38 ± 0.04 L⊙ Lbol/Lsmm = 173 ± 19 Tbol = 236 ± 3 K LIR L70 int = 0.77 L⊙ int = 0.69 L⊙ Lbol = 1.1 ± 0.2 L⊙ Lbol/Lsmm = 27 ± 4 Tbol = 64 ± 6 K – 74 – Table 11-Continued Source Number Wavelength Flux Density σ Aperturea (µm) (mJy) (mJy) (′′) Referenceb Evolutionary Indicators 5.8 8.0 24 25 60 70 100 160 450 850 850 1100 4.5 5.8 8.0 24 60 70 160 450 850 850 1100 3.6 4.5 5.8 8.0 24 70 850 1100 3.6 4.5 5.8 8.0 24 70 850 1100 1.25 1.65 2.17 3.6 4.5 5.8 11.1 12.0 396 545 4180 4300 6970 8970 12100 860 170 400 0.02 0.10 0.37 48.9 1880 1990 5960 11900 4900 410 520 6.81 3.99 12.8 2.15 45.4 1960 1460 1410 0.73 1.93 3.82 6.41 107 847 940 500 0.33 1.17 6.63 31.0 50.4 61.7 0.77 0.82 41.7 76.3 376 591 697 2550 ... ... ... 60 0.005 0.03 0.04 5.15 151 276 1730 6400 1200 ... 40 0.59 0.63 0.93 0.18 4.82 276 ... 40 0.05 0.15 0.27 0.45 11.7 124 ... 20 0.02 0.09 0.18 2.14 3.45 4.21 ... ... ... ... ... ... ... ... T T T 120 ... ... ... ... ... ... ... 40 120 T 80 ... ... ... ... ... ... T 80 ... ... ... ... ... ... T T ... ... ... ... ... ... 1 1 1 5 5 1 5 1 18, 19 18, 19 20 21 1 1 1 1 5 1 1 6 6 20 21 1 1 1 1 1 1 20 21 1 1 1 1 1 1 20 21 4 4 4 1 1 1 LIR L70 int = 0.28 L⊙ int = 0.34 L⊙ Lbol = 0.64 ± 0.18 L⊙ Lbol/Lsmm = 11 ± 3 Tbol = 34 ± 7 K LIR L70 int = 0.29 L⊙ int = 0.33 L⊙ Lbol = 0.32 ± 0.05 L⊙ Lbol/Lsmm = 56 ± 8 Tbol = 63 ± 3 K LIR L70 int = 0.17 L⊙ int = 0.15 L⊙ Lbol = 0.18 ± 0.03 L⊙ Lbol/Lsmm = 77 ± 11 Tbol = 76 ± 4 K LIR L70 int = 0.33 L⊙ int = 0.11 L⊙ Lbol = 0.26 ± 0.03 L⊙ Lbol/Lsmm = 195 ± 20 Tbol = 345 ± 4 K 080 104 107 109 – 75 – Table 11-Continued Source Number Wavelength Flux Density σ Aperturea (µm) (mJy) (mJy) (′′) Referenceb Evolutionary Indicators 124 181 183 184 8.0 24 70 1100 1.25 1.65 2.17 3.6 4.5 5.8 8.0 24 25 60 70 160 1300 1.25 1.65 2.17 3.6 4.5 5.8 8.0 24 70 1300 1.25 1.65 2.17 3.6 4.5 5.8 8.0 24 70 850 1.25 1.65 2.17 3.6 4.5 5.8 8.0 24 70.9 228 600 560 0.87 3.56 9.19 17.9 25.4 33.5 51.2 612 1010 2720 1540 3420 59.7 0.09 2.74 21.8 79.2 111 128 148 446 1100 50 4.41 32.0 97.3 127 143 140 99.0 816 15700 80 0.21 1.75 15.5 127 206 286 268 780 4.86 24.1 85.8 70 0.06 0.14 0.22 1.25 1.75 2.31 3.53 64.3 90.9 407 215 995 15.0 0.01 0.11 0.54 5.61 7.75 9.10 10.3 47.3 219 ... 0.23 2.09 3.40 9.36 11.2 11.4 12.2 287 2160 ... 0.01 0.16 0.57 8.96 14.9 20.5 19.9 84.1 ... ... ... 120 ... ... ... ... ... ... ... ... ... ... ... ... T ... ... ... ... ... ... ... ... ... 12 ... ... ... ... ... ... ... ... ... T ... ... ... ... ... ... ... ... 1 1 1 21 4 4 4 1 1 1 1 1 5 5 1 1 15 4 4 4 1 1 1 1 1 1 29 4 4 4 1 1 1 1 1 1 25 4 4 4 1 1 1 1 1 LIR L70 int = 0.31 L⊙ int = 0.14 L⊙ Lbol = 0.34 ± 0.05 L⊙ Lbol/Lsmm = 6000 ± 900 Tbol = 167 ± 6 K LIR L70 int = 0.17 L⊙ int = 0.05 L⊙ Lbol = 0.12 ± 0.01 L⊙ Lbol/Lsmm = 5000 ± 600 Tbol = 430 ± 19 K LIR L70 int = 0.73 L⊙ int = 0.64 L⊙ Lbol = 0.62 ± 0.09 L⊙ Lbol/Lsmm = 10000 ± 1500 Tbol = 225 ± 13 K LIR L70 int = 0.62 L⊙ int = 0.50 L⊙ Lbol = 0.76 ± 0.12 L⊙ Lbol/Lsmm = 3750 ± 600 Tbol = 157 ± 7 K – 76 – Table 11-Continued Source Number Wavelength Flux Density σ Aperturea (µm) (mJy) (mJy) (′′) Referenceb Evolutionary Indicators 161 145 25 70 100 1200 1.25 1.65 2.17 3.6 4.5 5.8 8.0 24 70 160 5.8 8.0 24 60 70 160 6560 12200 22000 375 0.42 1.67 4.83 8.36 9.92 8.98 7.70 75.9 1220 5390 0.28 0.70 8.78 468 372 705 1120 1720 4840 ... 0.06 0.10 0.20 0.61 0.70 0.63 0.53 7.99 173 1560 0.09 0.19 1.00 42.1 54.9 225 ... ... ... T ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... 5 1 5 26 4 4 4 1 1 1 1 1 1 1 1 1 1 5 1 1 LIR int = 0.08 L⊙ L70 int = 0.08 L⊙ Lbol ≥ 0.11 L⊙ Lbol/Lsmm = ... Tbol ≤ 126 K LIR int = 0.02 L⊙ L70 int = 0.03 L⊙ Lbol ≥ 0.02 L⊙ Lbol/Lsmm = ... Tbol ≤ 57 K aDiameter of the aperture used for photometry at λ > 100 µm, where the sources are likely to be extended, when specified. A "T" indicates the photometry is an attempt to derive the total source flux density rather than the flux density in a fixed aperture; see individual references for descriptions of the methods used to derive this quantity. bReferences.− (1) Spitzer observations from this work; (2) Andr´e et al. (1999); (3) Wu et al. (2007); (4) 2MASS; (5) IRAS ; (6) Young et al. (2006a); (7) Motte & Andr´e (2001); (8) Bourke et al. (2006); (9) M. Dunham et al. (2008), in preparation; (10) Kirk et al. (2005); (11) Kauffmann et al. (2008); (12) Young et al. (2003); (13) Lehtinen et al. (2005); (14) K. Brede et al. (2008), in preparation; (15) Henning et al. (1993); (16) H.J. Kim et al. (2008, in preparation); (17) Young et al. (2004); (18) Hatchell et al. (2005); (19) Hatchell et al. (2007a); (20) Kirk et al. (2006); (21) Enoch et al. (2006); (22) Clark (1991); (23) Hodapp et al. (2005); (24) Tachihara et al. (2007); (25) Johnstone et al. (2000); (26) Stanke et al. (2006); (27) Young et al. (2006b); (28) Kun (1998); (29) Andr´e & Montmerle (1994). – 77 – Fig. 1.- (a): Log Lint vs. Log LIR for the objects listed in Table 1. The solid line shows the results of a linear least-squares fit in log-log space; it has a slope of 0.88 and a y-intercept of 0.32. (b): Lint/LIR vs. Log LIR for the same objects. The solid line shows the average ratio of 1.7, weighted by the uncertainties in Lint. – 78 – Fig. 2.- Histogram of the un-normalized "probability" of being a galaxy, Pgal, using the method of Harvey et al. (2007b) for the 851 sources assigned such a probability in the ensemble of 82 regions with dense cores observed by c2d. The dotted line shows Log(Pgal) = −1.47 while the dashed line shows Log(Pgal) = −1.25. – 79 – Fig. 3.- Fν (normalized to 140 pc) vs. Lint for the 11 embedded protostars listed in Table 1. From left-to-right, top-to-bottom: IRAC1 (3.6 µm), IRAC2 (4.5 µm), IRAC3 (5.8 µm), IRAC4 (8.0 µm), MIPS1 (24 µm), and MIPS2 (70 µm). The lines represent the results of linear least-squares fits at each wavelength. The parameters of the fits are given in Table 2. – 80 – Fig. 4.- Same as Figure 3, except for the 1460 model SEDs calculated from the grid of 2-D radiative transfer models rather than the 11 embedded protostars listed in Table 1. – 81 – Fig. 5.- Histogram of the 70 µm 3σ point source sensitivity for the 82 regions with dense cores observed by c2d. The distribution has a mean and median of 38.6 mJy and 31.9 mJy, respectively. – 82 – Fig. 6.- Log of the luminosity sensitivity limit for embedded protostars in the 82 regions with dense cores observed by c2d (L) vs. the distance to each core (d), calculated by translating the 70 µm 3σ point source sensitivity for each core into a luminosity sensitivity using the correlation found between F70 and Lint and then scaling from 140 pc to the distance to the core. The solid line shows the relation Lint = 4 × 10−3 (d/140 pc)2 L⊙. – 83 – Fig. 7.- SED of SSTc2d J032856.64+311835.6, a source in Perseus that is representative of candidates that pass all 7 selection criteria described in §3.1 but feature SEDs inconsistent with being embedded sources. This particular source is SSS 108, a previously known pre- main sequence object with a 2 µm excess indicative of a circumstellar disk. Objects with these types of SEDs are rejected (see §3.4 for further information). – 84 – Fig. 8.- Average SEDs, weighted by 1/Lbol, for the embedded protostars identified by Enoch (2007) and Enoch et al. (2008, in preparation) in Perseus, Serpens, and Ophiuchus based on a comparison between 1.1 mm Bolocam dust continuum emission maps and the Spitzer c2d maps. The average Class 0 SED is shown in black, the average SED for Class I sources with rising (or flat) fluxes from 24 to 70 µm is shown in blue, the average SED for Class I sources with decreasing fluxes from 24 to 70 µm is shown in green, and the average SED for the Class II objects identified by Enoch et al. is shown in red. – 85 – Fig. 9.- Distribution of luminosities for the 851 objects in the 82 regions with dense cores observed by c2d classified as either candidate YSOs or candidate galaxies by the classification method of Harvey et al. (2007b). The solid line shows the distribution for the sources classified as candidate YSOs while the dashed line shows the same for the sources classified as candidate galaxies. Of the 604 out of 851 sources with LIR ≤ 0.05 L⊙, 518 (∼ 86%) are classified as candidate galaxies. – 86 – Fig. 10.- Top Left: Three-color image of L673-7 comprised of IRAC 1 (3.6 µm; blue), IRAC 2 (4.5 µm; green), and IRAC 4 (8.0 µm; red ). The white cross marks the position of Source 031 from this work, which is not detected by IRAC; it is clearly seen to fall within a dark core at 8.0 µm. Top Right: Same as top left, except the red now shows the MIPS1 (24 µm) image. A very red source is seen at the position of Source 031. Bottom: Radial profiles of the background intensity at 8 and 24 µm, centered at the position of Source 031. The background intensities are calculated in concentric annuli, each with a radius of 2′′. The radius on the x-axis is the mid-point of the annulus (for example, at a radius of 10′′ the background intensity was calculated in an annulus with its inner edge located 9′′ from the source and its outer edge located 11′′ from the source). The calculations start at radii of 5 and 7′′ respectively, for 8.0 and 24 µm. – 87 – Fig. 11.- LIR int, the internal luminosity estimated from the calculated value of LIR, vs. L70 int, the internal luminosity estimated from the 70 µm flux scaled to its value at 140 pc, for the 50 objects listed in Table 11. The dashed line shows the 1:1 line, while the dotted lines show the 2:1 and 1:2 lines. – 88 – Fig. 12.- Distribution of internal luminosities of the 50 embedded, low-luminosity protostars listed in Table 11. The internal luminosity of each protostar was estimated from the observed 70 µm flux and the correlation between the two derived in §3. The dashed lines shows our sensitivity limit to embedded protostars for the closest regions observed by c2d (Lint = 2.5 × 10−3 L⊙) and the most distant regions observed by c2d (Lint = 5 × 10−2 L⊙). – 89 – Fig. 13.- Distribution of internal luminosities, in linear bins with a size of 0.1 L⊙, for all sources listed in Table 11 with Lint ≤ 0.5 L⊙. – 90 – Fig. 14.- Figure 19 from Young & Evans (2005), a Lbol vs. Tbol BLT, with the 50 objects listed in Table 11 overlaid as large, colored points. The meaning of the colors and symbols are described in the text. The uncertainties in Lbol and Tbol, which are not shown on this plot, are dominated by the 20-60% errors introduced by incomplete, finite sampling of the source SEDs (see Appendix 2). The thick lines show the evolutionary tracks for the three models considered by Young & Evans (2005), which differ in their initial envelope mass. The thin lines show the evolutionary tracks for three models considered by Myers et al. (1998), which also differ in their initial envelope mass. The small circles are observations from the literature compiled by Young & Evans (2005). The vertical dashed lines show the Class 0/I and Class I/II Tbol boundaries from Chen et al. (1995). – 91 – Fig. 15.- Lbol/Lsmm vs. Tbol for the 50 objects listed in Table 11. The meaning of the colors/symbols are the same as for Figure 14. The Class divisions in Tbol are from Chen et al. (1995), while the Class divisions in Lbol/Lsmm are from Young & Evans (2005). – 92 – Fig. 16.- Same as Figure 14, except only showing the 15 VeLLOs (objects with Lint ≤ 0.1 L⊙). – 93 – Fig. 17.- Distribution of the sensitivity to protostars embedded within the 300 dense cores in Perseus, Serpens, Ophiuchus, and the sample of 82 regions with dense cores targeted for observations. The dashed lines show, from left to right, the sensitivity for cores in Lupus I and IV, Chamaeleon II, and Lupus III. – 94 – Fig. 18.- Monte Carlo simulation of the observation of 10,000 embedded protostars with internal luminosities distributed evenly between 10−5 and 10−1 L⊙ and randomly placed in one of the 300 dense cores. Top: Normalized distribution of Lint for the 10,000 sources. Bottom: Results of the simulation. The solid line shows the observed distribution of Lint (normalized) for the 15 VeLLOs identified by this work, while the dashed line shows the simulated observed distribution of Lint (normalized) of the simulated sources. – 95 – Fig. 19.- Distribution of internal luminosities, in linear bins of 0.01 L⊙, for the 15 embedded protostars with Lint ≤ 0.1 L⊙. The vertical dashed line shows Lint = 10−1.8 L⊙ ≈ 0.016 L⊙. The vertical dotted line shows Lint = 0.0025 L⊙. – 96 – Fig. 20.- Same as Figure 18, except with the internal luminosities of 5,000 of the 10,000 sources evenly distributed between 10−4 and 10−3 L⊙ and the internal luminosities of the other 5,000 sources evenly distributed between 10−3 and 10−1 L⊙. – 97 – Fig. 21.- Percent errors between Lbol calculated from the category 2 (left; category 3, middle; category 4, right) SED and the category 1 SED, as defined in the text. The dashed line shows the average value of this error for each category. Fig. 22.- Same as Figure 21, except for Tbol rather than Lbol.
astro-ph/0105015
1
0105
2001-05-01T23:10:36
Optical spectroscopy of X-MEGA targets I. CPD -59 2635: A New Double-Lined O type Binary in the Carina Nebula
[ "astro-ph" ]
Optical spectroscopy of CPD -59 2635, one of the O-type stars in the open cluster Trumpler 16 in the Carina Nebula, reveals this star to be a double-lined binary system. We have obtained the first radial velocity orbit for this system, consisting of a circular solution with a period of 2.2999 days and semi amplitudes of 208 and 273 km/s. This results in minimum masses of 15 and 11 Msol for the binary components of CPD -59 2635, which we classified as O8V and O9.5V, though spectral type variations of the order of 1 subclass, that we identify as the Struve-Sahade effect, seem to be present in both components. From ROSAT HRI observations of CPD -59 2635 we determine a luminosity ratio log(L_x/L_bol)~ -7, which is similar to that observed for other O-type stars in the Carina Nebula region. No evidence of light variations is present in the available optical or X-rays data sets.
astro-ph
astro-ph
Mon. Not. R. Astron. Soc. 000, 000–000 (0000) Printed 16 March 2018 (MN LATEX style file v1.4) Optical spectroscopy of X-MEGA targets I. CPD -59◦ 2635: A New Double-Lined O type Binary in the Carina Nebula J.F. Albacete Colombo1⋆ †, N.I. Morrell1⋆ ‡ §, V.S. Niemela1⋆ ¶ and M.F. Corcoran2,3 1 Facultad de Ciencias Astron´omicas y Geof´ısicas, Universidad Nacional de La Plata, Paseo del Bosque S/N, 1900 La Plata, Argentina 2 Universities Space Research Association, 7501 Forbes Blvd, Ste 206, Seabrook, MD 20706, USA 3 Laboratory for High Energy Astrophysics, Goddard Space Flight Center, Greenbelt MD 20771, USA 16 March 2018 ABSTRACT Optical spectroscopy of CPD -59◦ 2635, one of the O-type stars in the open cluster Trumpler 16 in the Carina Nebula, reveals this star to be a double-lined binary system. We have obtained the first radial velocity orbit for this system, consisting of a circular solution with a period of 2.2999 days and semi amplitudes of 208 and 273 km s−1. This results in minimum masses of 15 and 11 M⊙ for the binary components of CPD - 59◦ 2635, which we classified as O8V and O9.5V, though spectral type variations of the order of 1 subclass, that we identify as the Struve-Sahade effect, seem to be present in both components. From ROSAT HRI observations of CPD -59◦ 2635 we determine a luminosity ratio log(Lx/Lbol) ≈ −7, which is similar to that observed for other O- type stars in the Carina Nebula region. No evidence of light variations is present in the available optical or X-rays data sets. Key words: open clusters and associations: individual (Trumpler 16) – stars: binaries – early-type – stars: individual (CPD -59◦ 2635) – X-rays: stars 1 INTRODUCTION The very young Carina Nebula region contains several open clusters with a rich population of O-type stars. Among them, Trumpler 16 is one of the most conspicuous. One of its mem- bers, CPD -59◦ 2635 (V=9.27, α2000=10◦ 45′ 12.78′′, δ2000=- 59◦ 44′ 46.6′′; Massey & Johnson, 1993) has been observed in the context of the international X-Mega campaign, which involves optical spectroscopy of OB stars showing X-ray emission on ROSAT-HRI images (Corcoran 1999) k CPD - 59◦ 2635 is one of the OB stars in the neighborhood of η Car- ⋆ Visiting Astronomer, CASLEO, operated under agreement be- tween CONICET and National Universities of La Plata, C´ordoba and San Juan, Argentina. † Fellow of CIC, Prov. de Buenos Aires, Argentina. ‡ Visiting Astronomer, CTIO, NOAO, operated by AURA Inc., for NSF. § Member of Carrera del Investigador Cient´ıfico, CONICET, Argentina. ¶ Member of Carrera del Investigador Cient´ıfico, CIC, Prov. de Buenos Aires, Argentina. k (See http://lheawww.gsfc.nasa.gov/users/corcoran/xmega/ for a more comprehensive description of the XMega campaign). c(cid:13) 0000 RAS inae also detected as a bright X-ray source. Figure 1 shows an X-ray image centered on CPD -59◦2635 obtained through combination of 3 deep ROSAT HRI pointings (see below) along with the optical field from the Digitized Sky Survey. Because a massive binary system could influence its emergent X-ray flux as a result of colliding stellar winds (e.g. Chlebowski & Garmany, 1991), it is important to ver- ify the frequency of close multiple systems among the Carina OB stars which are detected as X-ray sources. We wonder, for example, if wind collision might be the physical reason that makes CPD -59◦ 2635 brighter in X-rays than its close neighbor HD 93343 very similar in both visual brightness and spectral type (see Fig. 1). CPD -59◦ 2635 has received different designations in the literature. The IDS catalogue (Jeffers et al. 1963) refers to it as IDS 10452 S 5946 as possibly forming a visual binary with HD 93343, of similar V magnitude, about 14" to the South. Stephenson & Sanduleak (1971) gave to CPD -59◦ 2635 the number 1872 in their catalogue of "Luminous Stars in the Southern Milky Way" (LSS), but misleadingly providing the cross-identification as "HD 93343?". Feinstein et al. (1973) assigned to CPD -59◦ 2635 the number 34, among the prob- able members of the open cluster Trumpler 16. Figure 1. Optical field around CPD −59◦ 2635 (left) from digitized sky survey and ROSAT HRI image (right). The fields of view are 6′ × 6′. 2 OBSERVATIONS Our observations consist of optical spectrograms of CPD -59◦ 2635, obtained during 1984 at Cerro Tololo Interameri- can Observatory (CTIO), Chile, and between 1997 and 2000 at Complejo Astron´omico El Leoncito (CASLEO). The first set of 11 observations was obtained in March 1984 at CTIO using the Carnegie Image Tube Spectrograph (CTIS) attached to the 1-m Yale telescope. These spectro- grams, covering a wavelength range from 3900 to 4900 A at a reciprocal dispersion of 43 A mm−1 were widened to 1 mm and secured on Kodak IIIa-J baked emulsion. A He-A lamp was used as a comparison source. These photographic spec- trograms were digitized with a Grant micro-densitometer at La Plata Observatory, and subsequently analyzed with IRAF ⋆⋆ routines. Spectral CCD images of CPD -59◦ 2635 were obtained at CASLEO observatory, between January 1997 and June 2000 with the 2.15-m Jorge Sahade telescope, mainly as part of the observations for the XMega campaign. We used a REOSC echelle Cassegrain spectrograph and a Tek 1024 × 1024 pixels CCD as detector to obtain 27 spectra in the wavelength range from 3500 to 6000 A at a reciprocal dis- persion of 0.17 A px−1 at 4500 A. The S/N of these data is ∼ 110 (although it changes, of course, with position within each echelle order). Four additional observations were obtained at CASLEO with a Boller & Chivens (B&C) spectrograph attached to the 2.15m telescope, using a PM 516 × 516 pixels CCD as detector, and a 600 l mm−1 diffraction grating, yielding a reciprocal dispersion of 2.5 A px−1. These spectra cover the spectral range from ∼ 3800 to 4900 A , and their S/N is ∼300. One more spectrum of CPD -59◦ 2635 was obtained ⋆⋆ Image Reduction and Analysis Facility, distributed by NOAO, operated by AURA, Inc., under agreement with NSF. at CASLEO with the REOSC spectrograph in its simple dis- persion mode, using a 600 l mm−1 grating and the Tek 1024 × 1024 CCD as detector, at a resulting reciprocal dispersion of 1.8 A px−1. The central wavelength of this observation is 4700 A and the corresponding S/N is ∼ 300. The usual series of bias, flat field and dark exposures were also secured during each observing night for every CCD data set. The CCD images were processed and analyzed with IRAF routines at La Plata Observatory. 3 RESULTS AND THEIR DISCUSSION 3.1 Radial Velocity orbit of CPD -59◦ 2635 A first inspection of our high resolution echelle spectrograms revealed double lines present in some of them, indicating that CPD -59◦ 2635 was probably a double-lined spectro- scopic binary. Figure 2 shows the behavior of He ii 4686 A line in echelle spectra of CPD -59◦ 2635, obtained at different observing dates. Radial velocities were determined from our spectra of CPD -59◦ 2635, fitting Gaussian profiles to the absorption lines. We used for radial velocity determination several lines of Hei along with Heii 4686 A which appear as the least affected by pair blending. The Pickering (4-n) series of Heii absorptions of the two binary components are not well re- solved in our spectra, these lines being broader than other Heii and Hei absorptions. We interpret this fact as indica- tive of higher pressure broadening acting in the region where these lines form, which must be deeper in the atmosphere than the formation region of Hei lines. As a consequence, these (4-n) series lines are more seriously blended than the other absorptions in the spectrum of CPD -59◦ 2635 and we decided not to include them in the average radial velocities presented in Table 1. We also derived radial velocities from Table 1. Radial-velocity measurements for CPD -59◦ 2635. HJD 2400000+ phase φ Primary km s−1 s.d. km s−1 Secondary km s−1 s.d. km s−1 45776.826† 45777.783† 45778.731† 45779.715† 45780.748† 45781.644† 45782.730† 45783.677† 45784.682† 45785.634† 45786.623† 50495.835 50498.826 50506.834 50841.793 50842.769 50843.727 50844.718 50845.723 50846.828 50847.649 50848.779 50850.724 50851.655 50852.828 50859.789‡ 50861.781‡ 50862.762‡ 50868.714‡ 51208.709 51209.714 51210.727 51211.715 51215.682 51216.653 51217.673 51218.632 51712.448 51712.478 51715.463 51716.483 51716.521 51718.555‡ 0.86 0.27 0.63 0.11 0.56 0.95 0.42 0.84 0.27 0.62 0.12 0.64 0.92 0.43 0.07 0.49 0.91 0.34 0.77 0.26 0.61 0.10 0.95 0.35 0.89 0.61 0.76 0.18 0.77 0.61 0.04 0.47 0.90 0.63 0.05 0.49 0.91 0.62 0.63 0.93 0.37 0.39 0.27 193 -214 +179 -138 +27 +37 -116 +175 -204 +181 -135 +163 +108 -68 -86 +3 +132 -168 +205 -202 +150 -133 +86 -162 +177 +112 +189 -192 +188 +134 -49 -36 +118 +137 -82 -12 +109 +136 +179 +84 -165 -131 -220 34 56 22 25 25 28 46 25 47 27 28 23 10 19 27 9 18 10 11 27 10 12 17 17 19 21 24 18 22 17 27 18 6 17 20 21 32 20 56 16 33 16 59 -132 +256 -210 +160 - - +98 -223 +219 -210 +180 -225 -131 +132 +152 - -143 +236 -271 +270 -184 +189 -101 +231 -191 - -248 +243 - -214 - - -179 -197 +80 - -149 -193 - -120 184 160 220 56 41 28 10 - - 28 61 68 30 34 18 18 13 28 - 16 29 37 21 35 24 18 24 18 - 25 33 - 21 - - 24 17 12 - 37 30 - 10 19 23 31 NOTE: Orbital phases have been calculated with ephemeris of ta- ble 2. † indicates the lower resolution observations obtained with the CITS , and ‡ those obtained either with the B&C spectro- graph or the REOSC spectrograph in simple dispersion mode. 3.2 Spectral classification of the binary components of CPD -59◦ 2635. CPD -59◦ 2635 has been previously classified by Walborn (1982) as O7Vnn; and by Levato & Malaroda (1982) as O8/9:V "+ companion ?", already pointing out its probable binary nature. Massey & Johnson (1993) classified CPD - 59◦ 2635 as O8.5V. The spectrum of CPD -59◦ 2635 (identi- fied as star number 516), shown in the last mentioned paper displays indeed double lines, apparently not noticed by the authors. Two spectra of CPD -59◦ 2635 from our lower resolu- Figure 2. Heii 4686 A absorption in the spectrum of CPD - 59◦ 2635 at different observing dates, showing the doubling of spectral lines. our lower resolution, but higher S/N CCD spectra and from our digitized photographic plates. The radial velocities were computed as unweighted mean values of individual velocities determined for each spectral line. The journal of our radial velocity observations is pre- sented in Table 1. In successive columns, we quote the Helio- centric Julian Date (HJD), the corresponding orbital phase (as explained below), the measured average radial velocities for the primary and secondary components, and their stan- dard deviations (s.d.). We identified the primary component as the one having deeper absorptions of Heii lines. From the radial velocities listed in Table 1, it was al- ready apparent that the orbital period of CPD -59◦ 2635 was of the order of a few days. A period search routine based on the modified Lafler & Kinman (1965) method (Cincotta, M´endez & Nunez 1995) applied to all the radial velocity observations of the primary component of CPD - 59◦ 2635 as listed in Table 1, yielded as the most probable period P = 2.29995 ± 0.00002 days. Initial orbital elements were estimated, leaving also the period as a free parame- ter, resulting in an orbital solution of negligible eccentric- ity (e = 0.005 ± 0.008) with no significative change in the orbital period. We therefore considered the orbit to be cir- cular, and the above mentioned value of the period to be the most probable, and proceeded to find the best fit for the remaining orbital parameters. In order to avoid as much as possible pair blending effects, we computed the orbital ele- ments of CPD -59◦ 2635 using only radial velocities derived from our high resolution observations of both binary com- ponents, obtained at the orbital phase intervals 0.1 to 0.4 and 0.6 to 0.9, of the binary period. The resulting orbital elements are listed in Table 2. Figure 3 represents the complete set of observed radial velocities as a function of the adopted orbital period, along with the circular orbital solution from Table 2. Figure 3. Radial velocity orbit for CPD -59◦ 2635. The meaning of the symbols is as follows: filled and open symbols refer to the primary and secondary stars, respectively. Circles represent REOSC-echelle spectra (bigger symbols refer to data used in the calculation of the orbital solution); squares stand for B&C and REOSC-DS spectra and triangles, for CITS spectra. Typical error bars for each data-set are also indicated. Table 2. Circular orbital elements of CPD −59◦ 2635. P [days] K1 [km s−1] K2 [km s−1] γ [km s−1] Tmax [HJD] a1.sini [R⊙] a2.sini [R⊙] M1 sin3i [M⊙] M2 sin3i [M⊙] q(M2/M1) 2.29995 ± 2 ×10−5 208 ± 2 273 ± 2 0 ± 1 2450845.664 ± 0.01 9.4 ± 0.1 12.4 ± 0.1 15.0 ± 0.5 11.4 ± 0.5 0.76 ± 0.01 NOTE: Tmax represents the time of maximum radial velocity of the primary component. tion CCD images, corresponding to nearly opposite binary phases, are illustrated in Figure 4, where a difference in the relative intensities of Hei 4471A and Heii 4542A is evident. In the upper spectrum shown in Figure 4, obtained at bi- nary phase 0.76P, Hei and Heii absorptions appear similar indicating a spectral type O7; while in the lower spectrum, obtained at the binary phase 0.18P, Hei absorption is clearly stronger than that of Heii, corresponding to a spectral type O8-9. Such spectral variations might explain the slightly dif- ferent classifications given to this star by different authors. Keeping in mind the known difficulties in classifying spectra of close binaries, and the above illustrated spectral variations, we have nevertheless tried to estimate the spec- tral types of the binary components of CPD -59◦ 2635. An inspection of our high and intermediate resolution observations showed that both stars present absorption-line ratios of Hei/Heii ≥ 1, indicating spectral types probably later than O7. From the spectra with maximum separation of the lines of the binary components, and using the classi- fication criteria described by Walborn & Fitzpatrick (1990), we derived spectral types of O8:V and O9.5:V for the pri- mary and secondary components, respectively. In order to provide an additional estimate for the spec- tral types of the binary components of CPD -59◦ 2635, we considered the Hei(4922)/Heii(5411) ratio, following the classification criteria proposed by Kerton et al. (1999). The equivalent widths (W ) of the 5411 A and 4922 A lines were measured in normalized spectra, by means of a Gaussian fit to the absorption profiles, though Heii 5411 A looks somewhat blended even at phases of maximum radial ve- locity separation (as above discussed) which causes less con- fident W determinations. We found equivalent width ratios R(W4922/W5411) of 0.54 ± 0.03 and 1.18 ± 0.04 for the pri- mary and secondary components, respectively, correspond- ing to spectral types of O8V and O9.5V, with some varia- tions depending probably on binary phase. Figure 5 shows the spectral region comprising Hei 4922 A and Heii 5411 A lines on echelle spectrograms of CPD - 59◦ 2635 near the phases of maximum separation of the com- ponents. By inspection of the Hei and Heii spectral lines observed during different phases of the binary motion, we found line depth variations that can be appreciated in Figures 2 and 5. We identify this phenomenon as the Struve-Sahade ef- fect (cf. Gies et al. 1997 and references therein) observed in Figure 4. Spectra of CPD -59◦ 2635 obtained at CASLEO with the B&C spectrograph. Spectral features marked are: H Balmer lines, Hei(+ii) 4026 A, Hei 4143 A, Heii 4200 A, Hei 4388 A, Hei 4471 A, Heii 4541 A, Ciii + Oii 4650 A blend, Heii 4686 A, Hei 4713 A, and Hei 4921 A. Note the difference in relative intensities of Hei 4471A and Heii 4541A between both spectra. Kerton et al. (1999) we found variations around one sub- class in spectral types for both binary components, going from O7 to O8 for the primary and O8.5 to O9.5 for the secondary star. However, the errors involved in the EW measurements (especially those affecting the Heii 5411 A line) are also considerable and, as a consequence, we cannot address any conclusive statement about the phase depen- dence of these variations until higher resolution and S/N observations are available. As the binary components of CPD -59◦ 2635 are prob- ably in synchronous rotation with the orbital period, we believe that the later spectral types (i.e. O8V and O9.5V respectively) are more realistic in describing each star in this binary system, being the earlier ones probably produced by photo-spheric heating on each star from its companion and/or the colliding wind region between the stars. 3.3 Physical Parameters In what follows we will estimate the physical parameters of the binary components of CPD -59◦ 2635. We adopt for this star the visual magnitude and distance modulus of Trum- pler 16 published by Massey & Johnson (1993), namely V = 9.27 and Mv − V0 = 12.55 ± 0.08, close to the value of Mv − V0 = 12.6 obtained by Feinstein et al. (1973). Also from Massey & Johnson (1993), we take, for CPD -59◦ 2635, EB−V = 0.54, R = 3.2 and thus V0 = 7.54. In order to ob- tain individual absolute magnitudes, we need an estimate of the luminosity ratio of the binary components. We applied the corrected integrated-absorption method of Petrie (1940) in the way described by Niemela & Morrison (1988). We used equivalent widths measured for Hei 4387 and 4471 and Heii Figure 5. Hei 4922 A and Heii 5411 A lines in the spectrum of CPD -59◦ 2635 at phases 0.26 and 0.77 massive close binaries, which consists in the deepening of the spectral lines of the secondary star when it approaches to the observer. In our spectra, variable line depths seem to be present in both binary components although stronger variations are observed in the secondary star. Applying again the classification criteria proposed by 4686 A absorptions in spectra where those features are bet- ter resolved. We corrected for continuum overlapping using equivalent widths of the same lines measured for individ- ual stars of similar spectral types, taken from Mathys, 1988. Then we calculated for each selected spectral line the quo- tient (L2/L1=< (B.Ka)/(A.Kb) >), where A is the equiv- alent width of the line measured in the spectrum of the most intense component, B is the equivalent width of the same line as observed in the weaker component, and Ka,b are the equivalent widths measured for single stars of spec- tral types O8 V and O9 V, respectively. Performing these measurements in several spectra of CPD -59◦ 2635 observed near phases of maximum separation of the components we obtained an average value of L2/L1 = 0.45 ± 0.14, which we have used in the following calculations. With this luminosity ratio and the adopted distance modulus, we obtained indi- vidual absolute magnitudes of MV = -4.61 ± 0.1 and -3.74 ± 0.1 for the primary and secondary components, respectively. Our classification of the components of CPD -59◦ 2635 as O8 V and O9.5 V corresponds to Teff = 38450 K and BC = -3.68 for the primary, and Teff = 34620 K and BC = -3.36 for the secondary, according to the calibration of effective temperatures (Teff) and bolometric corrections (BC) proposed by Vacca, Garmany & Shull (1996, here- after VGS). However, according to Schmidt-Kaler (1982), the corresponding effective temperatures and bolometric co- rrections would be Teff = 35800 K and BC = -3.54 for the primary, and Teff = 31500 K and BC = -3.25 for the sec- ondary. Therefore, depending on which calibration we adopt, we would find somewhat different values for the bolometric magnitudes, and thus, luminosities of each binary compo- nent. Knowing the luminosities and effective temperatures, we can derive the radii of the stars, that we want to com- pare with the radii of the critical Roche lobes. Those were estimated using the expression given by Paczynski (1971): r1 sin i = a sin i(0.38 + 0.2 log(M1/M2)) for a "mean" Roche radius of r1 and separation a. We ob- tained individual Roche radii of r1 sin i = 8.8 R⊙ and r2 sin i = 7.8 R⊙. We need to know something about the inclination of the orbital plane in order to compare these critical Roche radii with the Stefan-Boltzmann radii of the stars. The physical parameters derived for the binary compo- nents are summarized in Table 3. Under the assumption that the system is in synchronous rotation (which seems reasonable in a massive binary of short period) we derived probable rotational velocities for its components (quoted in Table 3 which are higher than the observed rotational velocities of single stars of similar spectral types (e.g. Slettebak et al., 1975 ; Conti & Ebbets, 1977). We tried to estimate the projected rotational velocities (V sin i) comparing the observed absorption profiles of Hei 4388 A and 4471 A with flux profiles from non-LTE model atmospheres by Auer & Mihalas (1972). We chose models corresponding to Teff of 40000, 35000 and 30000 K and log g = 4 to represent the binary components. The model pro- files were convolved with different rotational velocity pro- files, finding satisfactory agreement with observations for V sin i values of 180 ± 25 km s−1 and 140 ± 30 km s−1 for the primary and secondary components, respectively. Com- Table 4. Three deep ROSAT HRI pointings which include CPD −59◦ 2635. Sequence Identification Begin Date End Date Exposure RH900385N00 RH900385A03 RH202331N00 1992-07-31 1994-07-21 1997-12-23 1992-08-02 1994-07-29 1998-02-10 11527 40555 47095 paring these results with the calculated rotational velocities, we obtain, for the inclination of the orbital plane, values of 64◦ and 79◦, for the primary and secondary, respectively, using the radii derived through the calibration by VGS, or 56◦ and 59◦, for primary and secondary, respectively, using the calibration by Schmidt-Kaler (1982). However, the er- rors involved in the V sin i determination give room for a large range of inclinations. Assuming the mass-spectral type relation for normal O-type stars by VGS (1996), we can expect masses near 25 and 21 M⊙ for the individual binary components of CPD - 59◦ 2635. These values are similar to (or slightly larger than) those obtained from the observation of eclipsing binary sys- tems with O8V and O9V components (e.g. EM Car, Ander- sen & Clausen, 1989; Y Cyg, Burkholder, Massey & Mor- rell, 1997; CQ Cep, Kartascheva & Svechnikov, 1989). This also points to an orbital inclination inclination of 56◦ ± 6◦, similar to the values estimated from the line widths. If these estimates are correct, the system is not likely to present eclipses. However, some kind of light variations may occur due to tidal deformation, considering that both com- ponents are hot luminous stars in a close binary system. Also, if our guess for the inclination is as supposed, the sys- tem would be detached, with both components within their critical Roche lobes, as we would expect for a young binary with non evolved components. 4 X-RAY EMISSION CPD −59◦ 2635 was detected as a serendipitous X-ray source during numerous pointing in the Carina Nebula by the ROSAT X-ray satellite observatory with both the Posi- tion Sensitive Proportional Counter (PSPC) and the High Resolution Imager (HRI). In the PSPC images the star is unresolved from other nearby X-ray sources (notably the O stars HD 93343 and Tr 16 #182) due to the rather coarse spatial resolution of the PSPC (∼ 1′). The HRI has finer spatial resolution (∼ 10′′) and so can better resolve CPD −59◦ 2635 from surrounding sources, providing a more ac- curate measure of the uncontaminated X-ray emission from the star. Table 4 lists 3 deep ROSAT HRI pointing which include CPD −59◦ 2635. The source lies about 4.2′ off axis, and at this position the 50% encircled energy radius is about 3′′. We extracted source counts from these 3 HRI sequences in an 8′′ re- gion centered on CPD −59◦ 2635. We extracted background counts from a region of blank sky centered at α2000 = 10h 45m 19.6s, δ2000 = −59◦ 44′ 43.2′′ using an extraction radius of 40′′ to reduce the statistical uncertainty in the net rate. Figure 6 shows the extracted net light curve for the Table 3. Physical Parameters of CPD -59◦ 2635. Primary component VGS 38450 Calibration Teff [K] Mbol L [L⊙] RS−B [R⊙] RS−B/RRoche−lobe Vrot [km.s−1] Schmidt-Kaler Secondary component VGS Schmidt-Kaler 35800 34620 31500 -8.3 ± 0.15 -8.1 ± 0.15 -7.1 ± 0.15 -7.0 ± 0.15 162000 ± 20000 143000 ± 20000 54000 ± 11000 49000 ± 11000 9.1 ± 0.4 0.8 ± 0.1 200 ± 10 9.8 ± 0.4 0.9 ± 0.1 215 ± 10 6.5 ± 0.5 0.7 ± 0.1 142 ± 15 7.4 ± 0.5 0.8 ± 0.1 164 ± 15 Notes: a) RS−B means Stefan-Boltzmann radii. b) Theoretical Roche-lobe radii were desafected by the sin i factor using a probable average value for the inclination of the orbital plane, namely 60◦ . 0.004 0.003 s / c t e N I R H 0.002 0.001 0.000 48500 49000 49500 50000 50500 51000 JD - 2,400,000 Figure 6. HRI X-ray light curve of CPD −59◦ 2635. source. There is no evidence for variability. Because the HRI has no spectral resolution, we could not determine X-ray lu- minosity by direct spectral fitting. Rather, to determine the X-ray luminosity we converted the total net count rate to luminosity by assuming a Raymond-Smith thermal source spectrum with a temperature kT = 0.5 keV and an ab- sorbing column NH = 2 × 1021 cm−2, which should reason- ably approximate the X-ray emission from OB type stars in the Carina nebula. At 4′ off-axis, the HRI vignetting cor- rection is 1.01 (David, et al., 1997); the total net count rate, corrected for vignetting, is 1.35 ± 0.14 × 10−3 HRI counts s−1. With our assumed spectral parameters, this cor- responds to an X-ray luminosity Lx = 3.8 × 1031 ergs s−1 in the ROSAT band (0.2 − 2.4 keV), uncorrected for absorp- tion, using Mv − V0 = 12.55. After correcting for absorp- tion, the X-ray luminosity at the source is approximately Lx,unabs = 9 × 1031 ergs s−1 in the ROSAT band. With a total bolometric luminosity Ltot ∼ 200000L⊙ the ratio of the unabsorbed X-ray luminosity to the total luminosity is log Lx/Ltot ∼ −7, similar to the canonical value of this ratio in the ROSAT band (Berghoffer et al. 1997). Accord- ing to this, CPD -59◦ 2635 does not show excess in its X-ray flux compared to other O-type stars, and the question about why it looks brighter than its neighbour HD 93343 in the ROSAT HRI image of Figure 1 remains to be clarified when more observations of both stars are available. 5 CONCLUSIONS In the present study we demonstrate that CPD -59◦ 2635 is a close binary system in a circular orbit. Both binary compo- nents are O-type stars, and the short period of binary motion suggests strong interactions between the stellar winds. Variations of the order of one subclass in the spectral types of both components are observed, probably related to the phenomenon known as Struve-Sahade effect. A possible explanation for the Struve-Sahade effect, analyzed by Gies et al. (1997), is that it is present in systems expected to contain colliding winds with X-ray generation from the bow shock between the stars. However, the observed X-ray light curve does not show any significant variations, and moreover, the ratio Lx/Lbol seems to be similar to that observed for other O-type stars. This star is a massive close binary, with hot, luminous components (O8V + O9.5V) and thus a good candidate for further exploration of colliding wind effects. With the estimated inclination ∼ 60◦, we can conclude that the stellar radii are smaller than the corresponding Roche lobes, the system being detached, as one would ex- pect from the evolutionary status of a member in a cluster still containing unevolved O3 stars. Though the system is not expected to present eclipses, future photometric studies could reveal the presence of phase locked light variations produced by tidal deformation of the binary components. This would provide the opportunity of making a better estimate of the inclination of the orbital plane, and consequently lead to the derivation of absolute individual masses for the components of CPD -59◦ 2635. No need to recall that this is of fundamental astrophysical interest concerning O-type stars. 6 ACKNOWLEDGEMENTS We thank the director and staff of CASLEO for techni- cal support and kind hospitality during the observing runs. N.I.M. is indebted to the director and staff of CTIO for the use of their facilities. We acknowledge use at CASLEO of the CCD and data acquisition system supported under U.S. NSF grant AST-90-15827 to R. M. Rich. The Digitized Sky Survey in the southern hemisphere is based on photographic data obtained using The UK Schmidt Telescope. The UK Schmidt Telescope was operated by the Royal Observatory Edinburgh, with funding from the UK Science and Engineering Research Council, until 1988 June, and thereafter by the Anglo-Australian Observatory. Orig- inal plate material is copyright the Royal Observatory Ed- inburgh and the Anglo-Australian Observatory. The plates were processed into the present compressed digital form with their permission. The Digitized Sky Survey was produced at the Space Telescope Science institute under US Government grant NAG W-2166. This research has made use of NASA's Astrophysics Data System Abstract Service. This research has made use of data obtained from the High Energy As- trophysics Science Archive Research Center (HEASARC), provided by NASA's Goddard Space Flight Center. Useful discussions with R. Barb´a are very much appreciated. REFERENCES Andersen J., & Clausen J.V., 1989, A&A, 213, 183 Auer L. H., & Mihalas D. 1972, Ap.J.S., 24, 193. Berghoffer, T., Schmitt, J. H. M. M., Danner, R., and Cassinelli, J., 1997, A&A, 322, 167. Burkholder V., Massey P. & Morrell, N.I., 1997, ApJ, 490, 328 Chlebowski, T., 1989, 342, 1091. Chlebowski T., & Garmany C. D. 1991, AJ, 368, 241. Chlebowski, T., Harnden, Jr., F. R., & Sciortino, S. 199, ApJ, 341, 427. David, L., et al., 1997, ROSAT High Resolution Imager Calibration Report, RSDC/SAO. Conti P. S., & Burnichon M. L. 1975, A&A, 38, 467. Conti P. S., & Ebbets, D. 1977, ApJ, 213, 438. Corcoran M. F. 1999, Rev Mex AASC, 8, in press David P., Goldwurm A., Murakami T.,Paul J., Laurent P., Goldoni P. 1997, A&A., 322, 229D. Feinstein A., Marraco H. G., & Muzzio J. C. 1973, A&AS., 12, 331. Gies D. R., Bagnuolo W. G. Jr., & Penny L. R. 1997, AJ, 479, 408. Jeffers H. M., van den Bos W.H., Greeby F. M. 1963, Index Ca- talogue of Visual Double Stars, Univ. California, Publ Lick Obs., Vol XXI. Kartascheva T.A. & Svechnikov M.A., 1989, Astrofiz. Issled. Izv. SAO, 28, 3 Kerton C. R., Ballantyne D. R., & Martin P. G. 1999, AJ, 117, 2493. Lafler J., & Kinman T. D. 1965, ApJS, 11, 199. Levato H., & Malaroda S. 1982, PASP, 94, 807. Massey P., & Johnson J. 1993, AJ, 1053, 980. Mathys, G. 1988, A&AS, 76, 427. Niemela, V. S., & Morrsion, N. D. 1988, PASP, 100, 1436. Packzynski B. 1971, Ann. Rev. Astron. Astrphys. , 9, 183. Petrie, R.M. 1940, Publ. Dom. Astroph. Obs. Victoria, 7, 205. Schmidt & Kaler 1982, in "Landolt-Bornstein, NS", Vol. 2, p. 455. Slettebak A., Collins G. W., Boyce P. B., White N. M., & Parkin- son T. D. 1975, Ap,J.S. 29, 137. Stephenson C. B., & Sanduleak N. 1971, Publ. Warner & Swasey Obs., 1, 1. Stevens, I. R., Blondin, J. M., & Pollock, A.M.T., 1992, ApJ, 386, 265. Usov, V. V., 1992, ApJ, 389, 635. Vacca W. D., Garmany C. D. & Shull J. M. 1996, ApJ, 460, 914 (VGS) Walborn N. 1982, AJ, 87, 1300. Walborn N., & Fitzpatrick E. 1990, PASP, 102, 379. This figure "figure1.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/astro-ph/0105015v1
astro-ph/9704059
2
9704
1997-10-14T08:28:52
Deconvolution with correct sampling
[ "astro-ph" ]
A new method for improving the resolution of astronomical images is presented. It is based on the principle that sampled data cannot be fully deconvolved without violating the sampling theorem. Thus, the sampled image should not be deconvolved by the total Point Spread Function, but by a narrower function chosen so that the resolution of the deconvolved image is compatible with the adopted sampling. Our deconvolution method gives results which are, in at least some cases, superior to those of other commonly used techniques: in particular, it does not produce ringing around point sources superimposed on a smooth background. Moreover, it allows to perform accurate astrometry and photometry of crowded fields. These improvements are a consequence of both the correct treatment of sampling and the recognition that the most probable astronomical image is not a flat one. The method is also well adapted to the optimal combination of different images of the same object, as can be obtained, e.g., from infrared observations or via adaptive optics techniques.
astro-ph
astro-ph
Deconvolution with correct sampling P. Magain1 Institut d'Astrophysique, Universit´e de Li`ege, 5, avenue de Cointe, B -- 4000 Li`ege Belgium email: [email protected] F. Courbin Institut d'Astrophysique, Universit´e de Li`ege, 5, avenue de Cointe, B -- 4000 Li`ege Belgium email: [email protected] URA 173 CNRS-DAEC, Observatoire de Paris, F-92195 Meudon Principal CEDEX, France S. Sohy Institut d'Astrophysique, Universit´e de Li`ege, 5, avenue de Cointe, B -- 4000 Li`ege Belgium email: [email protected] ABSTRACT A new method for improving the resolution of astronomical images is presented. It is based on the principle that sampled data cannot be fully deconvolved without violating the sampling theorem. Thus, the sampled image should not be deconvolved by the total Point Spread Function, but by a narrower function chosen so that the resolution of the deconvolved image is compatible with the adopted sampling. Our deconvolution method gives results which are, in at least some cases, superior to those of other commonly used techniques : in particular, it does not produce ringing around point sources superimposed on a smooth background. Moreover, it allows to perform accurate astrometry and photometry of crowded fields. These improvements are a consequence of both the correct treatment of sampling and the recognition that the most probable astronomical image is not a flat one. The method is also well adapted to the optimal combination of different images of the same object, as can be obtained, e.g., from infrared observations or via adaptive optics techniques. Subject headings: Methods: Numerical - Observational - Data analysis 1Maıtre de Recherches au Fonds National de la Recherche Scientifique (Belgium) -- 2 -- 1. Deconvolution Any recorded image is blurred whenever the instrument used to obtain it has a finite resolving power : for example, the image of a point source seen through a telescope has an angular size which is inversely proportional to the diameter of the primary mirror. If the instrument is ground-based, the image is additionally degraded by the turbulent motions in the earth's atmosphere. Much effort is presently devoted to the improvement of the spatial resolution of astronomical images, either via the introduction of new observing techniques (e.g. interferometry or adaptive optics, see L´ena 1996, Enard et al. 1996) or via a subsequent numerical processing of the image (deconvolution). It is, in fact, of major interest to combine both methods to reach an even better resolution. An observed image may usually be mathematically expressed as a convolution of the original light distribution with the "total instrumental profile" -- the latter being the image of a point source obtained with the instrument considered, including the atmospheric perturbation (seeing) if relevant. The total blurring function is called the Point Spread Function (PSF) of the image. Thus, the imaging equation may be written : d(~x) = t(~x) ∗ f (~x) + n(~x) (1) where f (~x) and d(~x) are the original and observed light distributions, t(~x) is the total PSF and n(~x) the measurement errors (noise) affecting the data. In addition, on all modern light detectors (e.g. CCDs whose pixels have finite dimensions), the observed light distribution is sampled, i.e. known only at regularly spaced sampling points. The imaging equation for a sampled light distribution then becomes : di = N Xj=1 tij fj + ni (2) where N is the number of sampling points, dj, fj, nj are vector components giving the sampled values of d(~x), f (~x), n(~x) and tij is the value at point j of the PSF centered on point i. The aim of deconvolution may be stated in the following way : given the observed image d(~x) and the PSF t(~x), recover the original light distribution f (~x). Being an inverse problem, deconvolution is also an ill-posed problem, and no unique solution can be found, especially in the presence of noise. This is due to the fact that many light distributions are, after convolution with the PSF, compatible within the error bars with the observed image. Therefore, regularization techniques have to be used in order to select a plausible solution amongst the family of possible ones and a large variety of deconvolution methods have been proposed, depending on the way this particular solution is chosen. A typical method is to minimize the χ2 of the differences between the data and the convolved model, with an additional constraint imposing smoothness of the solution. fj is then the light distribution which minimizes the function : S1 = N Xi=1 1 σ2 N i  Xj=1  2 tijfj − di  + λ H(f1, · · · , fN ) (3) -- 3 -- where the first term in the sum is the χ2 with σi the standard deviation of the image intensity measured at the ith sampling point, H is a smoothing function and λ a Lagrange parameter which is determined so that the reconstructed model is statistically compatible with the data (χ2 ≃ N ). j=1 fj, one If H(f1, · · · , fN ) = PN obtains the so-called maximum entropy method for image deconvolution (Narayan and Nityananda 1986, Skilling and Bryan 1984). i=1 pi ln pi, with pi the normalized flux at point i, pi = fi/PN In order to choose the correct answer in the family of possible solutions to this inverse problem, it is also very useful to consider any available prior knowledge. One such prior knowledge is the positivity of the light distribution : no negative light flux can be recorded, so that all solutions with negative values may be rejected. The maximum entropy method automatically ensures positivity of the solution. This is also the case, under certain conditions, for other popular methods, such as the Richardson-Lucy iterative algorithm (Richardson 1972, Lucy 1974). Most of the known deconvolution algorithms suffer from a number of weak points which strongly limit their usefulness. The two most important problems in this respect are the following : (1) traditional deconvolution methods tend to produce artefacts in some instances (e.g. oscillations in the vicinity of image discontinuities, or around point sources superimposed on a smooth background) ; (2) the relative intensities of different parts of the image (e.g. different stars) are not conserved, thus precluding any photometric measurements. In the next sections, we identify a plausible cause of these problems and show how to circumvent it. 2. Sampling The sampling theorem (Shannon 1949, Press et al. 1989) determines the maximal sampling interval allowed so that an entire function can be reconstructed from sampled data. It states that a function whose Fourier transform is zero at frequencies larger than a cutoff frequency ν0 is fully specified by values spaced at equal intervals not exceeding (2 ν0)−1. In practice, for functions whose Fourier transform does not present such a cutoff frequency, ν0 may be taken as the highest frequency at which the Fourier transform emerges from the noise. The imaging instruments are generally designed so that the sampling theorem is approximately fulfilled in average observing conditions. A typical sampling encountered is ∼ 2 sampling intervals per Full Width at Half Maximum (FWHM) of the PSF (this does certainly not ensure good sampling for high signal-to-noise (S/N) images, but is roughly sufficient at low S/N). The main problem with classical deconvolution algorithms is the following : if the observed data are sampled so that they just obey the sampling theorem, the deconvolved data will generally violate that same theorem. Indeed, increasing the resolution means recovering highest Fourier frequencies, thus increasing the cutoff frequency, so that the correct sampling becomes denser. One might object that some deconvolution algorithms, which allow a different sampling in the deconvolved image, could overcome this problem : it would be possible to keep a correct sampling by shortening the sampling interval. This is however an illusory solution, since the only limit on the frequency components present in an arbitrary image comes from the PSF of the instrument -- 4 -- used to record it. Removing the effect of the PSF would allow the presence of arbitrary high frequency components, and thus an infinitely small sampling interval would have to be used. This is particularly true if the image contains point sources, which is generally the case for astronomical images. Indeed, the angular diameters of most stars (≪ 0.001 arcsec) are so small compared to the sampling interval (∼ 0.1 arcsec) that they may be considered as point sources ("δ-functions"). In such an instance, it would be hopeless to reduce the sampling interval in an attempt to obtain a good sampling of such "δ-functions". This is one of the sources of some of the artefacts present in the deconvolved images and, in particular, of the "ringing" around point sources superimposed on a diffuse background. The origin of this "ringing" may be understood in the following way. Let us assume that we have a continuous (i.e. not sampled) noise-free image of a field containing point sources, and observed with an instrument having a known PSF. For simplicity, we restrict our considerations to one-dimensional images. If we can perfectly deconvolve this image, we shall obtain a solution f (x) in which each point source is represented by a Dirac δ-function. Now, let us assume that we have the same image, but sampled on N points, with a sampling step ∆x. The Fourier transform of its deconvolution may be obtained from the Fourier transform F (ν) of the continuous deconvolution in the following way : repeat periodically F (ν) with the Nyquist frequency νN y = (2∆x)−1, take the sum of all these periodical replicas at each frequency point, isolate one period and sample it on N equally spaced frequency points. Isolating one period means multiplying the Fourier transform by a rectangular ("box") function which equals 1 in an interval of length νN y and 0 outside. Now, a convolution in the image domain translates into a simple product in the Fourier domain, and vice-versa. This multiplication by a box function in the Fourier domain is thus equivalent, in the image domain, to a convolution by the Fourier transform of the box function, which is a function of the form sin x/x. The solution of the deconvolution problem for a sampled image with point sources is thus the (sampled) convolution of the exact solution f (x) with a function of the form sin x/x. Each δ-function is thus replaced by an oscillatory sin x/x function, which explains the ringing. Another, more intuitive, explanation of the same effect is the following. If a point source is located between two sampling points (as will generally be the case), in order to correctly reproduce its position, the deconvolution algorithm will have to distribute its intensity over several sampling points. But, then, the width of the source will be too large and ringing will appear as the algorithm attempts to decrease the intensity on the edges of the reconstructed source, in order to keep the convolved model as close as possible to the observed data. In fact, it is not possible to correctly reproduce both the position and the width of a sampled point source. To reproduce the zero width, the full signal must be concentrated on a single sampling point. On the other hand, to reproduce the position with a precision which is better than the sampling interval, the signal has to be distributed over several points. -- 5 -- 3. Solution The correct approach to this sampling problem is thus not to deconvolve with the total PSF t(~x), but rather with a narrower function s(~x) chosen so that the deconvolved image has its own PSF r(~x) compatible with the adopted sampling. These three functions are simply related by : t(~x) = r(~x) ∗ s(~x) (4) Note that a similar decomposition was proposed, in a completely different context (reduction of artefacts in maximum-likelihood reconstructions for emission tomography), by Snyder et al. (1987). The shape and width of r(~x) can be chosen by the user. The only constraint is that Eq. (4) admits a solution s(~x). The function s(~x) by which the observed image has to be deconvolved is thus obtained as the deconvolution of the total PSF t(~x) by the final PSF r(~x). Of course, the sampling interval of the deconvolved image does not need to be equal to the sampling interval of the original image, so that r(~x) may be much narrower than t(~x), even if the original sampling would not allow it. Choosing a sufficiently narrow r(~x) effectively insures that Eq. (4) will admit a solution s(~x). Note also that, contrary to other traditional methods (the success of which depends crucially on the effectiveness of the positivity constraint), we have no positivity constraint on the PSFs r(~x) and s(~x). Thus, the deconvolution algorithm should not attempt to determine the light distribution as if it were obtained with an ideal instrument (e.g. a space telescope with a primary mirror of infinite size). This is forbidden as long as the data are sampled. Rather, the aim of deconvolution should be to determine the light distribution as if it were observed with a better instrument (e.g. a 10 m space telescope). Deconvolution by s(~x) ensures that the solution will not violate the sampling theorem. It also has a very important additional advantage : if the image contains point sources, their shape in the deconvolved image is now precisely known : it is simply r(~x). This is a very strong prior knowledge, and it may be used to constrain the solution f (~x), which can now be written : f (~x) = h(~x) + ak r(~x − ~ck) M Xk=1 (5) where M is the number of point sources, for which ak and ~ck are free parameters corresponding to their intensities and positions, and h(~x) is the extended component of the solution, i.e. generally a rather smooth background. We can use another prior knowledge to constrain the solution : we know that the background h(~x) can also be written as the convolution of some function h′(~x) with the PSF of the solution r(~x) : h(~x) = r(~x) ∗ h′(~x) (6) However, we cannot use that decomposition directly and determine h′(~x) instead of h(~x) because h′(~x) might violate the sampling theorem, even if it does not contain point sources. Rather, we may use this knowledge to impose smoothness of h(~x) on the scale length of r(~x). -- 6 -- So, instead of regularizing the solution by a global function such as the entropy, we use a function imposing local smoothness of h(~x) on the known scale length. We thus choose the solution which minimizes the function : S2 = N Xi=1 1 σ2 N i  Xj=1  sij hj + M Xk=1 ak r(~xj −~ck)! − di  2 + λ N Xi=1  hi − N Xj=1 2 rij hj  (7) with respect to the unknowns hi (i = 1, . . . , N ), ak and ~ck (k = 1, . . . , M ). Although the smoothing term in the right-hand side of Eq. (7) will not force h(~x) to exactly obey Eq. (6), it essentially contains the Fourier components with frequencies higher than those of the deconvolved PSF r(~x). Minimizing this term will force the background component to contain only the frequencies compatible with r(~x) and, thus, with the adopted sampling. One additional improvement may be introduced. In general, the Lagrange multiplier λ is chosen so that χ2 ≃ N . This ensures that the fit is statistically correct globally. However, some regions of the image may be overfitted, and others may be underfitted. In practice, this will generally be the case : although the residuals will be correct on the average, they will systematically be too small in some parts of the image and too large in other parts. To avoid this problem, one may replace the smoothing function by : H(f1, · · · , fN ) = N Xi=1 2 λi hi − N Xj=1 rij hj  (8) where λi is the value at the ith sampling point of a function λ(~x) which is chosen so that the residuals of the fit are correct locally, i.e. that they are statistically distributed with the correct standard deviation in any sub-part of the image. In practice, an image of the square of the normalized residuals (observed data minus convolved model, divided by sigma) is computed and then smoothed with an appropriate function, so that any value is replaced by a weighted mean on a neighborhood containing a few dozens of pixels. The parameter λ is then adjusted until this image is close to one everywhere. 4. Examples Our deconvolution program implements the ideas exposed in the preceding section. The light distribution aimed at is written as the sum of a smooth background plus a number of point sources. The sampling step of the deconvolved image is chosen, as well as the final PSF r(~x), compatible with this sampling (in general, we adopt a gaussian function, with a few pixels FWHM). Approximate values of the unknowns are chosen and the function S2 is computed, together with its derivatives with respect to all variables. The minimum of S2 is then searched for, using an algorithm derived from the classical conjugate gradient method (Press et al. 1989). The fit's residuals are then computed and a check of their statistical correctness is performed. If this test is not satisfied, the Lagrange multiplier λ is replaced by a variable λ(~x) which is varied until the residuals conform to the statistical expectations. -- 7 -- The present version of the program runs on PCs and workstations and can handle images of reasonable size (e.g., 256 × 256 pixels) containing up to several hundreds of stars. The main weakness of the present implementation is related to the conjugate gradient algorithm, which is not always able to find the global minimum of the function, especially when the number of point sources is large. We are presently working on a new optimization technique which would allow our method to be applied to the photometry of crowded fields with thousands of stars. It may not seem obvious at first sight to select the correct number of point sources to be included in the solution. However, the algorithm allows to constrain this number in a very efficient way : if too few point sources are entered, it will generally be impossible to obtain statistically correct residuals locally, in all sub-parts of the image. On the other hand, if too many point sources are considered, the algorithm will either attribute essentially the same position or negligible intensities to several of them. Our methodology is thus to model the data with the minimum number of point sources necessary to yield statistically correct residuals locally in all sub-parts of the image, in the sense described at the end of the preceding section. Figure 1 compares the results of our new deconvolution algorithm to those of three classical methods in the case of a simulated star cluster partly superimposed on a smooth background (e.g. a distant elliptical galaxy). The input point sources were selected from the observed image alone, without any prior knowledge of the exact solution. It is clear that our result is free from the artefacts present in the other methods and that it allows an accurate reconstruction of the original light distribution. Another important property of our technique is that it allows, contrary to the other ones, an accurate measurement of the positions and intensities of the point sources. This point will be discussed more extensively in the next section. An application to real astronomical data is shown on Fig. 2, which displays a mediocre resolution image of the "Cloverleaf", a gravitationally lensed quasar (Magain et al. 1988), together with the deconvolved version, using a sampling interval twice as short. The four lensed images, which were unresolved in the original data, are completely separated after deconvolution. The deduced fluxes are fully compatible with those measured on higher resolution images and, although the original resolution is 1.3 arcsec only and the pixel size is 0.35 arcsec, the deduced image positions are accurate to 0.01 arcsec. Figure 3 illustrates the deconvolution of an image of the compact star cluster Sk 157 in the Small Magellanic Cloud (Heydari-Malayeri et al. 1989). The original image was obtained with the ESO/MPI 2.2m telescope at La Silla, in average seeing conditions (1.1 arcsec FWHM). While the original maximum entropy deconvolution (Heydari-Malayeri et al. 1989) allowed to resolve the cluster into 12 components, our new algorithm detects, from the same input data, more than 40 stars in the corresponding area. Another important application of our algorithm is the simultaneous deconvolution of different images of the same field. These images may be obtained with the same instrument or with different ones. The solution is then a light distribution which is compatible with all the images considered. Our technique even allows to let, e.g., the intensities of the point sources converge to different values in the different images, so that variable objects may be considered. This technique should be very useful for the photometric monitoring of variable objects in crowded fields or overimposed -- 8 -- on a diffuse background (e.g. Cepheids in distant galaxies, gravitationally lensed QSOs,...). Figure 4 illustrates this simultaneous deconvolution on simulated images, the first of which has a good resolution but a poor S/N (as might be obtained with a space telescope) and the second one a low resolution and a high S/N (a typical image from a large ground-based telescope). Contrary to Lucy's method (Hook and Lucy 1992) which is very sensitive to the noise present in one of the images, our technique allows to reliably recover both the high resolution of the space image and the hidden information content of the ground-based one. In the same spirit, our algorithm is well adapted to the processing of images obtained with infrared or adaptive optics techniques. In the latter, numerous short exposures of the same field are usually obtained, the shape of the mirror being continuously adapted to correct for atmospheric distortions. So, the observations consist in a number of images of the same field, each of them having its own PSF. Performing a simple sum results in an image whose spatial resolution is typical of the average observing conditions, while a simultaneous deconvolution not only allows to take count of the best conditions, but even results in an improved resolution by optimally combining the information content of the different images. A simple illustration of these considerations is provided by Fig. 5, which shows the simultaneous deconvolution of four adaptive-optics-like images of the same field, where the PSF as well as the image centering vary from one observation to the other. Of course, the PSF needs to be known for each individual observation, but only with an accuracy comparable to that of the observation itself. 5. Astrometric and photometric accuracy Traditional deconvolution methods are notoriously unable to give photometrically accurate results. Two main reasons for that are readily identified. First, as we have already mentioned, these methods generally produce rings when point sources are overimposed on a diffuse background. In fact, these rings tend to appear as soon as the positivity constraint is inefficient to inhibit them, that is, as soon as some flux is distributed around the point sources. This is most clearly seen when this flux is in the form of a smooth background, but the effect is also present if the flux is distributed among, e.g., a number of fainter stars. In this case, the rings around the star considered will interfere with the intensity in the neighbouring sources, and the photometry of the latter ones will be affected. A second photometric bias comes from the fact that, among the family of possible solutions to the inverse problem, most classical algorithms select, in one way or another, the smoothest one according to some criterion. These algorithms thus produce images where the peaks corresponding to point sources deviate as little as possible with respect to the background -- provided, of course, that the model fits the data. This implies a systematic underestimate of the intensity peaks, and thus, a photometric bias. An example of these effects is illustrated by the deconvolution of an image of two point sources with varying separation. A simple image was constructed, with two point sources having an intensity ratio of 0.1, and convolved with a Gaussian PSF of 7 pixels FWHM, plus some gaussian -- 9 -- noise so that the peak S/N ratio reaches 100. Figure 6 shows the deduced intensity ratio as a function of the source separation, as derived after deconvolution with the maximum entropy method. It clearly shows that the photometry is not preserved, even when the two stars are separated by nearly two FWHMs. For more details on the photometric accuracy of deconvolution algorithms (in the special case of HST images), see Busko (1994). Our algorithm naturally avoids these two biases. Indeed, the fact that the sampling theorem is obeyed in the deconvolved image, combined with the fact that no smoothing of the point sources is attempted, naturally ensure that no ringing is present around the star peaks, and that no bias will appear as a consequence of smoothing. This is illustrated in Fig. 7, which shows the results of a photometric test applied to a synthetic field containing 200 stars in a 128× 128 pixels image. The positions and central intensities were selected at random, and nearly all the stars are blended to varying degrees (197 stars out of 200 have the nearest neighbour within 2 FWHMs). Moreover, these stars are superimposed on a variable background. Figure 7 clearly shows that no systematic error is present, and that the intensities of all but the most severely blended objects are reproduced with errors compatible with the photon noise. Figure 8 illustrates the astrometric accuracy of our algorithm. For the brightest stars, the positional accuracy is generally better than 0.1 pixel, even in very severe blends. The positions of stars with a blend between 1 and 2 FWHM is generally accurate within 0.02 pixel at high S/N, and within 0.1 pixel otherwise. There exits another deconvolution algorithm which claims to achieve a high photometric quality, namely the so-called two-channel Lucy method (Lucy 1994, Hook and Lucy 1994). As our algorithm, the two-channel Lucy method is based on a decomposition of the deconvolved model into point sources and background. The main problem with that method is that, contrary to ours, the total PSF is used in the deconvolution, so that the sampling problem is avoided only if each point source is exactly centered on a pixel. To increase the accuracy, the model can use a finer pixel grid than the data. However, in high S/N cases, the model pixels will generally need to be very small if high accuracy is aimed at (which is normally the case in high S/N observations...). As an example, let us recall that the positions derived from our new algorithm for the different images of the Cloverleaf gravitational lens (Fig. 2) are accurate to 0.01 arcsec, which is 1/35 of a data pixel. To achieve the same accuracy with the two channel Lucy algorithm would require each original pixel to be devided in ∼ 35 × 35 ∼ 1000 finer pixels. This would rapidly result in huge data frames and computationally intractable problems. Another weakness of the two-channel Lucy method is that the point source positions have to be supplied by the user, and cannot be adjusted by the algorithm. So, no astrometry can be performed and, moreover, in the case of high S/N data with many point sources, it might require an unreasonably large number of trials for the user to find a fairly good estimate of the source positions. Finally, let us note that the user has to choose arbitrarily not only the number of iterations of the algorithm, but also a scale length for the smoothing of the background component. These choices are generally made by looking at the results. This approach can obviously give nice-looking -- 10 -- results, but their scientific soundness may be questioned. On the contrary, the scale length for the smoothing of the background in our method is unambiguously fixed by the PSF r(~x) of the deconvolved image. A comparison of the two-channel Lucy method with our algorithm is illustrated below by an example which is not meant to provide a general comparison between the two methods, but only to illustrate some of the points in the preceding discussion. Figure 9 shows the deconvolution of simulated data containing three point sources superimposed on a background which varies rather fast. The peak S/N of 170 is quite reasonable for modern CCD detectors. In order to obtain a satisfactory result with the two-channel Lucy method, each original pixel was divided into 16 model pixels and the positions were adjusted iteratively by the user. In contrast, when running our algorithm, we kept the same pixel size as in the data (this is why the results of the two-channel method seem smoother in Fig. 9). The deduced background light distributions are compared in Fig. 10, which also shows the difference between these deduced backgrounds and the known solution, reconvolved to the same resolution of 2 pixels FWHM. It is immediately seen that the residuals are much less important in the case of our method (largest residual : 4.8 σ as compared to 14 σ with the Lucy method, mean variance : 1.8 σ2 instead of 15 σ2). The photometry of the point sources is also more accurate with our method : the mean deviation is 1%, as compared to 7% with the two-channel method. 6. Discussion We summarize here some of the reasons why classical deconvolution algorithms generally give rather disappointing results, and why our method allows to improve the situation. A major advantage of our method over traditional ones comes from the fact that the deconvolved image never violates the sampling theorem, so that the fastest image variations may be correctly represented, without the introduction of spurious rings, or Gibbs oscillations. An additional drawback of most traditional deconvolution algorithms lies in their smoothing recipe. For example, in the maximum entropy method, one assumes that the most probable image is a perfectly flat one. However, the most probable astronomical image is certainly not a flat one. It would rather look like a dark background with a number of sharp sources. Trying to smooth the sharp sources is undesirable, and results in poor performance. The smoothing function used in the classical maximum entropy method and most of its derivatives is, moreover, a global function, i.e., a function linking the value of the intensity in a particular pixel to the values in all other pixels, even very remote ones. Thus, the flux distribution in one part of the image will depend on what is happening in other remote parts (in astronomical images, this often corresponds to quite different parts of the Universe). This link is obviously not based on physical grounds, and is totally avoided by our smoothing function, which is purely local and linked to the PSF of the deconvolved image. Another weakness of the most popular of the classical methods (e.g. maximum entropy or Richardson-Lucy) is that the solution depends on the zero point level of the image : this is due -- 11 -- to the fact that the positivity constraint is essential for their success. Indeed, this positivity constraint is the main inhibitor of the ringing around point sources: by forbidding the negative lobes, it automatically reduces the positive ones since the mean level must be compatible with the observed data. Adding a constant to the image data results in a strong degradation of the performance of these algorithms (which then depend, e.g., on a precise subtraction of the sky level). On the contrary, our technique is completely independent of an additive constant, and it is reliable enough that the positivity constraint, although it can be used, is not necessary in most cases (it has not been used in any of the examples shown in this paper). As can be seen from the above examples and from the discussion, our new deconvolution technique is well adapted to the processing of astronomical images. It is however not restricted to that field of imaging and, in fact, should be useful in several other areas where an enhancement of the image resolution is desirable, or where different images of the same object could be optimally combined. References Busko, I.C., in The Restoration of HST Images and Spectra-II, R.J. Hanisch and R.L. White, eds., p. 279 (1994) Enard, D., Mar´echal, A. & Espiard, J. Rep. Prog. Phys. 59, 601 (1996) Heydari-Malayeri, M., Magain, P. & Remy, M. Astron. Astrophys. 222, 41 (1989) Hook, R.N. & Lucy, L., ST-ECF Newsletter 17, 10 (1992) Hook, R.N. & Lucy, L., in The Restoration of HST Images and Spectra-II, R.J. Hanisch and R.L. White, eds., p. 86 (1994) L´ena, P. Astrophysique. M´ethodes physiques de l'observation, 2nd edition, CNRS Editions, Paris, 1996 (english translation of the first edition : Observational Astrophysics, Springer-Verlag, Berlin, 1988) Lucy, L. Astron. J. 79, 745 (1974) Lucy, L. ST-ECF Newsletter 16, 6 (1991) Lucy, L., in The Restoration of HST Images and Spectra-II, R.J. Hanisch and R.L. White, eds., p. 79 (1994) Magain, P., Surdej, J., Swings, J.-P. et al. Nature 334, 325 (1988) Narayan, R. & Nityananda, R. Annu. Rev. Astron. Astrophys. 24, 127 (1986) Press, W. H., Flannery, B. P., Teukolsky, S. A. & Vetterling, W. T. Numerical Recipes (Cambridge University Press, 1989) Richardson, W. H. J. Opt. Soc. America 62, 55 (1972) Shannon, C. J. Proc. I. R. E. 37, 10 (1949) -- 12 -- Skilling, J. & Bryan, R. K. Mon. Not. Roy. Astron. Soc. 211, 111 (1984) Snyder, D. L., Miller, M. I., Thomas, L. J., & Politte, D. G. IEEE Transactions on Medical Imaging MI-6, 228 (1987) Acknowledgements This work has been supported by contracts ARC 94/99-178 "Action de Recherche Concert´ee de la Communaut´e Fran¸caise de Belgique", SC 005 "Service Center and Research Networks" and Pole d'Attraction Interuniversitaire P4/05 (SSTC, Belgium). -- 13 -- Fig. 1. -- Deconvolution of a simulated image of a star cluster partly superimposed on a background galaxy. Top left : true light distribution with 2 pixels FWHM resolution ; bottom left : observed image with 6 pixels FWHM and noise ; top middle : Wiener filter deconvolution of the observed image ; bottom middle : 50 iterations of the accelerated Richardson-Lucy algorithm ; top right : maximum entropy deconvolution ; bottom right : deconvolution with our new algorithm. Fig. 2. -- Deconvolution of a pre-discovery image of the Cloverleaf gravitational mirage obtained with the ESO/MPI 2.2m telescope at La Silla (Chile). Left : observed image with a FWHM resolution of 1.3 arcsec ; right : our deconvolution with improved sampling and a FWHM resolution of 0.5 arcsec. Fig. 3. -- Deconvolution of an image of the compact star cluster Sk 157 in the Small Magellanic Cloud. Left : image obtained with the ESO/MPI 2.2m telescope at La Silla (1.1 arcsec FWHM); right : deconvolution with our algorithm (0.26 arcsec FWHM). Fig. 4. -- Simultaneous deconvolution of simulated images. Top left : true light distribution with 2 pixels FWHM resolution ; top middle : image obtained with a space telescope ; top right : image obtained with a large ground-based telescope ; bottom left : sum of the two images ; bottom middle : simultaneous deconvolution with Lucy's algorithm ; bottom right : simultaneous deconvolution with our new algorithm. Fig. 5. -- Simultaneous deconvolution of 4 simulated adaptive-optics-like images. Top left : true light distribution with 2 pixels FWHM resolution ; middle and right : 4 images obtained with the same instrument but in varying atmospheric conditions ; bottom left : simultaneous deconvolution with our new algorithm. Fig. 6. -- Intensity ratio derived after deconvolution with the maximum entropy method for a pair of point sources with variable separation. The true intensity ratio is 0.1, the peak S/N ratio is 100 and the original resolution is 7 pixels FWHM. -- 14 -- Fig. 7. -- Photometric test performed on a synthetic field containing 200 stars with random positions and intensities, nearly all blended to various degrees (see text). The relative errors are plotted against the total intensity (the latter being on an arbitrary scale, corresponding to an integrated S/N varying from 10 to 400). Open symbols represent heavily blended stars (the distance to the nearest neighbour is smaller than the FWHM), filled symbols correspond to less blended objects. The dashed curves are the theoretical 3σ errors for isolated stars, taking into account the photon noise alone. Fig. 8. -- Astrometric test performed on the same crowded field as in Fig. 7. The total error in position (expressed in fractions of a pixel) is plotted versus the total intensity. Open symbols represent heavily blended stars (the distance to the nearest neighbour is smaller than the FWHM), filled symbols correspond to less blended objects. Fig. 9. -- Deconvolution of an image containing 3 points sources superimposed on a variable background. Top left : observed image; top right : deconvolution with 50 iterations of the accelerated Richardson-Lucy algorithm; bottom left : deconvolution with 1000 iterations of the two-channel Lucy method; bottom right : deconvolution with our algorithm. Fig. 10. -- Comparison of the background light distributions deduced from the deconvolution of the image in Fig. 9, using the two-channel Lucy method (top left) and our method (top right). The bottom panels show the square of the difference between the deduced background (reconvolved to the same 2 pixels resolution when necessary) and the exact solution, with the two-channel Lucy method (left) and with our algorithm (right). This figure "fig1.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/astro-ph/9704059v2 This figure "fig2.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/astro-ph/9704059v2 This figure "fig3.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/astro-ph/9704059v2 This figure "fig4.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/astro-ph/9704059v2 This figure "fig5.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/astro-ph/9704059v2 This figure "fig9.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/astro-ph/9704059v2 This figure "fig10.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/astro-ph/9704059v2
astro-ph/0609188
2
0609
2006-12-19T17:36:43
Can extragalactic foregrounds explain the large-angle CMB anomalies?
[ "astro-ph" ]
We address the effect of an extended local foreground on the low-l anomalies found in the CMB. Recent X-ray catalogues point us to the existence of very massive superstructures at the 100 h^(-1) Mpc scale that contribute significantly to the dipole velocity profile. Being highly non-linear, these structures provide us a natural candidate to leave an imprint on the CMB sky via a local Rees-Sciama effect. We show that the Rees-Sciama effect of local foregrounds can induce CMB anisotropy of DeltaT/T ~ 10^(-5) and we analyse its impact on multipole power as well as the induced phase pattern on largest angular scales.
astro-ph
astro-ph
October 18, 2018 18:20 WSPC - Proceedings Trim Size: 9.75in x 6.5in main 6 0 0 2 c e D 9 1 2 v 8 8 1 9 0 6 0 / h p - o r t s a : v i X r a CAN EXTRAGALACTIC FOREGROUNDS EXPLAIN THE LARGE -- ANGLE CMB ANOMALIES? Aleksandar Raki´c 1, Syksy Rasanen2 and Dominik J. Schwarz1 1 Fakultat fur Physik, Universitat Bielefeld, Postfach 100131, D-33501 Bielefeld Germany 2 CERN Physics Department Theory Unit, CH -- 1211 Geneva 23, Switzerland email: rakic at physik dot uni -- bielefeld dot de, syksy dot rasanen at iki dot fi, dschwarz at physik dot uni -- bielefeld dot de We address the effect of an extended local foreground on the low -- ℓ anomalies found in the CMB. Recent X -- ray catalogues point us to the existence of very massive superstructures at the 100 h−1Mpc scale that contribute significantly to the dipole velocity profile. Being highly non -- linear, these structures provide us a natural candidate to leave an imprint on the CMB sky via a local Rees -- Sciama effect. We show that the Rees -- Sciama effect of local foregrounds can induce CMB anisotropy of ∆T /T ∼ 10−5 and we analyse its impact on multipole power as well as the induced phase pattern on largest angular scales. 1. Motivation and Overview At largest angular scales which correspond to small multipole moments ℓ there exist puzzling features in the Cosmic Microwave Background (CMB). The near vanishing of the two -- point angular correlation function in all wavebands for angular scales between 60◦ and 170◦ is one of the longest known anomalies, already detected in the data of the Cosmic Background Explorer's Differential Radiometer (COBE -- DMR). It has been confirmed and persists in the three -- year Wilkinson Microwave Anisotropy Probe data [WMAP(3yr)].1,2 Among the two -- point angular correlation functions it has been shown that none of the almost vanishing cut -- sky wavebands matches the full sky and again neither one of these is in accordance with the best fit Λ cold dark matter (ΛCDM) model.3 The disagreement turned out to be even more distinctive in the WMAP(3yr) data than in WMAP(1yr) and is unexpected at 99.97% C.L. for the updated Internal Linear Combination map [ILC(3yr)].3 Besides the lack of power, there are a number of remarkable anomalies regarding the phase relationships of the quadrupole and octopole within the WMAP data.4 -- 6 In order to be able to make distinct statements with respect to a phase analysis of multipoles we make use of the multipole vectors formalism.7 Looking at quadrupole plus octopole vectors from WMAP(3yr) the alignment with the equinox (EQX) and with the ecliptic is found to be unlikely at 99.8% C.L. and 96% C.L. respectively.3 The correlation with the dipole direction and with the galactic plane is found to be odd at 99.7% C.L. and 99% C.L. respectively. Moreover from the combined full sky map of ℓ = 2 + 3 one infers that the octopole is quite planar and that the ecliptic strongly follows a zero line of the map, leaving the two strongest extrema in the southern hemisphere and the two weakest in the northern hemisphere. Some of these effects are statistically dependent, e.g. given the observed quadrupole -- octopole alignment, the significance of alignment with the galactic plane is reduced to unremarkable 88% C.L. These findings support the conclusion that either the Universe as seen by WMAP 1 October 18, 2018 18:20 WSPC - Proceedings Trim Size: 9.75in x 6.5in main 2 is not statistically isotropic on largest scales, or that the observed features are due to unexpected foregrounds, hidden systematics or new physics challenging the standard cosmological model. Diverse attempts for explanation can be found in the litera- ture: considering anisotropic or inhomogeneous models [Bianchi family, Lemaıtre -- Tolman -- Bondi (LTB) models],8 -- 13 Solar system foreground,14 lensing of the CMB15 and moving foregrounds,16 Sunyaev -- Zel'dovich (SZ) effect17,18 and Rees-Sciama (RS) effect,12,19 considering a non -- trivial topology of the Universe,20,21 considering modifications and refinements of the standard simplest scenario of inflation,22 -- 28 considering possible phenomenology of loop quantum gravity.29,30 In this talk we update and expand our previous work12 in the light of the WMAP(3yr) data release. 2. Local Structures and Rees-Sciama Effect Recent X-ray catalogues of our neighborhood show that a major contribution to the dipole velocity profile originates from the Shapley Supercluster (SSC) and other density concentrations at a distance of of around 130 -- 180 h−1 Mpc.31 -- 34 The SSC is a massive concentration centered around the object A3558 with a density contrast of δ ≃ 5 over the inner 30 h−1Mpc region.35 We will show that the CMB displays correlations between the dipole and higher multipoles after passing through non -- linear structures, due to the RS effect.36 The physics of the RS effect is that in the non -- linear regime of structure formation, the gravitational potential changes with time, so photons climb out of a slightly different potential well than the one they fell into. Following ref.37 the CMB anisotropy pro- duced by a spherical superstructure is estimated by the integral of the gravitational potential perturbation φ ≃ δM/d along the path of the photon: ∆T (θ, ϕ)/T ≃ φ vc , where d is the physical size of the structure and δM is the mass excess. Here we assumed a structure collapsing at velocity vc and let the evolution time of the structure tc be the matter crossing time d/vc (using c ≡ 1 ≡ G). We estimate the typical collapse velocity from the energy balance condition v2 c ≃ φ and get: ∆T (θ, ϕ)/T ∼ φ3/2 ∼ (δM/d)3/2. We model the non -- linear structure by a spher- ically symmetric LTB model embedded in a flat (Ω = 1) Friedmann -- Robertson -- Walker Universe. Substituting the expression for the mass excess within this model we arrive at:37 ∆T (θ, ϕ) T ∼(cid:18) δρ ρ (cid:19)3/2(cid:18) d t(cid:19)3 , (1) where t is the cosmic time at which the CMB photons crossed the structure. Inserting the characteristics of the SSC it follows that a CMB anisotropy of 10−5 due to a local RS effect is reasonable. For simplicity we picture the local Universe as a spherically symmetric density distribution, with the Local Group (LG) falling towards the core of the overdensity at the centre. The line between our location and the centre defines a preferred direction z, which in the present case corresponds to the direction of the dipole. This setup exhibits rotational symmetry w.r.t. the axis z (neglecting transverse components of our motion). Consequently, only zonal October 18, 2018 18:20 WSPC - Proceedings Trim Size: 9.75in x 6.5in main axial= 0 µK [WMAP(1yr) best-fit ΛCDM] a20 20 µK 40 µK 70 µK WMAP(1yr) WMAP(3yr) 7 6 5 4 3 2 1 d o o h i l e k L i 0 0 0.05 0.1 0.15 C2 in [0.1 mK]2 0.2 0.25 0.3 d o o h i l e k L i 16 14 12 10 8 6 4 2 0 3 axial= a30 0 µK [WMAP(1yr) best-fit ΛCDM] 20 µK 40 µK 70 µK WMAP(1yr) WMAP(3yr) 0 0.05 0.1 0.15 C3 in [0.1 mK]2 0.2 0.25 0.3 Fig. 1. Likelihood of quadrupole and octopole power for increased axial contributions. Vertical lines denote experimental data: WMAP(1yr) cut -- sky and WMAP(3yr) maximum likelihood es- timate. Considering the quadrupole adding any multipole power was excluded at > 99% C.L. w.r.t WMAP(1yr) but it is possible to add up to 60µK within the same exclusion level w.r.t. the WMAP(3yr) value. The octopole is more resistant against axial contaminations as it is possible to add a whole 100µK before reaching the same exclusion level w.r.t the updated WMAP data. harmonics (m = 0 in the z-frame) are generated. Note that any other effect with axial symmetry would also induce anisotropy only in the zonal harmonics. 3. Multipole Analysis We study how maps of the CMB are affected by the anisotropy induced by additional axisymmetric contributions aaxial added to the quadrupole and octopole by using Monte Carlo (MC) methods. As predicted by the simplest inflationary models, we assume that the aℓm are fully characterised only by angular power, for which we use the values from the best fit ΛCDM temperature spectrum to the WMAP data.2 We produced 105 MC realisations of ℓ = 2 and ℓ = 3 for the statistical analysis. ℓ0 The angular power spectrum is estimated by Cℓ = 1/(2ℓ + 1)Pm aℓm2. In fig. 1 we show how the histograms for the quadrupole and octopole power compare with the measured values from WMAP(1yr,3yr). Considering the WMAP(1yr) cut -- sky, adding any power to the quadrupole was already excluded at > 99% C.L. whereas the WMAP(3yr) data allows for adding up to aaxial 20 = 60µK in order to reach the same exclusion level. The octopole is quite robust against axial contaminations as it lies better on the fit: in order to reach the same exclusion level of > 99% C.L. it is necessary to add aaxial 30 = 80µK w.r.t. the WMAP(1yr) cut -- sky and a whole aaxial 30 = 100µK w.r.t. the WMAP(3yr) value. Considering only the WMAP(3yr) maximum likelihood estimate and increasing the effect of local structures up to aaxial ℓ0 = 70µK leads to an exclusion of 99.5% C.L. for C2 and 92.9% C.L. for C3. ℓ0 The next question is what kind of phase pattern the contribution aaxial will induce on the CMB sky. Using the multipole vector formalism7 a (temperature) multipole on a sphere can be alternatively decomposed as: Tℓ = ℓ Xm=−ℓ aℓmYℓm(θ, ϕ) = A(ℓ)" ℓ Yi=1(cid:16)v (ℓ,i) · e(θ, ϕ)(cid:17) − Lℓ(θ, ϕ)# , (2) October 18, 2018 18:20 WSPC - Proceedings Trim Size: 9.75in x 6.5in main 4 where e(θ, ϕ) = (sin θ cos ϕ, sin θ sin ϕ, cos θ) is a radial unit vector. With the de- composition (2) it is possible to obtain an unique factorisation of a multipole into a (ℓ,i) that contain scalar part A(ℓ) which measures its total power and ℓ unit vectors v all the directional information. The signs of the multipole vectors can be absorbed into the scalar quantity A(ℓ), and are thus unphysical. Introducing the ℓ(ℓ − 1)/2 oriented areas n(ℓ;i,j) ≡ v (ℓ,j) we are ready to define a statistic in order to probe alignment of the normals n(ℓ;i,j) with a given physical direction x: (ℓ,i) × v (ℓ,i) × v (ℓ,j)/v Snx ≡ . (3) 1 4 Xℓ=2,3Xi<j(cid:12)(cid:12)(cid:12) n (ℓ;i,j) · x(cid:12)(cid:12)(cid:12) We test for alignment with three natural directions x: the north ecliptic pole (NEP), EQX and the north galactic pole (NGP). The results of the correlation analysis are shown in fig. 2: in the first row the preferred direction z coincides with the direction of local motion, the dipole.38 Here the anomaly becomes worse when increasing the amplitude of the axial contribution. But for x = NEP the exclusion becomes somewhat milder; e.g. aaxial ℓ0 = 40µK leads to an exclusion of 99.2% C.L. for ILC(1yr) but only 98.2% C.L. for the updated ILC map. Finding an alignment with the EQX though is strongly excluded at > 99.2%C.L. even with a vanishing axial contribution for both one -- and three -- year data. In the second row of fig. 2 we let the preferred direction point to the NEP as a complementary test. Here the probability to find an ecliptic alignment becomes dramatically increased: with a aaxial ℓ0 = 70µK it is 17% and 10% for the ILC(3yr) and ILC(1yr) values respectively. Regarding the three -- year data the probability for finding an EQX alignment increases from 1% to 3% for aaxial ℓ0 = 70µK. The alignment with the NGP remains quite stable for both tested directions of z. 4. Conclusion Recent astrophysical data cataloguing our neighborhood in the X -- ray band point us to the existence of massive non -- linear structures like the SSC at distances ∼ 100 h−1Mpc. Besides its significant contribution to the dipole velocity profile such a structure is able to induce anisotropies ∼ 10−5 via its RS effect. Regarding CMB modes, the spherical symmetry (LTB) which we use to approximate the local superstructure reduces to an axial symmetry along the line connecting our position and the centre of the superstructure where we locate the SSC. We produced statisti- cally isotropic and gaussian MC maps of the CMB and computed their S -- statistics (3) for alignment with generic astrophysical directions like the NEP, EQX and NGP. The additional zonal harmonics have been added with increasing strength (see ref.39 for full -- sky maps). When gauging the preferred axis to the direction of local mo- tion (WMAP dipole) the consistency of the data with theory becomes even worse, albeit with less significance w.r.t. WMAP(3yr). On the other hand an orthogonally directed (Solar system) effect would be more consistent with the three -- year data. October 18, 2018 18:20 WSPC - Proceedings Trim Size: 9.75in x 6.5in main 5 Considering extended local foregrounds Abramo et al.17 recently proposed that a cold spot in the direction of the local Supercluster could account for the cross alignments of quadrupole and octopole. The cold spot would be realised by the SZ effect of CMB photons scattering of the hot intracluster gas. On the other hand Inoue and Silk19 suggest a certain geometrical pattern of two identical voids to account for the cross alignment as well as for the octopole planarity via the RS effect. Each of the latter approaches alone is not fully satisfactory. Nevertheless a combined approach enfolding the RS effect as well as the SZ effect from extended foregrounds seems promising for the future. Moreover, since the local RS effect can contribute up to 10−5 to the temperature anisotropies on large angular scales, a detailed study is important for cross -- correlating CMB data (including upcoming Planck data) with astrophysical observations on the local large -- scale structure. Acknowledgment It is a pleasure to thank the organisers of the 11th Marcel Grossmann meeting for their effort and the opportunity to speak. We acknowledge the use of the Legacy Archive for Microwave Background Data Analysis (LAMBDA) provided by the NASA Office of Space Science. The work of AR is supported by the DFG grant GRK 881. References 1. G. Hinshaw et al., astro-ph/0603451. 2. WMAP data products at http://lambda.gsfc.nasa.gov/ 3. C. Copi, D. Huterer, D. J. Schwarz and G. Starkman, astro-ph/0605135. 4. D. J. Schwarz, G. D. Starkman, D. Huterer and C. J. Copi, PRL 93, 221301 (2004). d o o h i l e k L i 4 3.5 3 2.5 2 1.5 1 0.5 axis in direction of WMAP dipole axial= al0 0 µK 40 µK 70 µK ) r y 1 ( C L I ) r y 3 ( C L I axial= 0 µK al0 40 µK 70 µK axis in direction of WMAP dipole ) r y 3 ( C L I ) r y 1 ( C L I 4 3.5 3 2.5 2 1.5 1 0.5 axis in direction of WMAP dipole axial= al0 0 µK 40 µK 70 µK ) r y 3 ( C L I ) r y 1 ( C L I 4 3.5 3 2.5 2 1.5 1 0.5 0 0 0.2 0.4 SnNEP 0.6 0.8 1 0 0 0.2 0.4 SnEQX 0.6 0.8 1 0 0 0.2 0.4 0.6 0.8 1 SnNGP 4 3.5 3 2.5 2 1.5 1 0.5 d o o h i l e k L i ) r y 1 ( C L I ) r y 3 ( C L I axis in direction of north ecliptic pole axial= al0 0 µK 40 µK 70 µK axis in direction of north ecliptic pole axial= al0 0 µK 40 µK 70 µK ) r y 3 ( C L I ) r y 1 ( C L I 4 3.5 3 2.5 2 1.5 1 0.5 axis in direction of north ecliptic pole axial= al0 0 µK 40 µK 70 µK ) r y 3 ( C L I ) r y 1 ( C L I 4 3.5 3 2.5 2 1.5 1 0.5 0 0 0.2 0.4 SnNEP 0.6 0.8 1 0 0 0.2 0.4 SnEQX 0.6 0.8 1 0 0 0.2 0.4 0.6 0.8 1 SnNGP Fig. 2. WMAP one- and three -- year ILC maps compared to the alignment (3) of quadrupole and octopole normals with physical directions (NEP, EQX, NGP in columns) for two orthogonal realisations of the preferred direction z (WMAP dipole, NEP in rows). The bold histograms represent statistically isotropic and gaussian skies. Increasing the axial contribution makes the anomalies worse for z = WMAP dipole, but with the exclusions being less significant for the ILC(3yr) than for the ILC(1yr). At the same time a Solar system effect is preferred by the data. October 18, 2018 18:20 WSPC - Proceedings Trim Size: 9.75in x 6.5in main 6 5. H. K. Eriksen, F. K. Hansen, A. J. Banday, K. M. G´orski and P. B. Lilje, ApJ 605, 14 (2004); (Erratum) 609, 1198 (2004). 6. A. de Oliveira -- Costa, M. Tegmark, M. Zaldarriaga and M. Hamilton, Phys. Rev. D 69, 063516 (2004);A. de Oliveira -- Costa and M. Tegmark, astro-ph/0603369. 7. C. J. Copi, D. Huterer and G. D. Starkman, Phys. Rev. D 70, 043515 (2004). 8. T. Ghosh, A. Hajian and T. Souradeep, astro-ph/0604279. 9. T. R. Jaffe, A. J. Banday, H. K. Eriksen, K. M. Gorski and F. K. Hansen, astro-ph/0606046 and references therein. 10. H. Alnes and M. Amarzguioui, astro-ph/0607334. 11. J. W. Moffat JCAP 0510, 012 (2005). 12. A. Raki´c, S. Rasanen and D. J. Schwarz, MNRAS 369, L27 (2006). 13. K. Tomita Phys. Rev. D 72, 043526 (2005), Phys. Rev. D 72 103506; (Erratum) D 73 029901. 14. P. C. Frisch, ApJ 632, L143 (2005). 15. C. Vale, astro-ph/0509039. 16. A. Cooray and N. Seto, JCAP 0512, 004 (2005). 17. L. R. Abramo and L. Sodr´e Jr., astro-ph/0312124; L. R. Abramo, L. Sodr´e Jr. and C. A. Wuensche, astro-ph/0605269. 18. F. K. Hansen, E. Branchini, P. Mazzotta, P. Cabella and K. Dolag, MNRAS 361, 753 (2005). 19. K. T. Inoue and J. Silk, ApJ 648, 23 (2006); K. T. Inoue and J. Silk, astro-ph/0612347. 20. A. Riazuelo, J. Weeks, J. P. Uzan, R. Lehoucq and J. P. Luminet, Phys. Rev. D 69, 103518 (2004); J. P. Luminet, J. Weeks, A. Riazuelo, R. Lehoucq and J. P. Uzan, Nature 425, 593 (2003). 21. J. Shapiro Key, N. J. Cornish, D. N. Spergel and G. D. Starkman, astro-ph/0604616; N. J. Cornish, D. N. Spergel, G. D. Starkman and E. Komatsu, Phys. Rev. Lett. 92, 201302 (2004). 22. D. Boyanovsky, H. J. de Vega and N. G. Sanchez, astro-ph/0607508; astro-ph/0607487. 23. L. Campanelli, P. Cea and L. Tedesco, astro-ph/0606266. 24. C. R. Contaldi, M. Peloso, L. Kofman and A. Linde, JCAP 0307, 002 (2003). 25. F. Ferrer, S. Rasanen and J. Valiviita, JCAP 0410, 010 (2004). 26. C. Gordon and W. Hu, Phys. Rev. D 70, 083003 (2004). 27. A. E. Gumruk¸cuoglu, C. R. Contaldi and M. Peloso, astro-ph/0608405. 28. C. -- H. Wu, K. -- W. Ng, W. Lee, D. -- S. Lee and Y. -- Y. Charng, astro-ph/0604292. 29. S. Hofmann and O. Winkler, gr-qc/0411124. 30. S. Tsujikawa, P. Singh and R. Maartens, CQG 21, 5767 (2004). 31. D. D. Kocevski, C. R. Mullis, H. Ebeling, ApJ 608, 721 (2004). 32. D. D. Kocevski, H. Ebeling, ApJ 645, 1043 (2006). 33. M. J. Hudson, R. J. Smith, J. R. Lucey, E. Branchini, MNRAS 352, 61 (2004). 34. J. Lucey, D. Radburn -- Smith, M. Hudson, astro-ph/0412329. 35. D. Proust et al., A&A 447, 133 (2006). 36. M. J. Rees, D. W. Sciama, Nature 217, 511 (1968). 37. M. Panek, ApJ 388, 225 (1992). 38. C. L. Bennett et al., ApJS 148, 1 (2003). 39. Full -- sky maps at http://www.physik.uni-bielefeld.de/cosmology/rs.html
astro-ph/0503044
2
0503
2006-02-08T17:06:38
Effect of entrainment on stress and pulsar glitches in neutron star crust
[ "astro-ph" ]
The build up of the stress whose relaxation is presumed to account for pulsar frequency glitches can be attributed to various mechanisms, of which the most efficient involve differential rotation of the neutron superfluid in the inner layers of the (magnetically braked) solid crust of a rotating neutron star. In such a case it is usually supposed that the stress is attributable to pinning of superfluid vortices to crust nuclei, but it has been suggested that, even if the pinning effect is too weak, a comparably large stress might still arise just from the deficit of centrifugal buoyancy in the slowed down crust. The present work is a re-examination that investigates the way such processes may be affected by considerations that were overlooked in the previous work -- notably uncertainties about the ``effective'' masses that have to be attributed to the ``free'' superfluid neutrons to allow for their entrainment by the ionic crust material. Though restricted to a Newtonian formulation, this analysis distinguishes more carefully than has been usual between true velocities, which are contravariantly vectorial, and so called ``superfluid velocities'' that are proportional to momenta, which are essentially covectorial, a technicality that is important when more than one independent current is involved. The results include a Proudman type theorem to the effect that the superfluid angular velocity must be constant on slightly deformed Taylor cylinders in the force free case, and it is shown how to construct a pair of integral constants of the motion that determine the solution for the pinned case assuming beta equilibrium.
astro-ph
astro-ph
Effect of entrainment on stress and pulsar glitches in stratified neutron star crust Nicolas Chamel⋆ †, Brandon Carter⋆ ⋆ LuTh, Observatoire de Paris, 92195 Meudon, France † Centrum Astronomiczne im M. Kopernika, Warszawa, Polska Extended version, 10 December 2005 Abstract 1 Introduction The build up of the stress whose relaxation is presumed to account for pulsar frequency glitches can be attributed to various mechanisms, of which the most efficient involve differential rotation of the neutron superfluid in the inner layers of the (magnetically braked) solid crust of a rotat- ing neutron star. In such a case it is usually supposed that the stress is attributable to pinning of superfluid vor- tices to crust nuclei, but it has been suggested that, even if the pinning effect is too weak, a comparably large stress might still arise just from the deficit of centrifugal buoy- ancy in the slowed down crust. The present work is a re-examination that investigates the way such processes may be affected by considerations that were overlooked in the previous work -- notably uncertainties about the "effec- tive" masses that have to be attributed to the "free" super- fluid neutrons to allow for their entrainment by the ionic crust material. Though restricted to a Newtonian formu- lation, this analysis distinguishes more carefully than has been usual between true velocities, which are contravari- antly vectorial, and so called "superfluid velocities" that are proportional to momenta, which are essentially cov- ectorial, a technicality that is important when more than one independent current is involved. The results include a Proudman type theorem to the effect that the superfluid angular velocity must be constant on slightly deformed Taylor cylinders in the force free case, and it is shown how to construct a pair of integral constants of the motion that determine the solution for the pinned case assuming beta equilibrium. 1 This work is intended as a comparison of the qualitative mathematical consequences of hypotheses of various alter- native kinds (particularly concerning vortex pinning and chemical equilibrium) that have been invoked in the con- struction of differentially rotating neutron star models of the type developed to acccount for phenomena observed in pulsars. It is hoped that this qualitative study will be helpful for the preparation of quantitative models designed to im- prove on those of the simple kind constructed by Prix et al [39], for which (as for the analogously constructed mod- els [40] in a relativistic framework) separately conserved proton and superfluid neutron currents were able to de- viate from Proudman - Taylor type corotation illustrated on figure 1, without giving rise to the (glitch producing) stresses that are actually anticipated. In earlier work [38] such stresses were obtained by including allowance for the effect of vortex pinning, but it has since been remarked [16] that deviations from corotation would give rise to stresses in any case (with or without pinning) when allowance is made for the consideration that the proton and neutron numbers will not be separately conserved but will undergo transfusion (by beta processes) so as to achieve chemical equilibrium. The work of Prix [38] and of Carter et al [16] used a simplified treatment that neglected the mechanism of rel- ative entrainment between the differentially rotating con- stituents, an effect which had already been considered in earlier work (both in a Newtonian framework [34] and also [30] in a relativistic treatment) but which had been consid- ered to be only of secondary relevance as a minor quan- titative correction. The effect of entrainment (but not transfusion) in the liquid core was allowed for in the more recent work of Comer & Andersson [18] and of Prix et al. [39], and it has since been recognised [17] to be particu- larly important in the crust layers that are the primary concern of the present analysis, which will consider the (glitch producing) stresses that can be expected to arise from the combined effect of entrainment, pinning, trans- fusion, and also nuclear stratification. Explanations of the glitches of the pulsar period P = 2π/Ω observed in rotating neutron stars may be broadly classified into two main categories. In what may be de- scribed as the "deformation" category, the discontinuous spin up δΩ is attributed to a sudden change of the geomet- ric configuration of the matter distribution [23], whereby the relevant moment of inertia is decreased so that -- for a given angular momentum -- the angular velocity Ω must increase. While conceivably sufficient to account for cases such as the glitches of order δΩ/Ω ≃ 10−8 observed in the Crab pulsar, such a deformation mechanism is inadequate for explaining the frequent occurrence of the larger gliches δΩ/Ω ≃ 10−6 observed in Vela like pulsars, not to mention the enormous glitch δΩ/Ω ≃ 1.6 × 10−5 recently observed in PSR J1806-2125 by Hobbs et al. [24]. To account for the frequent occurence of such large glitches, it has long been generally recognised [8] to be necessary to invoke a mechanism belonging to what may be referred to as the "transfer" category, involving inter- action between two dynamically distinct constituents of the star. In such a case the discontinuous increase in the observed angular velocity Ω is attributed to transfer of angular momentum from a more rapidly rotating internal constituent to a more slowly rotating constituent that is directly coupled to the outer magnetosphere whose pulsa- tions are directly observed. However it is not so clear what actually happens during the discontinous process involved. It has been suggested [42] that the stress due to pinning of neutron superfluid vortices to the ionic crust material may build up elastic stress to the point where the solid stuc- ture itself breaks down (as in a terrestrial earthquake). In another much studied scenario [2, 37] it is postulated that before this can happen there will be a crisis of a different kind whereby a threshold is reached at which the vortices in the crust become collectively unpinned (whereas in the fluid core it is expected that vortices in the core will be subject to a strong dissipative drag, which will simulates pinning but will not be limited by any threshold). Another possibility [27, 26, 21, 22] is that pinning in the crust may be too weak to be effective, but it has been pointed out [16] that differential rotation might in principle by itself engender sufficient elastic stress to cause a breakdown of the solid structure even in the absence of pinning. The main purpose of the present work is not to draw any specific conclusions about the relative plausibility of these various postulates in particular cases, but to consider, gen- erally, how such alternatives will affect the predicted out- come, and how estimates of relevant quantities such as subcomponent moments inertia and stresses may need to be modified to take account of some effects that tended to be neglected in preceding discussions, including stratifica- tion and particularly the entrainment between the ionic crust material and the "free" superfluid neutrons, which thereby aquire "effective masses" that can be defined in various ways, but that in any case are likely[13, 14, 15] to be large enough to make an important difference. In order to carry out such a revision, and as a secondary purpose in its own right, this work employs and develops a Newtonian formulation of the covariant kind [11, 12] whose use would be taken for granted in a relativistic treatment, in which care is taken to distinguish between true veloc- ities, which are contravariantly vectorial, and so called "superfluid velocities" that are proportional to momenta, which are essentially covectorial. Although deeply rooted in the well known principles of Lagrangian and Hamiltonian mechanics, the fundamental distinction between velocities and their canonical conju- gates, namely momenta, has commonly been obscured in the context of superfluidity as traditionally formulated [34] in a Newtonian framework, for which it has long been cus- tomary to use the term "velocity" indiscriminately for any- thing having the physical dimensions thereoff, regardless of its actual mathematical role. The distinction is rela- tively unimportant for a single constituent perfect fluid or superfluid such as that of ordinary Helium at zero temper- ature, or even for a two contituent fluid in which entrain- ment is absent [16], but it becomes important when there are two or more independently moving fluids with entrain- ment, which means that the momenta need no longer be aligned with their conjugate velocities. As well as simplifying the algebra in the work presented below, the introduction of the canonically covariant for- mulation also has the advantage of providing a helpful step towards the development of a fully General relativis- tic treatment, such as will ultimately be necessary for ac- curacy. Relativistic effects will be particularly important for the core, in which however, compared with the crust, the effect of entrainment is expected to be much more moderate [44, 19] at least in the outer part for which the basic physics is reasonably well understood, unlike the in- ner core for which many exotic possibilities have been en- visages, such that of hyperon crystallisation [36], or of a LOFF phase in a color superconducting quark condensate [1]. 2 would appear that corrections from general relativity will be much less important (at most a few tens of per cent) than those due to the entrainment. A recent program of work [13, 14, 15] (using a microphysical analysis combin- ing methods from nuclear and solid state physics) on the relative motion of unbound neutrons in the inner crust lay- ers (between the neutron drip density about 1011 g/cm3 and nuclear saturation density about 1014 g/cm3) has pro- vided results [17] indicating that the effect of entrainment is likely to be very much stronger than had previously been supposed. As in the preceding work [16], we shall therefore pro- ceed here within a non relativistic Newtonian treatment, but instead of using a decoupled perfect fluid description we shall now include allowance for the effect of entrain- ment, using a non-dissipative two constituent description of the generic kind already developed for use in the liquid core [34, 18, 39] where the entrainment corrections are rel- atively moderate, though still potentially important as a cause for two-stream instability [6] if the relative velocity becomes too large. For the reasons outlined in the introduction, it is both instructive and convenient (even though our treatment is non-relativistic) to use a canonical formalism of the 4- dimensionally covariant kind recently developed [11, 12] for the purpose of facilitating the exploitation of the varia- tional formulation that is applicable to the non-dissipative limit. In this limit there is no resistance to relative mo- tion, so that the only direct coupling between the two constituents is via the entrainment effect [7], whose micro- scopic origin is the Bragg scattering of dripped neutrons by crustal nuclei [13] which is the analog in the nuclear con- text of conduction electron scattering in ordinary solids. It is also convenient to use a chemical constituent label X that runs over two values, namely X = f for the superfluid constituent and X = c for the crust constituent. Each constituent will have a baryon current density 4-vector having the form n ν X = nX u ν X (1) in which the scalars nX are the respective number densities of free baryons (only neutrons) and of baryons that are effectively confined within the atomic nuclei (namely all the protons, together with a roughly equal but somewhat larger number of confined neutrons, whose precise specifi- cation needs further clarification as discussed below) while the vectors u ν X are the corresponding 4-velocities, whose components are labelled by a Greek index ν running over the values 0, 1, 2, 3. The conservation of the total baryon number current n ν B = n ν f + n ν c , (2) is concisely expressible in this 4-dimensional notation as ∇νn ν B = 0 . (3) Figure 1: Intersection of rotating neutron star model with a Taylor cylinder, over which the superfluid angular veloc- ity must be uniform when the conditions for the Proudman theorem are applicable. 2 Two fluid model for stratified neutron star crust 2.1 The Lagrangian master function f and n ν The previous treatment referred to above [16] used a non relativistic Newtonian description in terms of free neutron and confined baryon 4-currents n ν c respectively representing a neutron superfluid, and a normal crust com- ponent consisting of ionic nuclei characterised by a con- fined baryon number Ac and a positive charge number Z (in a degenerate electron background) that were treated as decoupled but interpenetrating barotropic perfect fluids interacting only via vortex pinning and long range gravita- tional forces, with the solid rigidity of the crust component taken into account by the inclusion of an extra stress force density. One of the improvements introduced here is to drop the barotropy restriction for the ionic constituent by allowing its behavior to depend on the nuclear mass number Ac (which will tend to be larger in the deeper layers) thus allowing for the possibility of a stratification effect that may be important for stabilisation. Another improvement that would be needed for a more accurate analysis is of course the use of a general rela- tivistic treatment, of the kind that has long been com- monly used with a simple perfect fluid description, and that has recently been developed for use with a two con- stituent fluid description of the kind needed fore treating the stellar core [30, 5]. However for the semilocal modeli- sation needed for the treatment of differential rotation in the crust layers with which we shall be concerned here, it 3 X = v i X = 1, u i In Aristotelian coordinates (representing the usual kind of 3+1 space time decomposition with respect to the rest frame of the star) the 4-velocity components will have the form u 0 X are ordinary 3-velocity components, using a Latin index i = 1, 2, 3. This means that time component of the current density is simply equal X = nX , to the corresponding particle number density n0 while the space components are the those of the current density 3-vector ni X where the v i X . c and n i In addition to the six independent vectorial components provided by the space vectors n i f and the pair of scalars provided by the "confined" and "free" baryon num- ber densities nc and nf , our model will contain a nineth independent variable, namely the scalar provided by the nuclear mass number Ac, or equivalently by the number density nI of nuclei, whose number current will be given by I = nIu ν n ν c , nI = nc/Ac . (4) We shall take the mass per baryon to be the same (con- sidering the electron mass and the proton neutron mass difference to be negligible compared with the ordinary proton mass) with fixed value mc = mf = m, so that the corresponding mass density contributions will simply be given by ρX = m nX . The total mass density, as given explicitly by ρ = m nB = ρf + ρc , (5) will be important as the Poissonian source of the Newto- nian gravitational potential scalar, φ. Within this New- tonian background field, the total force balance equation will take the form ∇µT µ ν = f ν − ρ∇νφ , (6) in which f ν is the non gravitational external 4-force den- sity -- if any -- that may be acting on the system, and T µ ν is the stress momentum energy density tensor as constructed in terms of the 4-dimensional formalism recapitulated be- low. The dynamics of system will be governed -- in the man- ner described in our preceding work [11, 12] -- by a La- grangian master function Λ = Λmat − φ ρ , Λmat = Λkin + Λint (7) in which Λkin is the ordinary Galilean frame dependent kinetic energy density, namely Λkin = 1 2 (cid:0)ρf v 2 f + ρcv 2 c(cid:1) , (8) and Λint is some appropriately specified internal action density contribution that must be Galilean frame indepen- dent, which means that it can depend only on the scalar 4 number densities nI, nf , nc, and on the velocity difference [vi] = v i f − v i c (9) so that in terms of an "entrainment density" function ¯ρfc and a set of chemical potentials χf , χc, χI its generic vari- ation will take the form δ Λint = ¯ρfc[vi] δ[vi] − χf δnf − χcδnc − χIδnI , (10) which gives the corresponding formula δΛmat = nf µf iδv i f +ncµc iδv i c − µf δnf − µcδnc − µIδnI , (11) in which the (Galilean frame dependent) material energy coefficients are given in terms of the corresponding (frame independent) chemical potentials by µf = χf − 1 2 m v 2 f , µc = χc − 1 2 m v 2 c , µI = χI , (12) and the associated space momenta are given by µf i = m vf i + ¯ρfc nf [vi] , µc i = m vc i − ¯ρfc nc [vi] . (13) In the low relative velocity regimes that are relevant it will always be possible to take the internal energy contri- bution to have the quadratic form Λint = 1 2 ¯ρfc[v]2 − U ins , (14) with ¯ρfc and the static contribution U ins specified by ap- propriate equations of state as functions only of the scalar densities nI, nc, nf . It is to be remarked that to translate our notation to that used by Prix et al. [39] we would need to make the conversions χf 7→ µf , χc 7→ µc and Λint 7→ −E where E is what these authors have referred to rather loosely as "internal energy" density, although it is something that should not be confused with the true internal energy density U int, which will actually be given by U int = 1 2 ¯ρfc[v]2 + U ins = ¯ρfc[v]2 − Λint . (15) 2.2 Currents and their conjugate mo- menta in 4 dimensions Instead of treating quantities such as energy and momen- tum separately, as has traditionally been done in a New- tonian framework, it has been found to be technically advantageous, as remarked above, to unify them in a 4- dimensional treatment [11, 12] (which was originally in- spired by the corresponding relativistic theory [30]). The first step in this 4 dimensional treatment is to rewrite the generic variation (11) in tems of 4-current variations in the form dΛmat = µf ν dn ν f + µc ν dn ν c − µI dnI , (16) which provides a direct specification of the material 4- momentum covectors µf ν that play a crucial role in the concise canonical formalism to be described below. In terms of the quantities introduced above their time com- ponents will be given by ν and µc µf 0 = −(µf + µf iv i f ) , µc 0 = −(µc + µc iv i c ) , (17) These material 4-momenta, µf while their space components will simply be given by (13). ν , give rise to cor- responding complete "free" and "confined" baryon mo- menta, πf ν, whose time components include al- lowance for gravitational energy according to the specifi- cations ν and πc ν and µc πf 0 = µf 0 − mφ , πc 0 = µc 0 − mφ , while their space components are given simply by πf i = µf i , πc i = µc i . (18) (19) It is also useful to introduce the total ionic momentum ν , which is defined so as to include allowance covector πI for nuclear energy by πI 0 = Acπc 0 − µI , πI i = Acπc i . (20) This enables us to express the complete stress energy ten- sor of the system in the very concise form T µ ν = n µ f πf ν + n µ I πI ν + Ψ δµ ν , in which the material pressure function is given by Ψ = Λmat − n ν f µf ν − n ν c µc ν + nI µI . (21) (22) It follows that the force balance equation (6) can be rewrit- ten in the form f f ν + f I ν = f ν , (23) 2.3 Equations of state and effective mass To complete the specification of the variables introduced in the previous section it is evidently necessary to pre- scribe the particular form of the material action density Λmat, which is -- for this purpose -- most conveniently de- composible in the form Λmat = U dyn − U ins , (28) wherein U ins is the static internal energy contribution in- troduced in (14), which is specified by an appropriate "pri- mary" equation of state as a function of the three relevant number densities nf , nc nI, while the remaining dynamical energy contribution is given [15] in terms of the "normal" (crust) velocity 3 vector vi c and the relative current 3 vec- tor ni = nf [vi] , by the formula U dyn = ρ vc iv i c /2 + m niv i c + nini/2K , (29) (30) in which the K is the "mobility coefficient". This quantity K is given by a "secondary" equation of state as a function of the relevant number densities nI, nc, and nf . In the analysis that follows a particularly important role will be played by the "free" neutron 3-momentum covector µf i for which, on a local "mesoscopic" scale (small com- pared with the intervortex separation) the superfluidity property entails the irrotationality condition ∇[iµf j] = 0. It follows from (30) that the contravariant version (as ob- tained by contracting with the ordinary Euclidean space metric γij ) of this (superfluid) momentum will be given by an expression of the form in which the 4-force density f f trons is specified in terms of their vorticity tensor f ν acing on the"free" neu- µν by µf i = m⋆v i f + (m − m⋆)v i c f f ν = nµ f f µν + πf ν∇µn µ f , f µν = 2∇[µπf ν] , in terms of an effective mass variable, (24) m⋆ = nf /K , (31) (32) while the combined 4-force density acting on the ions and their "confined" baryons will be specified by the formula f I I πc This can be decomposed in the form ν = 2nµ I ∇[µ πI ν] + nµ µ∇νAc . f I ν = f c ν + nI∇ν µI − δ 0 ν ∇µ(µIn µ I ) , whose difference from the ordinary baryon mass m is pro- portional to the "entrainment density" (25) ¯ρfc = ρf (m⋆ − m)/m , (33) (26) which -- in much of the inner crust -- now seems likely [17] to be much larger than had previously been supposed. in which the last term contributes only to the time com- ponent, while the only part that remains in the absence of stratification is the contribution of first term, which is given by an expression of the standard form f c ν = 2nµ c ∇[µπc ν] + πc ν∇µn µ c . (27) 5 The reason why it is convenient for the formulation of the equations of state to start with the specification of K rather than the effective mass m⋆ or the corresponding "entrainment density" ¯ρfc in the decomposition (14) is that -- unlike m⋆ and ¯ρfc -- the "mobility coefficient" K has the advantage of being physically well defined in a manner that does not depend on the choice of the chemical basis that will be discussed in Section 2.6. Another kind of decomposition that is also independent of the choice of chemical basis, and that is instructive for the purpose of comparison with the notation used in re- lated work, including the seminal article of [7], is to intro- duce the concept of the so called "superfluid 3-velocity" i , which in stricter terminology should be referered to V S as "superfluid momentum per unit mass", as defined by setting µf i = m V S i V Si = v i c + m⋆ m [vi] . (34) This enables the dynamical energy density (30) to be rewritten (without a cross term) in the seductively sim- ple form magnetic braking of pulsar rotation. However the main reason for introducing the external force contribution f ν in the present analysis is to take account of the presence, due to the solidity of the crust, of additional stresses that we wish to estimate but for which a complete treatment would be beyond the scope of the purely fluid model used here. I , n ν f , n ν In conjunction with (3), strict application of the varia- tion principle would entail as consequence that each of the three currents n ν c should be separately conserved. This would no doubt be a very good approximation for the purpose of treating high frequency oscillation modes. However, in the medium to long timescale processes we wish to consider here, whereas it will still be reasonable to postulate conservation of the number of ionic nuclei, meaning that we we shall have (35) ∇νn ν I = 0 , (37) U dyn = m 2 (cid:0)nS V S i V Si + nN vc iv i c(cid:1) , in which the so called "superfluid particle number den- sity" nS and the corresponding"normal particle number density" nN are defined by nS = mK = ρf /m⋆ nN = nB − nS . (36) 2.4 Dynamical Equations Even after the primary (3 variable) equation of state func- tion U ins{nI, nc, nf } and the secondary (2 variable) equa- tion of state function K{nI, nc} have been specified, the formulae given so far will merely provide a coherent set of definitions, but -- except for the total baryon conser- vation law (3) -- they contain no dynamical information. To govern the evolution of the nine independent variables (namely nI and the components of the 4-vectors n ν f and n ν c ) of the system we still need eight more conditions, which can be provided by specification of the values of the 4-forces f f ν and f c ν . ν = 0. ν and f c In the absence of an external force f ν in (23) the sum of the contributions f f ν would have to vanish, a condition that will automatically be satisfied if these con- tributions are specified just by the variational principle, according to which they should each vanish separately, f f ν = f c In a dissipative model o the kind recently developed by [12] (such as might be needed to allow for the mutual re- sistivity that would be present if the temperature were too high for superconductivity) the contributions f f ν and f c ν could still add up to zero even though they would no longer vanish separately. However we wish to consider more gen- eral situations in which a non-vanishing total force density f ν may be present for two reasons. One reason is to al- low for the "secular" (very long timescale) effect of weak on the other hand it may be realistic to allow for the possibility that -- due to beta processes whereby neutrons are transformed into neutrons or vice versa -- the "free" and "confined" baryon currents may not be separately conserved, but will evolve in such a way as to diminish the magnitude of the relevant chemical affinity[12], A say, meaning the chemical potential difference between the two constituents, as measured in the relevant "normal" (non superfluid) rest frame, namely that of the ionic lattice with unit velocity 4-vector u ν c , so that it will be given by A = u ν c [µν] , [µν] = µf ν − µc ν . (38) As discussed in the treatment of the analogous relativis- tic model [30] (with nI and µI respectively replaced by entropy density and temperature) a non-dissipative model that is self contained -- so that total force f ν vanishes -- will be obtainable in a manner that is compatible either with the short term conservation conditions ∇νn ν c = 0, or alternatively with the long term equilibrium condition A = 0 that is more relevant here, provided the three force acting on the superfluid neutron current is specified in such a way that f = ∇ν n ν n µ f f µν n ν c = 0 . (39) 2.5 Superfluidity condition On a mesoscopic scale (large compared with the ionic spac- ing but small compared with the distance between quan- tised vortices) the superfluidity property of the "free" neu- trons entails that their momentum covector πf ν must be locally proportional to the gradient of a quantum phase scalar, and hence that their vorticity 2-form must vanish, f µν = 0 , which automatically ensures that the condition (39) will be satisfied. 6 The supposition that the vortices will be "pinned" in the sense of being transported with the ionic lattice is evidently expressible by adopting a dynamic equation of the form (42) with u µ w = u µ c (43) which still automatically ensures that the condition (39) will be satisfied, but which will generically entail the in- volvement of a non-vanishing force density that will be attributable to the Magnus effect as discussed in the ap- pendix. This condition (39) particularly plausible as a description of evolution on very long timescales, for which the beta equilibrium condition, A = 0 , (44) will also be appropriate. For evolution on the more moderate timescales that will be relevant for many astrophysical processes the situation is not so clear. An obvious -- though not necessarily re- alistic -- alternative way of satisfying (39) is to apply the strict variation principle to the effect that f f ν should van- ish, which instead of (43) as well as the separate "free" neutron conservation condition ∇ν n ν f = 0 also provides a dynamical equation of the form (42) but with w = u µ u µ f , (45) and which instead of (44) provides a separate conservation law ∇νn ν f = 0 . (46) However the realism of such an alternative is question- able on the basis of microphysical considerations which suggest that both the vortex slipping implied by (45) and the beta disequilibrium entailed by (46) would in prac- tise give rise to dissipative processes that would require a more elaborate treatment. This issue is hinted at, on a purely macroscopic level, by the consideration that the actual physical meaning of the simple set of non dissipa- tive evolution equations provided by (45) and (46) is open to question, due to its gauge dependence in the sense dis- cussed in the next section. 2.6 Chemical basis dependence A problem with the force free postulate, namely the use of (45) in conjunction with (46) is that (unlike the use of (43) in conjunction with (44) which will be appropriate for longer timescale processes) its physical meaning depends on how many of the neutrons are considered to be "free". To be more explicit, it depends on the choice of the dimen- sionless parameter ac in the chemical base transformation n ν c = ac np ν n ν f = n ν n + (1 − ac)np ν , (47) Figure 2: Sketch of the 2-surface swept out by a quantised vortex line. What we are concerned with here however is the macro- scopic level -- meaning lengthscales large compared with the intervortex separation -- at which there will be a non- vanishing vorticity 2-form, and a corresponding spacelike vorticity vector wλ = 1 2 ελµν f µν , (40) whose components in Aristotelian coordinates are simply w0 = 0 and wi = εijk∇jvf k, representing the mean density of circulation around the quantised vortices. The superflu- idity property does however entail an algebraic restriction to the effect that we must have wµf µν = 0 , (41) a condition need not be satisfied for a more general fluid motion (such as would be possible for the free neutrons in a very young neutron star that has not yet dropped below the temperature of the BCS pairing transition). This re- striction expresses the requirement that, instead of having matrix rank 4 as in the normal case, the antisymmetric (so necessarily even ranked) vorticity tensor f µν should have its rank reduced to 2 in the superfluid case, a condition that is mathematically necessary for it to be tangential to a set of 2-surfaces which in this instance are swept out by the quantised vortex lines as illustrated on figure 2. This implies the existence of vorticity transport 4-vector field u ν w subject to the usual time normalisation condition u 0 w = 1 characterised by the defining property w f u µ µν = 0 (42) which does not fix it completely but evidently allows a freedom of adjustment by addition of a spacelike vector field aligned with the direction of the vortex lines as given by wν . 7 whereby the "confined" and "free" baryon currents are specified in terms of the more obviously well defined chem- ν and the ical basis constituted by the proton current np total ("confined" plus "free") neutron current n ν n . Obvious simple possibilities for the choice of such a "chemical gauge" include the "comprehensive" gauge, ac in which all the neutrons are counted as "free", the "pair- ing gauge" ac = 2 in which the most strongly bound neu- tron states, namely those paired with proton states, are counted as "confined", but all the others are considered to be free. A more intuitively "natural" -- but more fuzzily defined -- possibility would be to reserve the qualification "free" for neutrons that are actually located outside the -- somewhat blurred -- boundaries of the nuclei that contain the protons. However what really matters is not location in ordinary space but in phase space: the "operational" criterion for a neutron state to be effectively "free" is that it should have sufficient energy to overcome the barriers separating the ionic potential wells either classically or by quantum tunnelling within the relevant timescale. On this "operational" basis it is clear that all the neu- trons will be effectively "free" in the fluid core and that none of them will be effectively "free" in the outer crust below the "drip" density threshold, while at densities just above this threshold there will be a clear cut critical energy above which the neutrons will be able to travel over ionic separation distances on a microscopically short timescale and below which their wave functions will be exponentially suppressed outside the ionic wells so that they will be able to tunnell only on cosmologically long timescales. How- ever in the deeper and denser layers of the crust (which are likely to be particularly important for glitch processes) this "operational" distinction will no longer be so clear cut, because the ionic wells will get too close for the exponential suppression outside to be fully effective, so that there will be marginally bound states with intermediate penetration timescales that are macroscopically long but cosmologi- caly short. This means that the most appropriate way of precisely defining an "operational" basis will depend on the timescales involved in the astrophysical context under consideration. In view of the debatability of this question about which basis may be most appropriate, it is important to observe that the counting convention used to define the chemical gauge parameter ac in (47) makes no difference at all to the specification of the "normal" velocity v i c and the rel- ative current ni defined by (29) and hence that it has no effect on any of the terms in the original specification (30) for the dynamical energy density. It is no less important to observe that uncertainties about the appropriate choice for ac make no difference to the specification of the "free" momentum covector µf ν, nor in consequence, to the vortic- ity form f i , µν and the so called "superfluid 3-velocity" V S which means that the quantities appearing in the rewrit- ten formula (35) will also be unaffected. The rule chosen for the specification of ac and hence of nf will evidently affect the individual terms in the formula (21) for the stress energy tensor, but it can be seen that the total, T µ ν , will be unaffected by the value of ac so long as it is held fixed, and even (though this is less obvious) if it is allowed to vary, so long as this ratio, ac = nc/np = AcnI/np , (48) is postulated to depend only on the nuclear charge number Z, which is itself definable as the ratio Z = np/nI . (49) It thus follows from (6) that provided ac (or equivalently Ac) is chosen as some function just of the single variable Z, although the choice of this function will indeed affect the specification of the separate force density contributions f f ν and f c ν , nevertheless it will not have any effect on their to- tal (23). This means that it will not have any effect on the complete set of evolution equations given by specifying the total external force f ν (e.g. by postulating that it should vanish) in conjunction with the ionic number conserva- tion law (37) together with the (long timescale) pinning and beta equilibrium conditions (43) and (44). However if (44) is replaced by (46), then the physical specification of the system would depend on the ansatz for the function ac{Z}, which for very short timescales could realistically be taken to be simply the "comprehensive" value ac = 1, corresponding just to separate conservation of protons and neutrons. Similarly if (43) is replaced by (45) the result would again depend on the ansatz for ac{Z}, but in this case the appropriate choice is a question needing further investigation at a microscopic level. The strategy of the following work is not to adopt any particular choice of gauge but to see how far it is possible to draw general conclusions that will be valid whatever the choice of the function ac{Z} and whatever the postulate adopted for the specification of the vortex drift velocity u µ w in (42). 3 Axisymmetric configurations 3.1 Angular momentum convection and the lag formula As in the preceding work [16] on this problem, we now restrict our attention to configurations that are axisym- metric in the sense of being invariant with respect to the action of a rotation symmetry generator ν of the usual kind, meaning one that is spacelike, 0 = 0, with space 8 We are concerned here with configurations that differ only very slightly from a state of circularity, meaning a state in which the 3-velocities are all aligned with the ax- isymmetry generator i, so that we can write vi X = ΩX i + v i X⊥ , i v i X⊥ = 0 , (54) where ΩX is the relevant angular velocity, and the residual non circular velocity contribution v i X⊥ is supposed to be very small except perhaps in very brief intervals during a glitch. The non circular part, if any, will evidently make no contribution to the superfluid angular momentum scalar, which, according to (64) will be given simply by ℓ = 2(cid:16)m⋆Ωf + (m − m⋆)Ωc(cid:17) . (55) Since the axisymmetry evidently implies i∇iℓ = 0, the conservation law (53) can be rewritten in terms of the vor- tex drift 3-velocity v i w as a slow variation rule of the form ℓ = −v i w⊥∇iℓ , (56) with the usual convention that a dot denotes partial dif- ferentiation with respect to time at a fixed space position. In the rotating neutron star configurations under con- sideration it will be justifiable to use a small perturbation approximation, not just for the non-circular velocity con- tributions, but also for deviations of the circular part from rigid rotation with a uniform fixed angular velocity value, ¯Ω say. This means that for each constituent we can write ΩX = ¯Ω + ∆ΩX (57) where, like v i w⊥ the deviation ∆ΩX is to be considered as relatively small. It follows that to first perturbative order we can rewrite the slow variation law (56) as Figure 3: Axisymmetry generator i, whose magnitude is equal to the cylindrical radius and the unit vector νi directed along the rotation axis. components i given in terms of the unit vector νi along the relevant symetry axis, with respect to Cartesian co- ordinates ri centered on the axis, by an expression of the form i = εijkνjrk as shown on figure 3. This means that this symmetry generator will have a scalar magnitude that can be interpreted as a cylindrical radial coordinate (such as was denoted by in the preceding work by Carter et al. [16]) so that it will be given in terms of the relevant angle of latitude θ by = r cos θ. Invariance of a field with respect to such an action is expressible in any (Cartesian or other) coordinate sys- tem as the vanishing of its Lie derivative with respect to the generator. In the particular case of the superfluid 4- momentum covector πf ν the axisymmetry condition will therefore be expressible as ν∇νπf µ + πf ν ∇µν = 0 . (50) ℓ = −m ¯Ω v i w⊥∇i2 . (58) In terms of the angular momentum per free (superfluid) particle, as defined (independently of any choice of the chemical gauge parameter ac) by ℓ = νπf ν , (51) (so that on a microscopic scale ℓ will be quantised as a half integer multiple of the Dirac-Planck constant ) the axisymmetry condition (50) can evidently be rewritten as ∇µℓ = f µν ν . (52) We can use this in conjunction with (42) to obtain a gen- eralised Bernoulli type [12] conservation law to the effect that this superfluid angular momentum scalar ℓ will be convected by (in the sense of remaining constant along) the vortex flux trajectories: Using square brackets for differences between the con- stituents according to the convention [Ω] = Ωf − Ωc , [ Ω] = Ωf − Ωc , (59) we can write (58) more explicitly as 2m¯Ω m⋆[ Ω] + m Ωc = − v i w⊥∇i . (60) It follows that, after a finite time interval during which the vortices have undergone a small cylindrical radial displace- ment, ∆ say, the ensuing (local) change in the angular frequency lag will be given in terms of the correspond- ing (uniform) change ∆Ω in the observable rotation fre- quency, namely that of the solid constituent Ω = Ωc by the remarkably simple formula u ν w∇νℓ = 0 . ∆[Ω] Ω = − (53) 9 m m⋆ (cid:18) ∆Ω Ω + 2 ∆ (cid:19) . (61) In the idealised case of pinning to an absolutely rigid solid the displacement ∆ would of course vanish, so that ∆[Ω] would be approximately determined as a function just of the spherical radial coordinate r by a proportion- ality relation of the form ∆[Ω] ∝ m/m⋆, but in practice, even for perfect pinning, there will be deviations from this as the crustal rigidity modulus is expected [42] to be rather low (the solid structure will be effectively "squashy"). On the other hand if it is supposed that pinning is entirely in- effective (since the vortices may drift by either thermal [3] or quantum fluctuations [32]) then the generalised Proud- man theorem derived below in Subsection 4.2 tels us that the angular velocity difference [Ω] -- and hence also its variation ∆[Ω] -- will be a function not of r but of the cylindrical coordinate = r cos θ, so in that case (to al- low for non-uniformity of m⋆) the contribution from the displacement term in (61) would have to be relatively sig- nificant. 3.2 Simplified global model For the purpose of drawing conclusions from available em- pirical data such as that of Lyne et al. [33], rather than with the locally well defined but observationally inacces- sible quantities involved in a formula such as (61) it can be more instructive in practice to work with corresponding crudely defined averages, using a total angular momentum decomposition of the form J = J f + J c (62) in terms of "free" and "confined" contributions of the form J X = Z nX µX i i d3r , (63) with the "free" bayon momentum µf i given by (31) and with the "confined" baryon momentum given by tha anal- ogous formula µc i = mv i c − mc ⋆ [vi] , , (64) in which the effective mass function mc ⋆ for the "con- fined" baryons is determined by its "free" baryon ana- logue,analogue mf ⋆ = m⋆, via the symmetric relation nc(mc ⋆ − m) = nf (m⋆ − m) . (65) Using an approximation in which Ωf is taken to be roughly uniform, as Ωc(= Ω) must be in any case, these contribu- tions will be given roughly by J f ≃ I ff Ωf + I cf Ωc , J c ≃ I ccΩc + I cf Ωf , (66) in terms of coefficients that combine to give the respective moments of inertia of the confined and free parts in the form I f = I ff + I cf , I c = I cc + I cf . (67) In accordance with (64) the separate coefficients will be given, in terms of the cylindrical radius h, as the volume integrals I cc = Z mc ⋆ nc 2 d3r , I ff = Z m⋆ nf 2 d3r , (68) while, in view of the equivalence (65) , the entrainment coefficient will be given by I cf = Z (m − mc ⋆) nc 2 d3r = Z (m − m⋆) nf 2 d3r , (69) so the subtotals will be given by expressions of the familiar form I c = mZ nc 2 d3r , I f = mZ nf 2 d3r , (70) which combine to give the complete stellar moment of in- ertia as I = I c + I f = Z ρ 2 d3r . (71) In a simplified treatment it may be supposed that these moment of inertia coefficients remain constant (i.e. that the effects of a conceivable radial displacement ∆ in (61) can be ignored) and that the superfluid contribution to the angular momentum is conserved during the interglitch period, ∆J f ≃ 0, which by (66) evidently implies ∆Ωf ≃ − Icf I ff ∆Ω . (72) It may also be supposed that the total angular momentum variation is negligible, δJ = 0 during the very short time interval in which the star undergoes a discontinuous glitch transformation involving an angular momentum transfer given by δJ f ≃ −δJ c so that the corresponding angular velocity discontinuities will be related by δΩf ≃ − I c I f δΩ . (73) To avoid a long term build up of too large a deviation of the superfluid angular velocity Ωf from the externally observable value, Ωc = Ω, the averages over many glitches of their total (continuous plus discontinuous) change per glitch should be the same, h∆Ωf + δΩf i ≃ h∆Ω + δΩi , (74) a condition that is equivalently expessible using the nota- tion of (59) as hδ[Ω]i ≃ −h∆[Ω]i . (75) The immediately preceding relations can be used to rewrite this condition in the form hδΩi h∆Ωi ≃ − (I f )2 I ff I , (76) 10 in which the left hand side contains only quantities that are directly observable, with values that have been typ- ically found [33] to be given in order of magnitude by hδΩi/h∆Ωi ≈ −10−2 in the case of Vela type pulsars. This magnitude has commonly been used as a basis for estimating the ratio I f /I of the moment of inertia of the superfluid component involved to that of the whole star, on the basis of the usual supposition that there is no need to distinguish between the coefficients I f and I ff . However it can be seen from (68) that the relation between these co- efficients will actually be expressible in terms of a suitably weighted mean value ¯m⋆ of m⋆ by the formula I ff ≃ ¯m⋆ m I f , (77) in which, according to the recent work [17] referred to above, the ratio ¯m⋆/m is likely to be substantially larger than unity, which leads to a correspondingly augmented estimate, in such a stationary configuration will satisfy the rigidity condition, meaning that Ωc is uniform with the externally observed angular velocity Ω, ∇i Ωc = 0 , Ωc = Ω , (80) it can be seen that the space parts of the force density formulae (24) and (26) will be expressible simply as f f i = nf (∇i Cf + ℓ ∇i Ωf ) , f I i = nc∇i Cc + nI∇i µI , (81) (82) using the notation of Prix et al. [39] for the comoving particle energies, which are concisely specifiable for the respective "free" and "confined" constituents as the cor- responding diagonal components C f = E f f , Cc = E c c , (83) I f I ≃ − ¯m⋆ m hδΩi h∆Ωi , (78) of the (non-symmetric) energy matrix defined by the for- mula for the relative value of the moment of inertia I f of the relevant superfluid component as compared with the total moment of inertia I of the neutron star. 4 Stationary circularly symmetric configurations. 4.1 Total stress force in the general case We now investigate the equilibrium conditions that need to be satisfied in the short and medium term when we neglect the slow secular evolution effects considered in the preced- ing section and consider the system to to be in a state that is exactly stationary, and with flow that is strictly circular in the sense that convective velocity contributions v i X⊥ in (54) are absent, so that we simply have v i f = Ωf i , v i c = Ωf i . (79) For a single-fluid model in such a non convective state the well known Proudman theorem tells us that the angular velocity must be uniform over each "Taylor cylinder" (as characterised by a fixed value of ) for the case of a perfect fluid that is barotropic, meaning that there is no stratifi- cation. It will be shown in the present section how this result can be generalised to the two-fluid context under consideration here. Using the circularity conditions (79) in conjunction with the axisymmetry conditions exemplified by (50), and as- suming that the solidity property of the crust will ensure that the "normal" flow velocity of the "confined" current E X Y = −πX νu ν Y . (84) This is another example of a formula that becomes more complicated and less heuristically meaningful when in- stead of using 4 dimensional notation it is written out using the traditional 3 dimensional notation, in which for example the comoving particle energy for the "free" con- stituent will be given by the formula Cf = mφ + χf − 1 2 m v 2 f , (85) in which, by (10) and (14), the relevant chemical potential will be given by χf = ∂ U ins/∂nf − 1 2 [v]2∂¯ρfc/∂nf , (86) with ∂¯ρfc/∂nf = nf ∂m⋆/∂nf + m⋆ − m. Without going into such details, it can be seen that the gravitational contributions to these comoving particle energies will cancel out in their difference, which will be given in terms of the chemical affinity (38) and the angular velocity difference [Ω] = Ωf − Ω by Cc − Cf = A + ℓ [Ω] . (87) If the crust were purely fluid, the force density f c i would just have to be equal in magnitude and opposite in sign to the force density acting on the superfluid, namely −f f i. Detailed evaluation of the effects of solidity would require the use of a more elaborate model incorporating elastic rigidity, but as in the preceding work by Carter et al. [16] we can allow for the effect of finite rigidity by introducing a non vanishing stress force density f i representing the 11 discrepancy according to the definition (23), which can be rewritten in the form 1 nc (cid:18)f i − ρ ρf f f i(cid:19) = f I i nc − f f i nf , (88) in order to obtain a difference on the right hand side that is particularly amenable to simplification using (87). In terms of the chemical disequilibrium force density term defined by f χ i = nc∇i A + nI∇i µI , (89) it can be seen that the formulae (81) and (82) lead to an expression of the remarkably simple form f I i nc − f f i nf = f χ i nc + [Ω]∇iℓ . (90) in the direction of ∇iℓ which is approximately that of ∇i. Consequently the condition (95) can be seen to provide us with a Proudman type theorem telling us that the energy parameter C f and the angular velocity difference [Ω] must both be uniform over each of the deformed Tayor cylinders characterised by a fixed value of the angular momentum per "free" particle, ℓ. Provided the difference [Ω] is small compared with the uniform crust angular velocity Ω it can be seen from (96) that these modified Taylor cylinders will differ only by a small deformation from the ordinary Taylor cylinders that are characterised by fixed values of the cylindrical radial coordinate . It is interesting to consider what would happen if the force given by (82) were also supposed to vanish, so that there would be no net force at all: We thus obtain an expression of the same form i = 0 ⇒ Ac∇i Cc + ∇i µI = 0 , f c (97) f i = ρ ρf f f i + f χ i + f b i , (91) as that previously derived by Carter et al. [16] in a sim- plified treatment in which the effects of entrainment and stratification were not included, with the force density attributable to centrifugal buoyancy deficit given by the same formula as previously, namely f b i = nc[Ω]∇iℓ . (92) 4.2 Generalised Proudman theorem for force-free superfluid case Leaving aside the pinned case for treatment in the next subsection, let us restrict our attention in the remainder of the present subsection to the case in which the neutron superfluid flow is postulated to be force-free, which means that we have f f i = 0 . (93) In this case the chemical and buoyancy contributions on the right of (91) will be the only sources for the stress force, which in this case will be given simply by f i = f χ i + f b i . Under these circumstances (81) reduces to the form ∇i Cf + ℓ ∇i [Ω] = 0 , in which by (55) we have ℓ = 2(cid:16)m Ω + m⋆[Ω](cid:17) . (94) (95) (96) Taking the exterior derivative of equation (95) implies that εijk(∇iℓ)∇j [Ω] must vanish so that ∇i [Ω] has to lie where Ac is the nuclear mass number as introduced in (4). This would provide an analogue of the Proudman theorem, to the effect that the confined particle energy parameter Cc and the nuclear potential function µI should both be uniform over surfaces characterised by a fixed value of Ac. A noteworthy special case of the application of the fore- going theorems is to the scenario considered by Prix et al [39] in which there is no stratification, meaning that ∇i µI = 0 and for which the differential rotation is also postulated to be uniform so that one has ∇i [Ω] = 0, in which case it follows that in the absence of any (pinning or external stress) force both Cc and Cf would be uniform, as first integral constants of the motion, in the sense that their gradients too would vanish, ∇i C c = ∇i Cf = 0. Assuming that (as well as the gravitational background φ that may have been provided by a Cowling type ap- proximation) the distributions of the angular velocities (or the corresponding angular momenta) and in our case also of the nuclear number density nI (or the corresponding stratification potential µI) have been specified in advance to have values resulting from the previous history of the system, then the specification of two more independent quantities such as Cc and Cf will suffice to determine the system completely, as remarked by Prix et al. [39], who observed that their values as first integral constants could be fixed by the specification of the globally integrated to- tals of free and confined baryons. A force free configuration with separate conservation of the free and confined constituents is not however what is most realistic in a system that is stationary over a secu- lar timescale. The assumption of separate conservation of neutrons and protons is appropriate in the context of high frequency dynamical processes [18, 5] such as may give rise to a two stream instability [6]. However in an effectively stationary context it is to be expected that transfusion 12 (by beta processes) would be able to establish the chemi- cal equilibrium condition (44) to the effect that the affinity A should vanish, which by (87) means that one will have Cc = Cf + ℓ [Ω] . (98) In the special case for which [Ω] and hence (by the gen- eralised Proudman theorem) also Cf are uniform it fol- lows, since the distribution of ℓ will be highly non-uniform (roughly proportional to the squared cylindrical radius 2) that Cc will also be highly non-uniform. We thus obtain a scenario in which unlike that considered by Prix et al.[39] requires only a single constant of integration. This can be taken to be the globally integrated total baryon number, which suffices to determine the uniform value of Cf , from which Cc will be obtainable unambiguously (without any need to specify any other arbitrary parameter) as the non- uniform field provided by substituting the relevant fixed values of [Ω] and Cf . A scenario of this mathematically well defined type differs from the kind envisaged by Prix et al. [39] in that, despite the fact that no vortex pinning is involved, it requires that the solid structure of the crust should provide a stress force density that according to (94) will be given as a sum of stratification and buoyancy deficit forces by f i = nI∇iµI + nc[Ω]∇iℓ . (99) c ∇i µI were equal in magni- This could only vanish if A−1 tude and opposite in sign to [Ω]∇iℓ which is mathemati- cally conceivable but physically implausible (except near the stellar equator) because ℓ will depend mainly on the cylindrical coordinate , whereas one would expect that the other quantities involved would depend mainly on the spherical radial coordinate r. 4.3 Constants of integration for pinned superfluid case To deal with the case when the superfluid vortices are pinned to the "normal" constituent it is convenient to fo- cus attention not so much on the comoving diagonal ele- ment Cf in the energy matrix (84) but on the cross term C = E f c = Cf + [vi] µf i , (100) which is interpretable as the energy per particle of the "free" (superfluid) neutrons with respect to the"normal" rest frame of the "confined" baryons. Unlike the diagonal elements Cf = −πf c (and the other off-diagonal element) this particular off-diagonal element C = −πf c has the property of being invariant with re- spect to the changes of chemical basis that were discussed in Subsection 2.6. f and Cc = −πc νu ν ν u ν ν u ν The usual supposition that, instead of moving freely, as was supposed in the preceding subsection, the vortices will satisfy the pinning condition expressed by (42) and (43) can be seen to be equivalent to the requirement that the "free" neutron current should be subject to a force density of the form f f i = −f J i (101) where f J i is the Magnus force density reacting on the solid structure of the confined constituent, as given in accor- dance with the usual Joukowsky formula (see appendix) by f J i = nf f ij [vj] , (102) which works out in this stationary circularly symmetric case as f J i = nf [Ω]∇iℓ . (103) It is to be noticed that, unlike the buoyancy deficit force density (92) to which it is proportional, this Magnus force density is invariant with respect to changes of chemical basis. When substituted in the general formula (91) for the total force acting on the system it provides an expression of the simple form f i = f χ i − f J i . (104) By substituting the Joukowsky formula in (81) it can be seen that in terms of the crossed particle energy (100) the vortex pinning condition can be rewritten simply as ∇i C = 0 , (105) which means that for any pinned configuration C will be a uniform integral constant, whose value can be fixed by the total globally integrated baryon number. In terms of this constant C the comoving energy per particle of the "free" constituent will be obtainable from (100) as the variable function of position given by C f = C − [Ω] ℓ . (106) It can be seen from (98) that the tranfusive equilibrium condition (44) will be expressible simply as Cc = C , (107) which means that, when the specification of the system is completed by the imposition of this chemical equilibrium condition, then C c will also have a uniform constant value. The constancy of these quantities is interpretable as a modified Bernouilli theorem of a kind that is rather ro- bust: it is to be emphasised that whereas the constancy deduced by Prix et al. [39] for Cc and Cc in the unpinned case was dependent on very restrictive assumptions about the uniformity of [Ω] and the negligibility of stratification, no such restrictions are involved in the derivation of con- stancy for Cf and Cc in the pinned case. 13 In this pinned equilibrium case, the analogue of the for- mula (99) for the stress force density that needs to be provided by the solidity of the crust will be given, as a difference between stratification and Magnus-Joukowsky contributions by the expression f i = nI∇i µI − nf [Ω]∇iℓ , (108) which, unlike (99), has the noteworthy property of being chemically invariant. The work in this section has used no approximations apart from the postulates of exact stationarity and ax- isymmetry, but if we make the realistic supposition that the configurations under consideration differ only slightly from a rigidly rotating reference state with fixed angular velocity ¯Ω then, to lowest order, we shall obtain the ex- pressions ℓ = m ¯Ω 2 , ∇iℓ = 2m ¯Ω ∇i . (109) This dependence of the angular momentum (per superfluid particle) ℓ on the cylindrical radial coodinate contrasts with the dependence mainly on the spherical radial coor- dinate r that one would expect for the number densities nI and nf , as well as for the angular velocity difference [Ω] (which, in view of (61), might be expected to be roughly proportional to the depth dependent ratio m/m⋆). This is why, as in the case of (99), it is physically implausible that the gradient of the stratification potential µI could be adjusted so as to cancel the terms in (108) except near the stellar equator. 5 Conclusions The foregoing work shows how it is possible to construct an extensive range of simple, explicitly integrable, axisym- metric stratified two-fluid neutron star models in which a small differential angular velocity [Ω] can be chosen in ad- vance as an arbitrary (implicitly history dependent) func- tion of the spherical coordinates r and θ in the pinned case, and as an arbitrary function just of the cylindrical coordinate = r cos θ in the unpinned case to which the generalised Proudman theorem applies. This idealised category of non-dissipative models in- cludes, as extreme case, the kind of force-free model con- structed by Prix et al. [39] on the basis of the postulate of separate conservation of protons and neutrons. Except in this special case it is shown that the equilibrium of the two-fluid model necessitates the presence of a balancing force (given by (99) for unpinned configurations and by (108) for pinned configurations) that must presumably be provided by stress in the solid structure of the crust (whose detailed treatment is beyond the scope of the purely fluid description used here) which can build up only to a certain point at which the breakdown responsible for the glitch phenomenon will occur. The exceptional case in which the presence of such (glitch producing) stress was avoided [39] depended on the postulate of separate conservation of protons and neu- trons, a condition that would be realistic in the short timescale context of dynamic perturbations but not in the long timescale context of stationary configurations, for which the postulate of chemical equilibrium is advocated here as a more plausible idealisation. It is to be understood that the force density given ei- ther by (99) or more likely by (108) will have to be bal- anced by a solid stress gradient, meaning that it will sat- isfy a relation of the form f i = ∇jhSij i in which hSij i is a shear stress tensor that is the trace-free part of an elastic stress tensor Sj i of the kind recently described in the framework of an elastic solid model that is already available [10] for the description of the outer crust, of which an appropriate neutron conducting generalisation will soon be ready [9] for application to the inner crust (with density above the neutron drip threshold) under con- sideration here. The idea is that this shear stress tensor hSiij will be proportional to a trace-free shear strain ten- sor that can not excede a critical magnitude beyond which the solid structure will breakdown, and that a glitch will occur either when this critical value is reached or else, in the pinned case (108), at what may be an earlier stage when the Joukowski-Magnus force density (103) reaches a vortex slippage limit at which unpinning occurs. For the study of processes occurring on intermediate timescales it will of course be necessary to include al- lowance for the dissipative processes that have been ne- glected in the idealised models considered here. As well as the controversial question of vortex slippage, [27, 26, 20, 21, 22] a particularly important issue concerns the timescales [41, 47, 45] of the weak interaction processes leading to chemical equilibrium. Acknowledgments The authors are grateful to Silvano Bonazzola, Loic Villain and Pawel Haensel for discussions concerning stratification and deviations from chemical equilibrium in neutron stars. Nicolas Chamel acknowledges support from the Lavoisier program of the French Ministry of Foreign Affairs. 14 References [1] M. Alford, J. Bowers, K. Rajagopal, "Crystalline color superconductivity", Phys. Rev. D63 (2001), 074016. [2] M.A. Alpar, H.F. Chau, K.S. Cheng, D. Pines, "Post- glitch relaxation of the VELA pulsar after its first eight large glitches - A reevaluation with the vortex creep model", Ap. J. 409 (1993) 345-359. [3] M.A. lpar, D. Pines, P. .W Anderson, J. Shaham, "Vortex creep and the internal temperature of neu- tron stars. I - General theory", Ap. J.276 (1984) 325- 334. [4] M.A. Alpar, J.A. Sauls, "On the dynamical coupling between the superfluid interior and the crust of a neu- tron star" Ap. J. 327 (1988) 723-725. [5] N. Andersson, G.L. Comer, D. Langlois, "Oscillations of general relativistic superfluid neutron stars", Phys. Rev. D66 (2002) 104002. [6] N. Andersson, G.L. Comer, R.Prix, "The super- fluid two-stream instability and pulsar glitches", M.N.R.A.S. 354 (2004) 101-110. [7] A. F. Andreev, E. P. Bashkin, "Three velocity hy- drodynamics of superfluid solutions", Sov. Phys. J.E.T.P. 42 (1976) 164-167. [8] G. Baym, C. Pethick, D. Pines, M. Ruderman, "Spin up in neutron stars: the future of the Vela pulsar", Nature 224 (1969) 872. [9] B. Carter, E. Chachoua, "Newtonian mechanics of neutron superfluid in elastic star crust", ArXiv preprint astro-ph/0601658. [10] B. Carter, E. Chachoua, N. Chamel, "Covariant New- tonian and Relativistic dynamics of (magneto)-elastic solid model for neutron star crust." Gen. Rel. Grav. 38 (2006) 83-119. [11] B. Carter, N. Chamel, "Covariant analysis of New- tonian multifluid models for neutron stars: I Milne- Cartan structure and variational formulation", Int. J. Mod. Phys. D 13 (2004) 291-325. [12] B. Carter, N. Chamel, "Covariant analysis of Newto- nian multifluid models for neutron stars: II Stress - energy tensors and virial theorems; III Transvective, viscous and superfluid drag dissipation", Int. J. Mod. Phys. D 14 (2005) 717-774. [13] B. Carter, N. Chamel, P. Haensel, "Entrainment co- efficient and effective mass for conduction neutrons in neutron star crust: Simple nuclear models", Nucl. Phys. A 748 (2005) 675-697. [14] B. Carter, N. Chamel, P. Haensel, "Effect of BCS pairing on entrainment in neutron superfluid current in neutron star crust", Nucl. Phys. A 759 (2005) 441- 464. [15] B. Carter, N. Chamel, P. Haensel, "Entrainment coef- ficient and effective mass for conduction neutrons in neutron star crust: Macroscopic treatment", ArXiv preprint astro-ph/0408083. [16] B. Carter, D. Sedrakian, D. Langlois, "Centrifugal buoyancy as a mechanism for neutron star glitches", A&A. 361 (2000) 795 - 802. [17] N. Chamel, "Band structure effects for dripped neu- trons in neutron star crust", Nucl. Phys. A 747 (2005) 109-128. [18] G.L. Comer, N. Andersson, "On the dynamics of su- perfluid neutron star cores", M.N.R.A.S. 328 (2001) 1129-1143. [19] G.L. Comer, R. Joint, "A relativistic mean field model for entrainment in general relativity", Phys. Rev. D68 (2003) 023002. [20] D. Dall'Osso, G.L. Israel, L. Stella, A. Possenti, E. Perozzi, "The Glitches of the Anomalous X-Ray Pul- sar 1RXS J170849.0-400910", ApJ.599 (2003) 485- 497. [21] P. Donati, P.M. Pizzochero, "Is there Nuclear Pinning of Vortices in Superfluid Pulsars?", Phys. Rev. Lett. 90 (2003), 211101 [22] P. Donati, P. M. Pizzochero, "Fully consistent semi- classical treatment of vortex-nucleus interaction in rotating neutron stars", Nucl. Phys. A 742 (2004), 363-379. [23] L. M. Franco, B. Link, R. I. Epstein, "Quaking neu- tron stars" Ap. J. 543 (2000) 987-994. [24] G. Hobbs, A. G. Lyne, B. C. Joshi, M. Kramer, I.H. Stairs, F. Camilo, R. N. Manchester, N. D'Amico, A. Possenti, V. M. Kaspi, "A very large glitch in PSR J1806-2125", M.N.R.A.S. 333 (2002) L7-L10. [25] K. Iida & K. Sato, "Spin-down of Neutron Stars and Compositional Transitions in the Cold Crustal Mat- ter", Ap. J. 477 (1997) 294-312. 15 [26] D. I. Jones & N. Andersson, "Freely precessing neu- tron stars: model and observations", M.N.R.A.S. 324 (2001) 811-824. [40] R. Prix, J. Novak, G.L. Comer, "Relativistic numer- ical models for stationary superflid neutron stars" Phys. Rev. D71 (2005) 043005. [27] P. B. Jones, "The origin of pulsar glitches", M.N.R.A.S. 296 (1998) 217-224. [28] V. M. Kaspi, J. R. Lackey, D. Chakrabarty, "A Glitch in an Anomalous X-Ray Pulsar", Ap. J. 537 (2000) L31-L34. [29] V. M. Kaspi, F. P. Gavriil, "A Second Glitch from the "Anomalous" X-Ray Pulsar 1RXS J 170849.0- 4000910", Ap. J. 596 (2003) L71-L74. [30] D. Langlois, D. Sedrakian, B. Carter, "Differential rotation of relativistic superfluid in neutron stars", MNRAS 297 (1998) 1198-1201. [41] A. Reisenegger, "Constraining Dense Matter Super- fluidity through Thermal Emission f Ap. J. 485 (1997) 313-318. [42] M. Ruderman, "Neutron star crustal plate tectonics. I. Magnetic dipole evolution in lillisecond pulsars and low-mass X-ray binaries", Ap. J. 366 (1991) 261-269. [43] M. Ruderman, T. Zhou, K. Chen, "Neutron star mag- netic field evolution, crust movement and glitches", Ap. J. 492 (1998), 267-280. [44] J.A. Sauls, in Timing Neutron Stars ed. H. Ogelman, E.D.P. Van den Heuvel (Kluwer, Dordrecht, 1989) 457-490. [31] B. Link, "Constraining Hadronic Superfluidity with Neutron Star Precession", Phys. Rev. Lett. 91 (2003) 101101. [45] L. Villain & P. Haensel, "Non equilibrium beta processes in superfluid neutron star cores", ArXiv preprint astro-ph/0504572. [32] B. Link, R. I. Epstein, G. Baym, "Superfluid vortex creep and the rotational dynamics of neutron stars", ApJ.403 (1993) 285-302. [46] N. Wang, R.N. Manchester, R.T. Pace, M. Bailes, V.M. Kaspi, B. W. Stappers, A. G. Lyne, "Glitches in southern pulsars", M.N.R.A.S. 317 (1998) 843-860. [33] A.G. Lyne, S. L. Shemar, F. G. Smith, "Statistical studies of pulsar glitches", M.N.R.A.S. 315 (2000) 534-542. [47] D. G. Yakovlev, A.D. Kaminker, O. Y. Gnedin, P. Haensel, "Neutrino emission from neutron stars", Phys. Rep. 354 (2001) 1-155. [34] G. Mendell, "Superfluid hydrodynamics in rotating neutron stars. I Nondissipative equations", Ap. J. 380 (1991) 515-529. 6 Appendix [35] M. Morii, N. Kawai, N. Shibazaki, "Pulse Profile Change Possibly Associated with a Glitch in an Anomalous X-Ray Pulsar 4U 0142+61", Ap. J. 622 (2005) 544-548. [36] M.A. Perez-Garcia, N. Corte, L. Mornas, J.P. Suares- Curieses, J. Diaz-Alonso, "Formation of an ordered phase in nuclear matter" Nucl. Phys. A 699 (2002) 939. [37] P.M. Pizzochero, L. Viverit, R.A. Broglia, "Vortex nucleon interaction and pinning forces in neutron stars" Phys. Rev. Lett. 79 (1997) 3347-3350. [38] R. Prix, "Slowly rotating two-fluid neutron star model" A&A. 352 (2000) 623 - 631. [39] R. Prix, G.L. Comer, N. Andersson, "Slowly rotat- ing superfluid neutron star model with entrainment", A&A. 381 (2002) 178-196. 6.1 Joukowsky formula for Magnus force The purpose of this appendix is to derive the formula used above for the force exerted on the fluid in the pinned case and to verify that it does indeed balance the collective effect of the Magnus forces exerted on the individual vor- tices. We shall start by evaluating the latter, using the generalised Joukowski formula that has recently been de- rived [12] for the force exerted on an isolated vortex by a mixture of irrotational perfect fluids. The present application is to the case of a vortex that is fixed with respect to the crust rest frame characterised by the flow 4 vector u µ c of the "normal" constituent, and that is oriented in the direction specified by a spacelike unit 4-vector νµ say, for which the force per unit length will be given by F ν = Cf nµεµνρσνρu σ c (110) where nµ is the (chemical gauge invariant) relative flow current vector of the superfluid and C is its (chemical 16 gauge invariant) circulation as given by 6.2 Effect of entrainment on vortex den- Cf = I πf µdxµ , (111) in which the integral is carried out along any closed path surrounding the vortex. Since the effect of many such vortices in closely spaced parallel array will be represented by a vorticity vector wµ = w νµ (112) of the type introduced in (40), whose magnitude w is the value of the total circulation per unit area, it follows that the corresponding force density (obtained as force per unit length per unit area) will be given by f J ν = w Cf F ν so we end up with the formula f J ν = nµεµνρσwρu σ c (113) (114) in which it is to be observed that the value of the cir- culation Cf round an individual vortex has cancelled out. (Thus it is not necessary to know that it will actually be given in terms of the Dirac-Planck constant by the formula Cf = π in which the usual factor 2 is missing because of the pairing due the fermionic nature of the neutrons.) To obtain the force density on the fluid we start from the formula (24) in which, since we are considering a sta- tionary axisymmetric situation, the divergence term will automatically drop out leaving the expression f f µ = nν f f νµ . (115) which would be equivalent just to the ordinary Euler equa- tion if the force term vanished. It is evident that the pin- ning condition c f uν νµ = 0, can be used to rewrite (115) in the form f f µ = nf (u ν f − u ν c )f νµ = nνf νµ . (116) (117) sity It is of interest to consider the effect of entrainment on the vorticity magnitude w which is the circulation per unit area round an infinitesimal circuit orthogonal to the local axis of rotation. Remarking that the circulation Cf is sim- ply given by 2πℓ, where ℓ is the angular momentum per free particle, it can be seen from (55), by considering the circulation between nearby circuits with crylindrical ra- dius values and +δ, that the local vorticity magnitude will be expressible in terms of the angular velocity differ- ence [Ω] = Ωf − Ωc and the mass increment mf c = m⋆ − m as w = δCf 2πδ = 2mΩf + 2mf c[Ω] + δ(m⋆[Ω]) δ . (120) Knowing that the circulation round an individual quan- tised vortex line in the fermionic superfluid condensate will be given by Cf = π we see that the corresponding surface number density of quantised vortex lines will simply be given by w/Cf = w/(π). The first term in equation (120) can be recognised as the well-known formula for a single superfluid. In practice, the extra terms proportional to [Ω] and its gradient will be negligible for the astrophysical applica- tions under consideration, in which we shall generally have [Ω] ≪ Ω, provided we are using an "operational" gauge in which magnitude of the effective mass m⋆ will remain com- parable with the ordinary baryonic mass m. If the evalua- tion were carried out in a "comprehensive" gauge, meaning a chemical basis in which all the neutrons are deemed to be "free", the relevant value of the effective mass would become extremely large near the "neutron drip" transi- tion, but the value of the (gauge dependent) difference [Ω] would correspondingly become even smaller, since the product m⋆[Ω] is chemically invariant, so that the same value would ultimately be obtained for the vorticity mag- nitude w itself. It can also be seen to follow from (116) that, in terms of the vector introduced in (40), the vorticity form will be expressible as f νµ = εµνρσwρu σ c , so that (114) can be rewritten in the form f J µ = −nνf νµ , which evidently does indeed exactly balance (117). (118) (119) 17
astro-ph/0212467
2
0212
2003-06-12T09:59:56
Discs in early-type lensing galaxies: effects on magnification ratios and measurements of $H_0$
[ "astro-ph" ]
Observations of early-type galaxies, both in the local universe and in clusters at medium redshifts, suggest that these galaxies often contain discs or disc-like structures. Using the results of Kelson et al. (2000) for the incidence of disc-components among the galaxies in the redshift z=0.33 cluster CL 1358+62, we investigate the effect of disc structures on the lensing properties of early-type galaxies. Statistical properties, like magnification cross sections and the expected number of quad (four-image) lens systems, are not affected greatly by the inclusion of discs that contain less than about 10 per cent of the total stellar mass. However, the properties of individual lens systems are affected. We estimate that 10-30 per cent of all quad lens systems, with early-type deflector galaxies, would be affected measurably by the presence of disc components. Intriguingly, the image magnification ratios are altered significantly. The amplitude of the predicted change is sufficient to explain the observed magnification ratios in systems like B1422+231 without requiring compact substructure. Furthermore, time delays between images also change; fitting a bulge-only model to early-type lenses that in fact contain a disc would yield a value of the Hubble constant H_0 that is systematically too low by about 25 per cent.
astro-ph
astro-ph
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000) Printed 13 June 2018 (MN LATEX style file v2.2) Discs in early-type lensing galaxies: effects on magnification ratios and measurements of H0 Ole Moller1, Paul Hewett2 and A. W. Blain3 1Kapteyn Institute, PO Box 800, 9700 AV Groningen, the Netherlands. 2Institute of Astronomy, Madingley Road, Cambridge CB3 0HE, UK. 3Department of Astronomy, California Institute of Technology, CA91125, Pasadena, USA 13 June 2018 ABSTRACT Observations of early-type galaxies, both in the local universe and in clusters at medium redshifts, suggest that these galaxies often contain discs or disc-like struc- tures. Using the results of Kelson et al. (2000) for the incidence of disc-components among the galaxies in the redshift z = 0.33 cluster CL 1358+62, we investigate the effect of disc structures on the lensing properties of early-type galaxies. Statistical properties, like magnification cross sections and the expected number of quad (four- image) lens systems, are not affected greatly by the inclusion of discs that contain less than ∼ 10 per cent of the total stellar mass. However, the properties of individual lens systems are affected. We estimate that 10 − 30 per cent of all quad lens systems, with early-type deflector galaxies, would be affected measurably by the presence of disc components. Intriguingly, the image magnification ratios are altered significantly. The amplitude of the predicted change is sufficient to explain the observed magnification ratios in systems like B1422+231 without requiring compact substructure. Further- more, time delays between images also change; fitting a bulge-only model to early-type lenses that in fact contain a disc would yield a value of the Hubble constant H0 that is systematically too low by ∼ 25 per cent. gravitational lensing -- galaxies: elliptical and lenticular -- galaxies: Key words: formation -- cosmology: dark matter -- cosmology: distance scale -- method: numerical 1 INTRODUCTION In current models of hierarchical galaxy formation the mor- phological type of a galaxy is determined to a large extent by its merger history (Abraham & van den Bergh 2001): late- type galaxies have not undergone a recent massive merger, so that stable, prominent discs formed, whereas early-type galaxies have been prevented from forming a prominent disc by a major merger that either consumes all the available gas for star formation, or redistributes angular momentum via large scale winds (Somerville & Primack 1999). This be- haviour has also been recovered in hydrodynamical simu- lations of galaxy formation (Steinmetz & Muller 1995). So far, there is no evidence for any evolution in the luminos- ity and size of discs in late-type galaxies up to redshifts of about 1 (Simard et al. 1999; Ferguson & Clarke 2001). This suggests that if mergers occur after that time, they are either not massive enough to significantly affect the disc structures or they destroy visible disc structures, producing elliptical systems. It is not known on what time scales disc structures form. Simulations and some observations suggest that disc structures form slowly between redshifts 0.1 and 1 (Mo, Mao & White 1998) but there is also evidence that some large discs are already present at higher redshifts (van Dokkum & Stanford 2001). Most likely, discs in early-type galaxies are the result of remaining gas settling after ma- jor merging events between two late-type galaxies (Naab & Burkert 2001). Another possible formation scenario for discs involves slow accretion from many minor merging events. Debris from minor mergers, or high velocity clouds in the halo of the galaxy, could rain down onto the disc (Benjamin & Danly 1997; Putman et al. 2002). However, such a scenario would lead to thicker, less pronounced disc structures. Since late-type galaxies exhibit clear disc structures, this process is unlikely to be very important in the formation of typical spiral structures. In all these scenarios, the obvious question arises as to whether disc structures are exclusively found in late-type galaxies. Observationally, that is clearly not the case. S0 galaxies in the local universe possess disc components, and even local ellipticals show evidence of disc structures and dust lanes when studied in detail (Rest et al. 2001; Tran et al. 2001). The isophotes of ellipticals are usually classified 2 Ole Moller, P. Hewett & A.W. Blain into "boxy" and "discy", types. Pointed, discy isophotes may well be caused by the presence of discs. Even in the case when the isophotes are nearly round and show no sign of "discyness", kinematic studies suggest that it is still possible to hide a disc in such galaxies (Romanowsky & Kochanek 1997). There is also some observational evidence for discs in el- lipticals at cosmological redshifts; Kelson et al. (2000) found that many early-type galaxies in CL 1358+62 at z = 0.33 possess evidence for the presence of exponential discs, sug- gesting discs are present in a large fraction of early-type galaxies, at least in clusters. Since there is no clear under- standing yet on how these discs form, and no systematic ob- servational investigation into the abundance and properties of discs in early-type field galaxies has been carried out, it is difficult to make precise statements about the abundance of discs in field ellipticals. However, some recent observations using the Hubble Space Telescope (HST ) show that a frac- tion of at least ∼ 50 per cent of field ellipticals contain discs of sizes spanning a range of at least an order of magnitude, from small nuclear discs to large discs extending beyond the bulge (Rest et al. 2001). In addition, kinematic studies using the SAURON integral field spectrograph, also suggest that a high fraction of ∼ 50 per cent of field ellipticals contain discs of various sizes (de Zeeuw et al. 2002). Most lens galaxies known to date are early-type galax- ies. The main reason for this is that early-type galaxies tend to be the most massive and hence have a larger cross sec- tion for gravitational lensing. However, as shown by Blain, Moller & Maller (1999) spiral galaxy lenses are also pre- dicted to contribute significantly to the total lensing cross section. This is mainly due to the high surface mass density of the thin disc component when viewed edge on (Maller, Flores & Primack, 1997; Moller & Blain, 1998). As shown in those papers, the disc components of late-type lenses can make important contributions to the lensing potential and change the generic properties of the lens systems. The dis- covery of spiral lens systems in the CLASS lensing survey, e.g. B1600+434 (Koopmans, de Bruyn & Jackson 1998), confirm that this is the case. However, early-type lensing galaxies have so far nearly always been modelled with a sim- ple one-component model, possibly including external shear (Keeton, Kochanek & Seljak, 1997; Leh´ar et al., 2000). In this paper we determine the lensing properties of early-type galaxies that contain realistic discs with stellar masses of up to ∼ 10 per cent of that of the bulge. We will determine the lensing properties of early-type lenses and show how discs in ellipticals could explain some properties of observed lens systems that have not yet been modelled successfully. We concentrate on three main questions: (i) statistical lensing properties: how do disc structures in early-type lensing galaxies affect the statistical proper- ties of a sample of early-type lenses? We calculate magni- fication cross sections and the expected ratio of quad: two- image systems. In particular, we investigate whether discs in early-type lenses may explain the high ratio of quad image systems, which is still an unsolved enigma in gravitational lensing (Rusin & Tegmark, 2001, but see Chae, 2002). (ii) magnification ratios: can disc structures explain some or all of the unusual magnification ratios observed in some early-type lens systems? To date, these deviant magnifica- tion ratios have been interpreted as providing strong evi- dence in favour of the presence of compact halo substructure as predicted by current cold dark matter (CDM) simulations (e.g. Metcalf & Madau 2001). (iii) time delays: how would disc structures affect the measured time delays in early-type lens systems? Currently there seems to be a strong disagreement between the value of H0 as found in the HST Key Project (Freedman et al. 2001) and measurements of H0 from gravitational lensing (Kochanek 2003). Some local estimates of H0 (e.g.Saha et al 2001) also favour lower values. We investigate whether disc structures in gravitational lens systems may reduce this dis- crepancy. We start in §2 by discussing the appearance of the model lens galaxies and show that the manifestations of the presence of a disc are similar to those observed in some ellipticals. In §3 we present the parametric lens models used in this paper. In §4 we describe some general and statistical lensing properties of early-type galaxies that include disc structures. In §5 we show how discs in ellipticals could explain some of the observed magnification ratios and in §6 we calculate how the presence of discs in ellipticals would affect estimates of H0 from lensing. We conclude with a summary in §7. Throughout this paper we assume a Friedmann- Robertson-Walker Universe with ΩM = 0.3, ΩΛ = 0.7 and H0 = 50 km s−1 Mpc−1. The choice for this value of H0 is motivated due to limitations of the original lensing code. Subsequently, the code has been extended to allow for any value of H0, as necessary for the calculations in §6, but we did not recalculate the results of §3-5, as these calculations are very time-consuming and a change in H0 has little effect on the results. A value of H0 = 70 km s−1 Mpc−1, is prob- ably more in line with current observational constraints, as given for example by the HST Key Project (Freedman et al. 2001) and the WMAP results (Bennett et al. 2003). 2 SURFACE BRIGHTNESS PROFILES The existence of disc components with large scale-lengths in luminous elliptical galaxies has been known for some time (Bender et al. 1989; Rix & White 1990; Rix 1991). Recent theoretical work (Naab & Burkert 2001; Barnes 2002) has begun to address the origin of such components in bulge-dominated galaxies. However, the ability to reli- ably determine the prevalence of discs in strongly bulge- dominated systems has only been possible following the de- velopment of full two-dimensional decomposition techniques (Byun & Freeman 1995). Given data of adequate signal-to- noise ratio and angular resolution, two-dimensional decom- position methods are capable of identifying discs in systems with bulge fractions (B/T) as high as B/T=0.95. A signif- icant quantity of HST imaging, obtained using the Wide Field Planetary Camera 2 (WFPC2), of luminous bulge- dominated galaxies in intermediate redshift clusters is well- suited to such decompositions but little work appears to have been undertaken. Exceptions are the investigations of the z = 0.33 cluster CL1358+62 by Kelson et al. (2000) and Tran et al. (2003). Fig. 4 of Kelson et al. provides a partic- ularly striking demonstration of the presence of discs, even among galaxies with conventional 'E' morphological classi- fications. In order to parameterize the light profile of ellipticals, a commonly used set of parameters are the half-light radius rh and the surface brightness at the half light radius Ih. The surface brightness of a de-Vaucouleurs bulge can then be written as (Binney & Tremaine 1987), I(r) = Ih exp(cid:8)−7.67[(r/rh)(1/4) − 1](cid:9). The parameterization of the surface brightness distribution for an exponential disc, in terms of the central surface bright- ness, Id, and scale -- length, rd, can be written (1) I(r) = Id exp −(r/rd). (2) The bulge fraction, B/T, for a galaxy parameterized by such bulge and disc components is then B T = Ihr2 h Ihr2 h + 0.28Idr2 d (3) where Ih and rh, Id and rd are as defined above. Table 2 in Kelson et al. gives surface brightnesses and half-light-radii for bulge-only and bulge-plus-disc fits to galaxies in CL1358+62. Note that Kelson et al. specify the properties of the disc component in terms of half-light radius of the disc, rhd = 1.688rd, and the mean surface brightness within the half-light radius, < Ihd >. Using this parameter- ization the B/T ratio may be calculated by replacing the term 0.28Idr2 d by < Ihd > r2 hd. For the investigation of the general and statistical lens- ing properties of discy ellipticals, we use the full sample of galaxy models in Kelson's catalogue. However, our main results, as presented in § 5 and § 6 involve time-consuming calculations and are therefore based on the parameters de- scribing a single model galaxy. The presence of many galax- ies with bulge and disc components with comparable scale- lengths in CL1358+62 is clear from the form of the residuals shown in Figure 4 of Kelson et al. It is worth stressing that the presence of even highly inclined (almost edge-on) discs, with scale-lengths similar to those identified by Kelson et al., making up less than ∼ 10 per cent of the total light are hard to discern directly in images. Plots showing the residuals ob- tained by fitting a bulge-only model are, however, far more revealing. The presence of face-on discs is revealed by char- acteristic donut-like residuals, while edge-on discs appear as often quite prominent inclined structures. Notwithstanding the evidence provided by Kelson et al. in support of the ac- curacy of their bulge-disc compositions we have chosen to be conservative by reducing the disc contribution to each galaxy in the Kelson et al. catalogue by half. Thus, our in- vestigation of the statistical lensing properties is based on the properties of the galaxies listed in Table 2 of Kelson et al. but with a reduction in the disc contribution of a factor 2. For the individual model galaxy used in our calculations we have chosen galaxy no. 242 (Table 1). Adopting our rescal- ing of the disc contribution, galaxy no. 242 has B/T=0.95, c.f., B/T=0.91 in Kelson et al. Galaxy no. 242 has a total mass, B/T ratio and bulge scale-length close to the average value for the 'E' galaxies in Kelson et al's catalogue. In our simulations the ellipticity of the bulge is set to e = 0.4 and the inclination of the disc is i = 75 deg. Fig. 1 shows images based on our adopted parameteri- zation of galaxy no. 242 with B/T=0.95. Fig. 1a is an image based on the bulge-only model surface-brightness profile fit (parameters from Cols: 3 & 4 in Table 1), Fig. 1b shows a Discs in early-type lens galaxies 3 realization of the bulge-plus-disc surface-brightness profile (parameters from Cols: 7 & 8 (bulge) and Cols: 10 & 11 (disc) in Table 1). The disc is at high inclination, i = 70 deg to the line of sight (i = 90 deg corresponding to edge-on) and oriented at an angle of 45 deg with respect to the x- axis. The presence of the disc is barely discernible. Fig. 1c shows the residual image derived from subtracting Fig. 1a from Fig. 1b (both normalized to possess the same total counts). The images may be compared to those in Figs 1 and 4 of Kelson et al.. In Kelson et al., galaxy no. 242 shows the donut-like residual characteristic of the presence of a close to face-on disc. By contrast, in Fig. 1c, the disc ap- pears prominently because of the high inclination angle of i = 70 deg to the line-of-sight. Following submission of the original version of this pa- per an analysis of the photometric properties of galaxies in CL 1358+62 undertaken by Tran et al. (2003) appeared. Comparison of both the B/T values and the scale-lengths of the light distributions for the bulge-dominated galaxies in common between Kelson et al. and Tran et al. show a good correlation over the full range of B/T. However, for large values of B/T the contribution of the disc components are much larger in the Tran et al. analysis. Less satisfac- tory still is the comparison between the scale-lengths of the disc components found among the bulge-dominated galaxies. The disc scale-lengths found by Tran et al. are significantly larger, explaining, at least in part, the systematically smaller B/T ratios found for the bulge-dominated systems. Tran et al. present the results of a large number of Monte-Carlo sim- ulations designed to quantify the accuracy of the decompo- sitions into bulge and disc components. No reference is made to the strong systematic disagreement between the statistics of the disc scale-lengths found in the two studies. Tran et al. are confident that their simulations demonstrate the ability of the software employed to obtain accurate bulge-disc de- compositions. However, the parameter ranges employed in the simulations appear to be determined from the results of the actual decompositions applied to the real data. Tran et al. find no discs with scale-lengths comparable to or smaller than those of the bulge components of the galaxies. Thus, it is not clear that the ability of Tran et al.'s analysis to cor- rectly parameterize combinations of bulge and disc profiles with parameters found by Kelson et al. has been adequately tested. In the circumstances we have decided to retain the very conservative distribution of B/T ratios, with associated values of component scale-lengths, based on the distribution advocated by Kelson et al. 3 THE LENS MODELS In order to study the lensing properties of early-type galax- ies including discs, a working model of the mass distribution is needed. Unfortunately, little is known about the detailed properties of early-type galaxy discs and how they depend on morphology and environment. From past studies it is clear that discs in ellipticals are abundant and that the disc sizes and magnitudes cover the whole range from compact and luminous to extended and faint (Rest et al. 2001; Scorza & Bender 1995; Scorza et al. 1998). Most of the available infor- mation is based on observations of individual galaxies; there has not yet been a systematic high-resolution study of the 4 Ole Moller, P. Hewett & A.W. Blain Figure 1. Simulated images of a typical model galaxy at redshift z = 0.33. The galaxy model parameters correspond to our modified representation of galaxy no. 242 in Kelson's catalogue (Table 1). Panel (a) shows the image derived using a bulge-only model. Panel (b) shows the bulge-plus-disc model, normalized to have the same total counts as the bulge -- only model. The disc component of the model is inclined at 70 deg to the line of sight and oriented at an angle of 45 deg, to the x axis. Panel (c) shows the residual following subtraction of the bulge-only image from the bulge-plus-disc image. The pixel scale, 0.1 arcsec per pixel, and point spread function have been chosen to approximate images obtained using the WFPC2 on HST. The signal-to-noise ratio of the synthetic images is significantly higher than that typical of HST imaging data of the type analysed by Kelson et al. (2000). morphologies of a complete, unbiased sample of early-type galaxies. In order to make meaningful predictions about the lensing properties of early-type lensing galaxies including disc components, we base our models on the light profiles obtained in Kelson et al. (2000). All galaxies in Kelson's sample are at a redshift of z = 0.33, which corresponds roughly to the redshift at which lensing is most efficient; few lenses are expected with redshifts less than z < 0.1. then rh = 7.6764rb and Ih = Γ−1Σb exp (−7.676) for the bulge component and Id = Γ−1Σd for the disc component. We will refer to rh and rhd = 1.688rd as the 'effective ra- dius' or 'half-light radius' of bulge and disc, respectively. Assuming randomly oriented discs, the average inclination is about 70 deg. Unless otherwise stated, the discs in our bulge-plus-disc models have this inclination. 3.1 Parameterization of early-type lenses with 3.2 Possible caveats of the model discs the We model lensing galaxies with two- components: an elliptical bulge with a de-Vaucouleurs r1/4 profile and an exponential thin disc. elliptical Under the usual assumption of a thin lens, the lensing properties depend solely on the surface mass density profile. For the bulge component this is given by Σb(r) = Σb exp (−r′/rb)1/4, (4) where Σb is the central surface mass density of the bulge, rb is the bulge scale-length, and r′ = px2 + y2(1 − e)2. Here r = px2 + y2, and e is the ellipticity of the bulge. The disc, inclined at an angle i to the line of sight, can be parameterized in a similar way, to give (5) Σd(r) = Σd exp (−r′/rd), (6) where Σd is the central surface mass density of the disc, rd is the scale-length of the disc and r′ = px2 + y2[1 − cos (i)]2. We convert the light profiles as given in equations (1) & (2) to models of the mass distribution assuming a constant mass-to-light ratio, Γ. The relation between the parameters in equations (1) & (2) and those in equations (4) & (6), is (7) The photometric profile decompositions of galaxies in CL 1358+62 provide an observational motivation for mod- elling early-type galaxies including disc structures. However, galaxies in cluster environments might exhibit more pro- nounced discs than corresponding galaxies in the field. If this is indeed the case, and simulations of Naab & Burk- ert (2001) suggest otherwise, the discs in our model lensing galaxies could be too pronounced compared to those ex- pected in observed early-type lens systems, many of which are in lower-density field environments. A comparison of the disc sizes obtained from a sample of mostly field ellipticals by Scorza & Bender (1995) with the sizes of discs in E's and S0's found in Kelson et al. (2000) shows that the discs from both samples are similar and range in disc half-light radii from 0.1 kpc to 1 kpc. This could indicate that the en- vironment may influence the relative number of galaxies as a function of morphology, but has little effect on the prop- erties of discs in galaxies of a given morphological type. In any case, we are not concerned here with accurate modelling of the statistical incidence of discs, but rather aim to show how discs with sizes and masses consistent with current ob- servations influence the gravitational lensing properties of early-type lens systems. There is a second possible caveat with the mass model we use here: we assume that mass follows light and do not include a dark matter halo. As we will discuss in more de- tail below, the main effect of inclined, thin disc components Discs in early-type lens galaxies 5 Figure 2. Radial surface mass density profiles for representative models of the galaxies in the cluster CL 1358+62. The different lines show the surface mass density as a function of radius for dif- ferent model parameters that are representative for the majority of all galaxies in the sample. All curves are normalized so that the total mass is 1011M⊙. The ellipticity of the bulge is e = 0.4 and the inclination angle of the disc is i = 70 deg. The surface mass density differs by less than 10 per cent between models that include a disc with Md = 0.1Mb and models without a disc. At small radii, discs increase the surface mass density, due to their smaller scale lengths. At the cluster redshift of z = 0.33, 1 arcsec corresponds to 6.6 kpc for the cosmology we adopt in this paper. on the lensing properties of galaxies arises from the strong asymmetry that is introduced in the central regions of the galaxies due to the disc. We therefore expect that the pre- cise form of the spherical or nearly-spherical components in which the disc is embedded, that is, the bulge and halo, will not change the effect of the added disc component drasti- cally. We have checked this explicitly for the mass model used to derive the results in §5 and §6, by adding a dark matter halo that contains 2/3 of the total mass and scaling the mass of the bulge and disc so that the Einstein radius is the same as for the model without a dark-matter halo. Estimates of the mass-to-light ratio of early type galaxies range from M/L ∼ 2 in the K-band (e.g. Moriondo, Gio- vanardi & Hunt 1998) to about M/L ∼ 10 in the V-band (Natarajan et al. 1998). Many lens systems are consistent with lower mass-to-light ratios within an Einstein radius of about M/L ∼ 2 in V (Koopmans & Treu 2003), motivating our choice of a dark halo containing 2/3 of the total mass. When including such a spherical dark halo we find that the inner caustic shrinks in size, and so quad images become less likely, but that the magnification ratios and time delays of images of sources that are quadruply imaged do not change substantially. 3.3 Surface density and mass profiles of lens models We show the radial surface mass density profiles for some typical cases in Fig. 2. The solid lines show the density pro- files of three bulge-only models with effective radii of 0.2, 1.0 and 2.0 arcsec. At the cluster redshift of z = 0.33 these val- ues correspond to 1.3, 6.6 and 13.3 kpc respectively. These values cover the range of half-light radii of the bulge-only fits Figure 3. Radial mass profiles for a number of galaxy models. The different lines show the enclosed mass as a function of radius for different model parameters. The curves are normalized so that limR→∞ M (R) = 1011M⊙. The effective radii for bulge and disc are rhd = rh = 0.2 arcsec, which corresponds to 1.3 kpc at the cluster redshift of z = 0.33. The ellipticity of the bulge is e = 0.4 and the inclination angle of the disc is 75 deg. The enclosed mass at distances of ∼ 1 arcsec differs by less than 10% between models that include a disc containing 10% of the mass and models without a disc. Note that, as in the previous figure, the most significant contribution of the disc is in the inner regions. to the light profiles of the galaxies in CL 1358+62, which lie between 0.2 and 4.9 arcsec with only two galaxies hav- ing half-light radii larger than 1.6 arcsec. The dashed lines in Fig. 2 show the surface mass densities for discs of effec- tive radii of 0.05, 0.2 and 0.3 arcsec. The discs are assumed to be inclined at 70 deg to the line of sight. Most of the bulge-plus-disc fits to the light profile of the elliptical galax- ies in the sample have discs with effective radii less than 0.3 arcsec, corresponding to 2 kpc at z = 0.33. The disc com- ponents of the later type galaxies are usually larger, with effective radii of up to 1 arcsec. Comparison of the relative bulge and disc contributions shows that the disc components contribute significantly to the surface mass density within about 0.5 arcsec, or 3.3 kpc. The dotted lines show the to- tal surface mass density for a typical bulge-plus-disc fit, but with different B/T mass ratios. It is clear that the radial profiles for models that include a disc of less than ∼ 5 per cent of the total mass are very similar to bulge-only mod- els. For B/T mass ratios of 0.9 or less, that is for most of the S0/Sa or later type galaxies in Kelson et al., the surface mass density in the inner 0.2 arcsec differs by more than 10 per cent. Gravitational lensing measures directly the projected mass Menc contained inside a circular region of radius R. In Fig. 3 we show Menc as a function of radius R. All the models shown are normalized to the same total mass, but the B/T mass ratio varies. The effective radius of disc and bulge, ellipticity of bulge and inclination angle of disc are rhd = 0.2 arcsec, rh = 0.2 arcsec, e = 0.4 and i = 75 deg respectively and are the same for all models. These values correspond to the typical values for an elliptical galaxy in CL 1358+62. At distances of R ∼ 1 arcsec the enclosed mass for typical bulge-plus-disc models with B/T> 0.9 is only a few percent larger than the enclosed mass for bulge-only models. 6 Ole Moller, P. Hewett & A.W. Blain None of the discs in the light profile fits to the early-type galaxies in CL 1358+62 contain more than ∼ 10 per cent of the light (cf. Table A1), and so, if the mass-to-light ratio is constant, then disc structures are expected to contain at most 10 per cent of the mass. The effect of the disc on the maximum image separations or the radius of the Einstein ring will be small in those cases. More massive discs with a mass of Md ∼ 0.3Mb, contribute significantly to the mass in the inner region of the galaxy and influence the mass contained within the expected Einstein radius more strongly. The properties of all the elliptical galaxies from the Kel- son et al. (2000) catalogue with the reduced disc contribu- tion adopted for our simulations are summarized in Table 1. The table lists both the values for the best fitting bulge-only models and the values for the best bulge-plus-disc fits. The two parameters fe and ft give the B/T mass ratio as mea- sured at the Einstein radius RE and at R → ∞ respectively. The Einstein radius is calculated taking account of the ellip- ticity of the bulge and the inclination of the disc. Note that the Einstein radius does not depend strongly on whether a disc is included or not: there is no systematic dependence of the Einstein radius on the presence of a disc. 4 LENSING PROPERTIES The lensing properties of discs in spiral galaxies have been studied by Maller, Flores & Primack (1997), Moller & Blain (1998) and Blain, Moller, & Maller (1999). In those studies it was found that thin exponential discs can influence the statistical lensing properties even if they only contain 10 per cent of the total mass provided that they are inclined by more than about 70 deg to the line of sight. In particular, the cross section for image geometries with 4 or more magnified images, and the cross section for high magnifications of 10 or more is increased considerably. These results are generic and hold qualitatively for discs in early-type galaxies. How- ever, the scale lengths and masses of the disc components in early-type galaxies are likely to be smaller, relative to the bulge component, and so the effect of the disc on their statis- tical lensing properties is expected to be less. In this section, we describe the general and statistical lensing properties of discy early-type galaxies. 4.1 General lensing properties The gravitational potential of a mass distribution between an observer at z = 0 and a source at z = zs deflects the light of the source, at position ~β on the source plane, so that an image is observed at position ~θ on the lens plane, where ~θ = ~β + DLS DOS ~α. (8) The deflection angle of a light ray α that passes the lens at an angular position ~θ relative to the lens centre, is given by ~α(~θ) = DOS πDLS Z κ(~θ′) 2 d2θ′. ~θ − ~θ′ ~θ − ~θ′(cid:12)(cid:12) (cid:12)(cid:12) (9) where DOS and DLS are the angular diameter distances from observer to source and lens to source respectively. The di- mensionless quantity κ is defined as κ = Σ/Σc where Σc = c2DOS 4πGDOLDLS (10) is the critical surface mass density. These equations show explicitly why thin disc components may be important, even if they contain only a small fraction of the total mass: it is the projected surface mass density that enters equation (9) and this can be very high for thin edge-on discs. The numerical ray-tracing routines that we used to ob- tain the results in this paper are described in detail in Moller & Blain (1998, 2001). The routines were used to generate magnification maps on the source plane and determine the cross sections for high magnifications and multiple imaging. To illustrate the possible effect of disc components in a lens system, Fig. 4 shows the images of a background point source at a redshift of zs = 1.0 that is lensed by a fore- ground elliptical galaxy at redshift zl = 0.3. The images are shown for the bulge-plus-disc and pure-bulge model dis- cussed in § 2. The source position is adjusted slightly, by less than 0.01 arcsec, so that both the total magnification of the source and the image geometries are very similar. Note the change in the relative fluxes of the images. The insets show the magnification maps on the source plane, µ(~θ) for both cases; the disc increases the area enclosed by the caus- tic only slightly. Comparing the magnification maps on the source plane as shown in the insets of Fig. 4 with those of disc lenses shown in Fig. 3 in Moller & Blain (1998), makes clear that the effect of a disc component in early-type lensing galaxies is much less pronounced. However, there are differ- ences in the magnification ratios which we will investigate more in §5. The importance of discs can also be seen when the rel- ative shifts of the image positions for a source at a fixed position is considered. The shift gives some indication of how significant a given modification of a lens model is. We show the image geometries for small, extended sources for a number of our lens models in Fig. 5. There is a small, but measurable change in the image shapes and positions for E/S0 lenses of the order of 0.05 arcsec. In the case of later- type lenses, there is a much more pronounced change; even Sa lens galaxies significantly change the image geometries. Note that there is also a change in the overall scale set by the Einstein radius, of 5−10 per cent. This is consistent with the change of mass projected within the Einstein radius shown in Fig. 3. However, this change of scale only affects the mass normalization and does not reflect an observable effect of the disc component. 4.2 Statistical lensing properties First we determine cross sections for multiple images and high magnifications for ellipticals that include an inclined, thin, exponential disc. In Fig. 6 we show the cross section ratios for bulge and bulge-plus-disc models as a function of magnification, aver- aged over all the galaxies of a certain morphological classifi- cation in the sample. Except for late-type (Sa, Sab and Sb) galaxies, there is no significant increase in the cross sections for magnifications of up to 100. Even though a few early- type galaxies show a clear signature of an enhancement by Discs in early-type lens galaxies 7 Table 1. Elliptical galaxies in CL 1358+62. The values for < Ih >, rh and < Ihd > are taken directly from Kelson et al. (2000). The disc scale length rd is related to the half-light radius as given in Kelson et al. by rd = 1.688rhd. In a cosmology with ΩΛ = 0.7, ΩM = 0.3 and H0 = 50 km s−1 Mpc−1, 1 arcsec corresponds to 6.6 kpc at the cluster redshift of z = 0.33. A larger table that includes all the galaxies in CL 1358+62 investigated by Kelson et al. (2000), is given in the Appendix. ←− Bulge only −→ ←− Bulge and disc No. Type < Ih > rh Σb Re < Ih > rh Σh < Ihd > rd Σd Re fe 212 242 256 303 360 375 409 412 531 534 536 E E E E E E E E E E E 20.780 21.920 20.880 20.590 20.220 22.860 21.270 21.390 21.260 21.210 21.180 0.683 1.529 1.380 0.638 0.341 4.979 0.498 0.767 1.549 0.620 1.266 0.488 0.171 0.445 0.581 0.817 0.072 0.310 0.278 0.313 0.328 0.337 1.220 1.360 2.330 1.270 2.176 2.650 2.176 2.176 2.090 0.860 1.790 21.170 22.420 21.200 21.110 21.660 23.010 21.490 22.320 21.710 22.170 21.520 0.781 1.825 1.551 0.774 0.594 5.267 0.542 1.091 1.830 0.878 1.433 0.340 0.108 0.331 0.360 0.217 0.063 0.254 0.118 0.207 0.136 0.247 18.030 20.720 18.000 18.890 17.620 19.220 22.600 17.660 19.300 16.680 18.570 0.031 0.156 0.053 0.054 0.030 0.096 0.132 0.033 0.108 0.019 0.057 6.138 0.515 6.310 2.780 8.954 2.051 0.091 8.630 1.905 1.281 3.733 1.160 1.270 2.260 1.210 1.000 2.870 0.660 0.870 1.980 0.770 1.720 0.808 0.806 0.836 0.787 0.565 0.800 0.959 0.502 0.772 0.489 0.807 −→ ft 0.960 0.950 0.967 0.947 0.867 0.984 0.969 0.908 0.954 0.901 0.965 the disc component in the magnification map on the source plane, the average effect is small. Rusin & Tegmark (2001) showed that the fraction of 4 image systems found in the CLASS survey is inconsistent with the predictions from standard lens models. They found that under the assumption of dark matter halos that are not significantly more flattened than the visible mass, about 25 per cent of all CLASS lenses should be quads. However, the observed fraction is about 60 per cent. We now investi- gate whether disc structures in the early-type lensing galax- ies could explain this discrepancy. Fig. 7 shows the expected fraction of quad systems as a function of the total magni- fication threshold A for the different morphological types. There is no significant increase in the cross section for the formation of a quad lens for the subset of early-type galaxies; this cross section increases significantly only for the late-type galaxies. Many observed quad lens systems are very symmet- rical. This suggests that the sources do not lie very close to a cusp (B1422+231 is an exception), and so that the sources of the majority of known quad lens systems are likely to be magnified by only moderate amounts µ ∼ 10 − 20. Hence, discs in ellipticals cannot explain the discrepancy between the predicted and observed quad fraction. However, our cal- culations are conservative, as they do not take into account the proper statistical range of lens parameters, source lu- minosity function, the selection function of the survey and the magnification bias. Rusin & Tegmark (2001) show that such factors can increase the predicted fraction of quads at modest magnifications by up to a factor of ∼ 2. Statistical lensing constraints obtained from CLASS by Chae (2002) suggest that the observed fraction of quad lenses is consistent with an average lens ellipticity e ∼ 0.4, and thus that many S0 and later lenses must be present to match this ellipticity. Our results in Fig. 7 suggest that if the majority of lens galaxies are indeed later than S0, then the predicted fraction of quads is consistent with the obser- vations. In summary, these results indicate that the presence of discs with constant mass-to-light ratio in ellipticals, is un- likely to affect the population statistics of strong lens sys- tems. However, inspecting the magnification maps on the source plane shows that discs might have a significant effect for individual lenses (Fig. 5). To investigate this further we determined how much the magnification ratios of individual lens systems are affected when a disc component is included in the lens model. 5 MAGNIFICATION RATIOS Attempts to model a number of known quad lens systems, like B1422+231 and PG1115+080 have shown that smooth power-law ellipsoidal models cannot fit both the observed image positions and magnification ratios. This has been generally interpreted as evidence for substructure in their CDM haloes (Metcalf & Madau 2001; Bradac et al. 2002; Chiba 2002; Dalal & Kochanek 2002; Metcalf & Zhao 2002). The observational detection of substructure in dark matter haloes would be of great importance for the standard model of structure formation, which predicts such CDM substruc- ture on galactic and sub-galactic scales (Moore et al. 1999). If the presence of CDM halo substructure provides the only possible explanation for the observed magnification ratios, this would provide strong observational evidence in support of the current CDM model. We now investigate whether discs in early-type lens galaxies provide an alternative ex- planation. 5.1 Causes for unusual magnification ratios As pointed out by Mao & Schneider (1998), for point sources that are highly magnified and split into 4 magnified images, A, B, C and D, with B the brightest and D the faintest image (cf. Fig. 4 in the previous section; note that we differ here from the more conventional labelling, where images are labelled A,B,C and D in order of decreasing brightness), the individual image magnifications obey the relation µAC/B ≡ µA + µC µB = 1, (11) in the limit of µtot → ∞. The loci of source positions at which µtot is infinite form the caustic line, and so equa- tion (11) holds for sources lying exactly on the caustic. What is not known, however, is how well equation (11) holds for 8 Ole Moller, P. Hewett & A.W. Blain Figure 4. Lensing of a point source by an elliptical galaxy with and without a disc. The contours show the images of a point source at zs = 1 that lies behind a lens galaxy with the same properties as galaxy no. 242 in Table 1, convolved with a Gaus- sian beam of 0.1 arcsec FWHM. The contour levels are 0.5, 3, 10, 20, and 30 in arbitrary flux units. The top panel shows the re- sulting images for a bulge-only model. The bottom panel shows the images for a bulge-plus-disc model, where the disc is inclined at i = 75 deg to the line of sight. The major axis of both bulge and disc is at 45 deg to the x-axis. The insets show the respective 0.2 arcsec × 0.2 arcsec magnification maps on the source plane. The grey-scale is logarithmic and ranges from µ ∼ 2 (white) to µ ∼ 200 (black). The source position is marked by a star in the inset, and is slightly different in the two panels, to keep the image geometry very similar and the total source magnification roughly the same in both cases. sources that lie close to the caustic, or, for sources of small but finite size. In addition, it is not clear how the details of the lensing potential will affect this behaviour close to the caustics. The observed magnification ratios in systems like B1422+231 grossly violate equation (11). This discrepancy is the main reason why the observed magnification ratios are generally regarded as strong evidence for CDM-halo sub- structures with masses of the order of 106 − 107 M⊙, which can strongly affect the flux of individual images if located Figure 5. Lensing of an extended source by galaxies with and without discs. The grey-scale shows the images of a small ex- tended source of size ∼ 0.1 arcsec at zs = 1 that lies behind lens galaxies with the properties of the labelled galaxies in Table A1. The models in the right panels include a disc at an inclination angle of i = 75 deg: the bulge and disc components are oriented at 45 deg with respect to the x-axis in all cases. The left panels show the corresponding images for bulge-only models. close to the images. However, there are several other expla- nations: (i) microlensing. In many strong lens systems, microlens- ing variability has been detected and it is difficult to exclude the possibility of microlensing in the observed systems with certainty. (ii) differential extinction. There is a possibility that op- tical flux ratios of the images are affected by different dust columns along the lines of sight through the lensing galaxy. (iii) the source is extended. Extended radio sources may lead to very different magnification ratios in the observed im- ages, depending on the size and relative brightness of the dif- ferent source components. Recent simulations by Moustakas & Metcalf (2003) suggest that even for extended sources there is no change in the continuum magnification ratios Discs in early-type lens galaxies 9 5.2 Known lens systems with unusual magnification ratios There have been two recent studies of lens systems with unusual magnification ratios. In the comprehensive study by Metcalf & Zhao (2002) 4 systems are listed where the magnification ratios deviate by ∼ 0.1 mag or more from the expected values assuming a one component smooth ellipsoidal lens with external shear; B1422+231, PG1115+080, H1413+117 and Q2237+030. The first three systems are quasars lensed by early-type galaxies while the lens in Q2237+030 is a barred spiral. Of the four systems, H1413+117 lies close to a cluster of galaxies which may cru- cially affect the lensing potential (Kneib et al. 1998; Chae & Turnshek 1999). Q2237+030 is lensed by a galaxy at a very low redshift of z = 0.04 and microlensing affects the flux ra- tios by up to a factor of two (Irwin et al. 1989; Wo´zniak et al. 2000). PG1115+080 and B1422+231 are affected only very moderately if at all by microlensing (Vanderriest et al. 1986; Yee & Bechtold 1996), but both reside very close to, or even within, a galaxy group (Tonry 1998; Kundi´c et al. 1997a,b). As shown by Moller et al. (2002), the presence of the group can significantly alter the lensing behaviour and it is doubt- ful whether a simple power-law-plus-shear model, as used for example by Keeton, Kochanek & Seljak (1997) reflects the true lensing potential adequately. For B1422+231 the predicted and observed magnification ratios differ by up to 0.3 mag, as compared to only 0.1 mag for PG1115+080, and so it is less likely that the discrepancy in B1422+231 is a re- sult of the details of the group mass distribution surrounding the lens. This points out B1422+231 as the strongest candi- date for dark substructures: it is studied in great detail by Bradac et al. (2002). A recent theoretical study (Keeton, Gaudi & Petters 2002) investigated the magnification ratios of four caustic lens systems in more detail, using an analysis similar to that presented here but with a simple one-component lens model: B2045+265, B0712+472, RX J0911+0551 and B1422+231 (Keeton, Gaudi & Petters 2002). RX J0911+0551 has only optical and near-infrared flux measurements, and so the extinction corrections for this source are highly uncertain. B0712+472 is possibly affected by a foreground group (Fass- nacht & Lubin 2002). Keck spectra of B2045+265 indicate that it is likely to be a late-type lens (Fassnacht et al. 1999), and so it is unlikely that a simple one-component model is viable: a disc is almost certainly present. Note that only B1422+231 was included in the study of Metcalf & Zhao (2002). The number of lens systems that are claimed to have magnification ratios that can only be explained using com- pact CDM substructure is rather large. In some cases, mi- crolensing variability or differential extinction could be a viable alternative explanation, but not for all. We summarise the properties of all these lens systems in Table 2. Together with the image fluxes, we list the mag- nification ratios µAC/B and µABC/D ≡ (µA + µB + µC)/µD. 5.3 Magnification ratios for bulge-plus-disc models of early-type lens galaxies Figure 6. The ratio of the magnification cross sections. The curves show the ratio of the cross sections, σb+d(µ > A)/σb(µ > A), for total magnifications µ > A of a point source at zs = 1 as a function of the threshold magnification, A. The different curves are calculated by averaging over all the galaxies in Table A1 that have the same morphological type. When a disc is included, the inclination angle i = 75 deg in all cases. Figure 7. The fraction of four-image lens systems as a function of minimum magnification for different lens models. For each mor- phological class of galaxies listed in Table A1 the average cross section ratio σN=4/σtot is shown, where σN=4 is the cross section for the formation of 4 images and σtot is the total cross section for multiple imaging. The left panel shows the results for earlier types, the right panel for later types. In all bulge-plus-disc mod- els, a disc with the parameters as listed in Table A1 is included, with an inclination angle i = 75 deg. for smooth lens models; only the relative contributions from broad and narrow lines is affected. (iv) off-caustic sources. If the source does not lie exactly on the caustic but is still magnified strongly, then the mag- nification ratios will not necessarily obey equation (11). We now investigate the validity of the last explanation in the context of discs in early-type lensing galaxies. To answer the question whether discs in early-type lenses could explain some or all of the observed magnification ra- 10 Ole Moller, P. Hewett & A.W. Blain Table 2. Summary of observed quad systems with unusual magnification ratios. 1 Christian, Crabtree & Waddell 1987. The fluxes are normalized to image D. 2 Jackson et al. 1998. The listed values are for the Merlin 1996 5 GHz data. 3 Fassnacht et al. 1999. The listed values correspond to the VLA 8.5GHz data. 4Patnaik et al. 1999. The listed values correspond to the VLA 8.5GHz data. 5 Chae & Turnshek 1999. All images have been relabelled to correspond to the convention adopted in other systems. 6 Falco et al. 1996. All images have been relabelled to correspond to the convention adopted in other systems. 7 Burud et al. 1998. The fluxes are normalized to image A. System PG1115+0801 B0712+4722 B2045+2653 B1422+2314 H1413+1175 Q2237+0306 RX J0911+05517 FA FB FC FD µAC/B µABC/D 2.49±0.03 10.5±0.1mJy 18.4±0.3mJy 152±2 1.65±0.16 78±14 µJy 1.0 3.22±0.03 8.5±0.1mJy 9.42±0.3mJy 164±2 1.62±0.16 60±14 µJy 0.965±0.013 0.64±0.03 4.7±0.1mJy 14.8±0.3mJy 81±1 1.0 85±14 µJy 0.544±0.025 1.0 0.9±0.1mJy 2.41±0.3 mJy 5±0.5 0.67±0.1 43±14 µJy 0.458±0.004 0.98±0.1 1.78±0.03 3.5±0.1 1.42±0.03 1.63±0.16 2.7±0.8 1.60±0.08 6.35±0.1 26.3±2.1 17.7±2.1 79.4±7.9 6.1±0.6 5.2±2.4 5.48±0.28 tios, we computed all the image multiplicities, position and magnifications for all sources on a 400 × 400 pixel grid for our particular representative model galaxy (galaxy no. 242). We performed the calculations both on a low-resolution grid that contained the whole caustic structure and on a smaller, higher resolution grid, that contained one of the cusps along the major axis of the lens galaxy. For both grids, we se- lected all sources with quad images, yielding a total sample of about 40000 simulated image systems. Fig. 8 shows a scatter plot of the ratio of the sum of the magnifications of the outer two images to the central image µAC/B, versus the ratio of the total magnification of the brightest three images, A-C to the magnification of the faintest image D, µABC/D. The points in the upper panel are for the pure-bulge model, whereas the points in the lower panel are for the bulge-plus-disc model. For large magnifica- tion ratios there is a significant difference. All the images of sources with high total magnifications, corresponding to a high µABC/D ratio, scatter around µAC/B ∼ 1 for the bulge- only models. For large µ the scatter of a few per cent equals that expected from numerical uncertainties in the simula- tions, due to small deviations in the deflection angle that lead to large deviations in the source position and images occasionally being missed or associated with a wrong source position. Note that there are two distinct and well defined 'populations' of images, corresponding to images in the two different types of cusp regions (see Fig. 9). The two panels of Fig. 9 show quantitatively how much the image magni- fication ratios deviate from equation (11) as a function of source position, both for the bulge-plus-disc model (upper panel) and the bulge-only model (lower panel). These maps are similar overall, but differ significantly in the two cusps that are aligned with the disc: when a disc is present, im- ages of sources that lie very close to these cusps have mag- nification ratios that are inconsistent with the predictions of equation (11). The images that are produced by sources that lie close to these two cusps form the second population of images in the bottom panel of Fig. 8. As an example, Fig. 4 illustrates the behaviour for sources in the lower-left cusp of our model. By comparing the relative fluxes of the images in the top and bottom pan- els of Fig. 4, one can see by eye that the sum of the fluxes of the images A and C is roughly equal to the flux of image B in the bulge only model, but not in the bulge-plus-disc model. The relative positions and flux ratios of the images Figure 8. Scatter diagrams of magnification ratios for quad im- ages of point sources lensed by a model galaxy with parameters matching galaxy no. 242 in Table 1. The markers in the upper panel show the magnification ratios between images produced by the bulge-only model, whereas the markers in the bottom panel show the results for the bulge-plus-disc model. For total magnifi- cations below 20 only 10 per cent of all simulated image systems are shown. Note that the images in the bulge-plus-disc case fall into two regions, whereas all the images produced by pure-bulge models obey equation (11) in the limit of µABC/D → 1. The large markers indicate the magnification ratios for the observed systems with the 1σ error, as listed in Table 2. A small scatter of about 10 per cent is introduced at high µ due to numerical uncertainties. A,B and C in the lower panel of that figure are very close to those observed in B1422+231. Figs 8 and 9 show the difference between the bulge and bulge-plus-disc models. In the bulge-plus-disc model a set of sources exists with an image magnification ratio µAC/B very different from one, even for high magnifications of µABC/D ∼ 50. All these sources lie inside one of the two cusps along the Discs in early-type lens galaxies 11 Figure 10. Scatter diagrams of magnification ratios for quad images of point sources lensed by a model galaxy with parameters as in the Fig. 9, but with a heavier disc that has twice the size. The symbols are in the previous plot. lensed by a spiral galaxy, and so our early-type lens model with a very modest disc would not be expected to provide a good fit. In Fig. 10 we show magnification ratios for about 40000 simulated image systems (without using a second higher- resolution grid) for a bulge-plus-disc model with a disc that has a scale-length twice that of the previous bulge-plus-disc model. All the other parameters are unchanged and corre- spond to galaxy no. 242. The effect of increasing the disc mass is to substantially increase the µAC/B ratio for image systems with magnifications µABC/D ∼ 20. The magnifica- tion ratio of B2045+265 is now consistent with the predicted value. We conclude that the presence of an inclined disc com- ponent containing of order 5 per cent of the total mass can describe all of the known early-type lens systems with un- usual magnification ratios without requiring compact sub- structure in the lens halo. 5.4 Effect of disc size, disc inclination, bulge ellipticity and a dark halo We have used a particular model galaxy (§ 2) as a lens. In order to make sure that our results apply in general to bulge- plus-disc models of early-type galaxies and are not specific to our particular choice of parameters we investigated a range of other models. The full parameter space for bulge-plus-disc models is large and a detailed study is beyond the scope of this paper. We investigated the dependence on several pa- rameters qualitatively. We find that, qualitatively, any disc that on its own produces a caustic of similar or larger size than that produced by the bulge is effective at modifying the image properties. In practice that means any disc that is inclined by more than 60 deg with a scale-length of the order of the size of the caustic of the corresponding bulge- only model. For nearly spherical bulges, discs with smaller scale-lengths are also effective, and naturally produce mag- Figure 9. Maps of the magnification ratio [(µA + µC)/µB − 1]−1 as a function of source position. The upper panel is for the bulge- only model, the lower panel for the bulge-plus-disc model. Large values (dark shades) indicate that the magnification ratios of the images of a source at that position obey equation (11). Small val- ues (light shades) indicate that the image magnification ratios deviate from that relation. Note that Fig. 4 shows an example configuration for sources lying in the bottom left cusp. major axis. Note that for all those images, the faintest image D lies at a large distance of ∼ 1.2 arcsec from the centre of the lensing galaxy and is not demagnified. This reflects the high total magnification of a source which lies close to the caustic line. Some unusual observed magnification ratios of some ob- served lens systems are indicated by the large symbols in Fig. 8. Error bars indicate the 1-σ flux measurement errors (Table 2). Only H1413+117 and RXJ0911+0551 lie in re- gions of the diagram that are consistent with a bulge-only model. However, the bulge-plus-disc model is clearly consistent with the magnification ratios for PG1115+080, H1413+117, RX J0911+0551, B0712+472, Q2237+030 and B1422+231. B2045+265 is the only system for which our model cannot reproduce the observed magnification ratios. This source is 12 Ole Moller, P. Hewett & A.W. Blain nification ratios similar to those observed for B1422+231. For disc sizes below 0.2 arcsec, the disc inclination must be increased beyond i = 80 deg to obtain magnification ratios consistent with B1422+231. Even an edge-on disc smaller than 0.1 arcsec cannot reproduce the magnification ratios of B1422+231. There is also the possibility of misalignment between the positions of the disc and the bulge/halo. Such misalignments are to be expected as a result of recent merger events and may lead to complicated and unusual lensing con- figurations and provide more extreme changes in the magni- fication ratios, as discussed by Quadri, Moller & Natarajan (2002). We have also investigated models with a CDM halo. Adding a singular CDM halo, containing 2/3 of the total mass within an Einstein radius, and adjusting the mass nor- malization of the bulge and disc so as to keep the Einstein radius constant, does not significantly alter the effect of the disc. It is purely the presence of a thin disc structure in combination with a more spherical mass component that produces the change in magnification ratios. 5.5 Expected number of affected systems Moller & Blain (1998) have shown that discs in spiral galax- ies are important for the statistical effects on lensing cross sections only at inclination angles of ≥ 60 deg. Assuming that a fraction of 50 − 100 per cent of all early-type lens galaxies do contain discs, 15 − 35 per cent will contain discs with inclination angles of ≥ 60 deg. A conservative estimate of the fraction of early-type lens systems in which discs may be important is thus ∼ 10 − 30 per cent. This fraction is consistent with the number of CLASS quad lens systems with deviant magnification ratios: of the 7 CLASS quad lenses, at least two early-type lens system (B1422+231 and B0712+472) exhibit magnification ratios inconsistent with simple power-law plus external-shear models. An additional factor to consider is magnification bias. Depending on the bulge-to-disc ratio, the disc may lead to higher cross sections for strong magnifications, as shown in Fig. 6. This could bias a flux-limited sample towards lens galaxies that contain discs. However, as shown in Fig. 6, the cross section is not increased significantly for E/S0/Sa galax- ies, and so the magnification bias is likely to be small. 5.6 Distinguishing the effect of discs and CDM substructure The results described in this section demonstrate that discs in early-type lensing galaxies may explain some of the ob- served magnification ratios that have hitherto only been ex- plained assuming CDM substructure. There are some key observational differences between these two scenarios. First, discs in early-type lens galaxies are, in principle, detectable by bulge-plus-disc light profile decompositions as discussed in § 2. However, the light of the background quasar makes accurate photometry of lens galaxies extremely dif- ficult in practice. Treu & Koopmans (2002) obtained kine- matic information of a lensing galaxy from spectroscopic ob- servations, and demonstrated that it is thus possible to ob- tain additional constraints on its mass distribution. It might be possible to constrain possible disc structures in lensing galaxies in this way. Secondly, disc structures significantly affect only the magnification ratios of sources that lie within one of the two cusps along the discs. Assuming that discs are oriented along the major axis of the bulge ellipticity, this means that disc structures provide an explanation of discrepant magnifica- tion ratios only for quad image systems with three strongly magnified images along a line perpendicular to the major axis of the lens galaxy, whereas CDM substructure could modify any image configuration. Note that this is not true for late-type lenses with extended inclined discs, in which the caustic structures are completely dominated by the disc, and the disc strongly affects the magnification ratios of quad systems produced by sources in any one of the four cusps. We stress that these results do not prove the existence of discs in early-type lensing galaxies, nor do we propose that such disc structures provide the only explanation for all the observed magnification ratios. For all systems, the observed magnification ratios could be caused by CDM sub- structures. We have not carried out detailed modelling of any of the observed systems, and our analysis does not include joint fitting to both magnification ratios and image posi- tions. However, in contrast to Keeton et al. (2002) we find that magnification ratios alone do not provide conclusive ev- idence for the presence of CDM halo substructure: realistic disc structures can explain the observed magnification ra- tios equally well. Detailed case-by-case lens modelling using bulge-plus-disc models is now needed to verify that early- type galaxies with disc structures are indeed viable models for all or at least some of the observed lens systems. The fact that discs affect image positions very slightly, even for a fixed source position, indicate that detailed bulge-plus-disc mod- els are likely to be successful in fitting magnification ratios and image positions. 6 EFFECT ON HUBBLE PARAMETER ESTIMATION Strong gravitational lensing of variable background quasars provides a unique method to determine the Hubble constant. Brightness fluctuations are observed at different times at each image as the path lengths and gravitational time delays to each image depend on its relative position and the local lensing potential. If the image positions and lensing poten- tial are known, then H0 can be determined, as ∆t ∝ 1/H0. Gravitational and geometric time delays are often of roughly equal magnitude, and so an accurate knowledge of the lens- ing potential is crucial to deduce an accurate value of H0. Lens systems with measured time delays are nearly always associated with early-type lens galaxies, although B1600+434 (Koopmans et al. 2000) and B0218+357 (Wuck- nitz, priv. comm.) are notable exceptions. These systems have been modelled using bulge-only models, with a con- stant external shear in some cases. We now investigate how a disc component is likely to affect time delays for the bulge and bulge-plus-disc model of galaxy no. 242. The time delay between two images at ~θ1 and ~θ2 of a source at ~β is ∆t = DOSDOL cDLS (1+zl)(cid:20) (~β − ~θ1)2 − (~β − ~θ2)2 2 − ∆ψb + ∆ψd(cid:21) ,(12) where DOL is the angular diameter distances between ob- Discs in early-type lens galaxies 13 Figure 12. Histogram of time delays. This plot shows the ex- pected relative fraction of systems as a function of timedelay be- tween images B and D. The left panel shows the statistics for the regions marked I in Fig. 11 and the right panel the statis- tics for the regions marked II. The histograms are shown for our bulge-plus-disc and bulge-only lens galaxy model (see § 2). Note that the time delay is smaller for bulge-only systems, due to the shallower potential. Figure 13. Histogram of the separations of images. This plot shows the expected relative fraction of systems as a function of separation. Note that the separation is larger for bulge-only sys- tems, due to the higher mass in the central region. to unity. Comparing the histograms in the right panel of Fig. 12, there is no significant difference between the bulge- plus-disc and bulge-only model time-delay statistics in the cusps away from the disc (the regions marked II on Fig. 11). However, the histograms in the left panel of Fig. 12 show that the time delays in the cusps lying along the disc are in- creased by ∼ 20 per cent for the bulge-plus-disc model with respect to the bulge-only model. The overall distribution for the bulge only model is shifted towards shorter time delays with respect to the bulge-plus-disc models. Assuming that an elliptical lens galaxy that contains a disc is modelled us- ing a bulge-only model, this suggests that the Hubble con- stant will be underestimated systematically. In Fig. 13 we show explicitly that disc components al- ter the individual image positions only very slightly for our model galaxy no. 242. The figure shows the distribution of Figure 11. Time delay as function of source position. The two maps show the time delays between images B and D (cf. Fig. 4) for all quad geometries for model galaxy no.242. The left panel shows the results for the bulge-only model and the right panel the results for the bulge-plus-disc model. The boxes in the left panel mark the different cusp regions, as discussed in the text. server and lens. The distances and thus ∆t are proportional to H −1 0 . The lens redshift is zl and the potentials due to bulge and disc components are ψb and ψd respectively. In § 4 we pointed out that the shift in image positions between bulge-plus-disc and bulge-only models for E and E/S0 type lensing galaxies is expected to be very small. We are concerned here with the effects of modelling an early-type lens that contains a disc with a bulge-only model. Since the lens models are obtained from fits to the image positions, the image positions remain the same irrespective of the model used. Therefore, the geometrical time delay of a given pair of images is not affected by a disc component; any change in the time delay is due to the change in the potential at the image positions. Note that compact CDM substructure does not affect time-delays measurably. 6.1 Four-image/quad systems About half of the known lens systems with measured time delays are quad systems. Since the quad caustic structure is the region in the source plane that is most strongly affected by the presence of a disc, the greatest effect of a disc on time delays is expected in these systems. In most cases a single time delay is measured, usually between the image pair with the largest separation. Using our ray-tracing code in combination with equation (12) we calculated the time delays between images B and D for all quad systems of the model galaxy described in § 2. Fig. 11 shows the time delays as a function of source position for bulge-plus-disc and bulge- only models. The difference is on the order of a few tens of per cent, and is largest in the cusps lying along the disc (the regions marked I). Especially in the cusp regions close to the caustic, where the image magnification is greatest, the time delays for the bulge-plus-disc model are 20−30 per cent larger than for the bulge-only model. The time-delay statistics for regions I and II are shown in Fig. 12. For each region in Fig. 11 the time delays are divided into 50 bins and the histograms are normalized 14 Ole Moller, P. Hewett & A.W. Blain Figure 14. Histogram of time delays for 2-image systems. This plot shows the expected relative fraction of systems as a function of time-delay between the two images in 2-image lens systems. The histograms are shown for our bulge-plus-disc and bulge-only lens galaxy model as described in § 2. image separations for both models. The images are system- atically further apart in the bulge-only models, suggesting a larger mass inside the Einstein radius and thus a larger pre- dicted time delay, if the radial profile of the potential is unaf- fected. In observed lens systems, the lens would be modelled to fit the image configurations, and so a bulge-only model fitted to the observed image separations would contain less mass within the Einstein radius than the bulge-only model with the time delay distribution shown in Fig. 12. This would shift the time-delay distributions shown in Fig. 12 for the bulge-only model to even lower time delays. The only pos- sibility both to fit the image separations and obtain similar distribution of time delays is by changing the gradient of the potential between the images. For point sources, the only other lensing property that is sensitive to the shape of the potential is the magnification ratio between the im- ages. Thus, if magnification ratios are not used to constrain lens models, then bulge-only models might fit the image po- sitions of a lens that actually consists of a bulge-plus-disc lens, yielding an incorrect mass model and an H0 value that is systematically too low by up to 25 per cent. 6.2 Two-image systems Quad lens systems are generally regarded as being more useful than 2-image systems for H0 determinations because more constraints on the lensing potential are available. How- ever, many 2-image lens systems have been used for time de- lay measurements. In these cases, the source lies outside the diamond-shaped caustic, where magnifications µ ∼ 2 − 10. These images are expected to be affected only weakly by a disc component. We have calculated the time delays for 2-image systems and show the corresponding statistics in Fig. 14. Comparison with Fig. 12 shows that the effect of a disc component on the time delay is much smaller for 2-image systems than for quad systems. A number of 2-image systems with measured time delays, like B0218+357 (Biggs et al. 1999) have possible disc structures, but even if these are present, our results indicate that the the effect on estimates of H0 is small. Note that recent observations (Wucknitz, private comm.) indicate that the lens in B0218+357 is a spiral galaxy. Massive discs, as expected in spiral lenses, may affect the time delays much more strongly. 6.3 Parameter dependence The results in this and the preceding section show that any inclined discs within elliptical lens galaxies affect sources in the cusps lying along the disc most strongly, whereas the effect on the cusps lying away from the disc is small. We have not checked the dependence of the time-delay statistics on the disc parameters in detail, but the dependence is similar to that discussed in § 5.4 since magnification ratios and time delays are both sensitive on changes in the cusps. The results we obtained in that section also hold for the time delays. Disc structures are important if they are inclined at i > 60 deg to the line of sight and are massive/extended enough to produce a caustic of a size that is of the order of the caustic produced by the bulge. That is typically the case for discs larger than rhd > 0.6kpc and with masses of at least ∼ 5 per cent of the bulge, for which time delays will be increased by a few tens of per cent relative to bulge-only models. 6.4 The effect on estimates on H0 Our approach here only gives a rough indication of how de- terminations of H0 are affected by disc structures. To be any more quantitative about the uncertainties that arise it is necessary to perform detailed modelling of known systems using bulge and bulge-plus-disc models, which is difficult and time-consuming. However, our results offer a possible reso- lution as to why there is a significant discrepancy between determinations of H0 from CMB data (Lewis & Bridle 2002), the HST Key-project (Freedman et al. 2001) and gravita- tional lensing (Kochanek 2003). The effect of disc structures on time-delays is much more significant for quad systems than for 2-image sys- tems. In addition, we expect that there will be a correla- tion between low-H0 results and unusual magnification ra- tios if discs are present. This seems to be the case: Biggs et al. (1999) derive H0 = 69+13 −19 km s−1 Mpc−1 for the 2- image system B0218+357, whereas Impey et al. (1998) ob- tain H0 = 44±4 km s−1 Mpc−1 for PG1115+080, a quad sys- tem with unusual magnification ratios. The errors on the H0 estimate by Biggs et al. (1999) have recently been disputed by Lehar et al.(2000) and Wucknitz, Biggs & Browne (MN- RAS, submitted). More recent single-component models of PG1115+080 by Treu & Koopmans (2002) give a higher value of H0 = 59+12 −7 ± 3 km s−1 Mpc−1. The larger errors are statistical, and the smaller error of ±3 is an estimate of the systematic error. No other quad systems with measured time delays have unusual magnification ratios, and the values of H0 determined from these systems are generally higher. 7 CONCLUSIONS Gravitational lensing plays a role of ever increasing impor- tance in determining cosmological parameters, notably H0, and constraining the dark matter distribution on a variety of scales. More than 50 strong gravitational lens systems are known. Generally, it is not possible to constrain the mass distribution of the lensing galaxy uniquely, due to the de- generacies present. Whether this uncertainty in models has serious implications for cosmological parameter estimation from strong lensing is still unclear. The majority of all lensing galaxies have been classified as massive early-type galaxies. Reliable photometric deter- mination of the presence of a disc component is difficult due to the usually very bright lensed quasar images. Therefore, discs similar to those discussed in this paper would not gen- erally have been identified in existing images. However, ob- servations of early-type galaxies in the local universe and in the moderate-redshift cluster CL 1358+62 suggest strongly that a large fraction of early-type galaxies contain disc-like structures. We have investigated how the lensing properties of early-type galaxies would change if a thin disc component that contains about 5 per cent of the mass is included: (i) If early-type galaxies contain disc components, the sta- tistical lensing properties of early-type lensing galaxies are affected by less than about 5 per cent. Only if the major- ity of lensing galaxies are of S0/Sa or later types would the expected relative number of quad to 2-image lens systems increase by a factor of 2 or more. (ii) A disc component in early-type lensing galaxies af- fects significantly the expected magnification ratios of highly magnified quad systems. If a disc component is present in the lens and inclined at more than ∼ 70 deg to the line of sight, the resulting image configurations and magnification ratios can resemble that of systems like B1422+231. (iii) Time delays may be affected significantly by the pres- ence of a disc component in individual early-type lens galax- ies. Bulge-only lens models used to fit bulge-plus-disc lenses fit lensed image positions well, but would yield a value for H0 that is systematically low by about 25 per cent. The presence of inclined exponential discs in early-type lensing galaxies affect the lensing properties of quad lens sys- tems significantly. Assuming that the majority of all early- type lens galaxies contain discs, then both the time delays and magnification ratios are expected to be affected by disc components in 10 − 30 per cent of all quad lens systems. One or two of the quad lenses identified by CLASS is expected to be affected (consistent with the properties of B1422+231). Bulge-plus-disc models provide an alternative explanation to the observed magnification ratios and lead to systemat- ically longer time-delays. In order to obtain more stringent constraints on the presence of CDM substructure in the lens and on the effect of discs on H0 measurements, detailed lens modelling of known lens systems will be necessary. Fitting models that include exponential discs to real lens systems is much more computationally intensive than the calculations performed here, but may be possible. We have discussed the first order effect of discs in early- type galaxies on observed lensing properties. Once a large number of lens systems is known, it might also be possible to use lensing to constrain the masses and parameters of those systems. Determining the abundance and properties of discs in ellipticals in more detail, and at different redshifts, using Discs in early-type lens galaxies 15 gravitational lensing would shed new light on the details of the formation of discs in early-type galaxies. ACKNOWLEDGEMENTS OM acknowledges financial support from a European Com- munity Marie Curie Fellowship. We thank Konrad Kuijken and the referee for helpful comments on the manuscript. REFERENCES Abraham R. G., van den Bergh S., 2001, Science, 293, 1273 Barnes J. E., 2002, MNRAS, 333, 481 Bender R., Surma P., Doebereiner S., Moellenhoff C., Madejsky R., 1989, A&A, 217, 35 Bennett C. L., et al., 2003, ApJ, submitted, astro-ph/0302207. Benjamin R. A., Danly L., 1997, ApJ, 481, 764 Biggs A. D., Browne I. W. A., Helbig P., Koopmans L. V. E., Wilkinson P. N., Perley R. A., 1999, MNRAS, 304, 349 Binney J., Tremaine S., 1987, "Galactic dynamics". Princeton, NJ, Princeton University Press, 1987, p747 Blain A. W., Moller O., Maller A. H., 1999, MNRAS, 303, 423 Bradac M., Schneider P., Steinmetz M., Lombardi M., King L. J., Porcas R., 2002, A&A, 388, 373 Burud I., Courbin F., Lidman C., Jaunsen A. O., Hjorth J., Os- tensen R., Andersen M. I., Clasen J. W., et al., 1998, ApJ, 501, L5 Byun Y. I., Freeman K. C., 1995, ApJ, 448, 563 Chae K., Turnshek D. A., 1999, ApJ, 514, 587 Chae K.-H., 2002, MNRAS, submitted (astro-ph/0211244) Chiba M., 2002, ApJ, 565, 17 Christian C. A., Crabtree D., Waddell P., 1987, ApJ, 312, 45 Dalal N., Kochanek C. S., 2002, ApJ, 572, 25 de Zeeuw P. T., Bureau M., Emsellem E., Bacon R., Marcella Carollo C., Copin Y., Davies R. L., Kuntschner H., et al., 2002, MNRAS, 329, 513 Falco E. E., Lehar J., Perley R. A., Wambsganss J., Gorenstein M. V., 1996, AJ, 112, 897 Fassnacht C. D., et al., 1999, AJ, 117, 658 Fassnacht C. D., Lubin L. M., 2002, AJ, 123, 627 Ferguson A. M. N., Clarke C. J., 2001, MNRAS, 325, 781 Freedman W. L., et al., 2001, ApJ, 553, 47 Impey C. D., Falco E. E., Kochanek C. S., Leh´ar J., McLeod B. A., Rix H.-W., Peng C. Y., Keeton C. R., 1998, ApJ, 509, 551 Irwin M. J., Webster R. L., Hewett P. C., Corrigan R. T., Jedrze- jewski R. I., 1989, AJ, 98, 1989 Jackson N., et al., 1998, MNRAS, 296, 483 Keeton C. R., Kochanek C. S., Seljak U., 1997, ApJ, 482, 604 Keeton C. S., Gaudi B. S., Petters A. O., 2002, ApJ, submitted (astro-ph/0210318) Kelson D. D., Illingworth G. D., van Dokkum P. G., Franx M., 2000, ApJ, 531, 137 Kneib J. P., Alloin D., Mellier Y., Guilloteau S., Barvainis R., Antonucci R., 1998, A&A, 329, 827 Kochanek C. S., 2003, ApJ, 583, 49 Koopmans L. V. E., de Bruyn A. G., Jackson N., 1998, MNRAS, 295, 534 Koopmans L. V. E., de Bruyn A. G., Xanthopoulos E., Fassnacht C. D., 2000, A&A, 356, 391 Koopmans L. V. E., Treu T., 2003, ApJ, 583, 606 Kundi´c T., Hogg D. W., Blandford R. D., Cohen J. G., Lubin L. M., Larkin J. E., 1997a, AJ, 114, 2276 16 Ole Moller, P. Hewett & A.W. Blain Kundi'c T., Turner E. L., Colley W. N., Gott J. R. I., Rhoads J. E., Wang Y., Bergeron L. E., Gloria K. A., et al., 1997b, ApJ, 482, 75 Lewis A., Bridle S. L., 2002, Phys. Rev. D, 66, 10 Leh´ar J., Falco E. E., Kochanek C. S., McLeod B. A., Munoz J. A., Impey C. D., Rix H.- W., Keeton C. R., et al., 2000, ApJ, 536, 584 Maller A. H., Flores R. A., Primack J. R., 1997, ApJ, 486, 681 Mao S., Mo H. J., White S. D. M., 1998, MNRAS, 297, L71 Mao S., Schneider P., 1998, MNRAS, 295, 587 Metcalf R. B., Madau P., 2001, ApJ, 563, 9 Metcalf R. B., Zhao H., 2002, ApJ, 567, L5 Mo H. J., Mao S., White S. D. M., 1998, MNRAS, 295, 319 Moller O., Blain A. W., 1998, MNRAS, 299, 845 Moller O., Blain A. W., 2001, MNRAS, 327, 339 Moller O., Natarajan P., Kneib J., Blain A. W., 2002, ApJ, 573, 562 Moriondo G., Giovanardi C., Hunt L. K., 1998, A&A, 319, 409 Moore B., Ghigna S., Governato F., Lake G., Quinn T., Stadel J., Tozzi P., 1999, ApJ, 524, L19 Moustakas L. A., Metcalf R. B., 2003, MNRAS, 339, 607 Naab T., Burkert A., 2001, ApJ, 555, L91 Natarajan P., Kneib J.-P., Smail I., Ellis R. S., 1998, ApJ, 499, 600 Patnaik A. R., Kemball A. J., Porcas R. W., Garrett M. A., 1999, MNRAS, 307, L1 Putman M. E., et al., 2002, AJ, 123, 873 Quadri R., Moller O., Natarajan P., 2002, ApJ, submitted Rest A., van den Bosch F. C., Jaffe W., Tran H., Tsvetanov Z., Ford H. C., Davies J., Schafer J., 2001, AJ, 121, 2431 Rix H.-W., White S. D. M., 1990, ApJ, 362, 52 Rix H.-W., 1991, Ph.D. Thesis Romanowsky A. J., Kochanek C. S., 1997, MNRAS, 287, 35 Rusin D., Tegmark M., 2001, ApJ, 553, 709 Saha A., Sandage A., Tammann G. A., Dolphin A. E., Christenses J., Panagia N., Macchetto F. D., 2001, ApJ, 562, 314 Scorza C., Bender R., 1995, A&A, 293, 20 Scorza C., Bender R., Winkelmann C., Capaccioli M., Macchetto D. F., 1998, A&A Supp, 131, 265 Simard L., Koo D. C., Faber S. M., Sarajedini V. L., Vogt N. P., Phillips A. C., Gebhardt K., Illingworth G. D., et al., 1999, ApJ, 519, 563 Somerville R. S., Primack J. R., 1999, MNRAS, 310, 1087 Steinmetz M., Muller E., 1995, MNRAS, 276, 549 Tonry J. L., 1998, AJ, 115, 1 Tran H. D., Tsvetanov Z., Ford H. C., Davies J., Jaffe W., van den Bosch F. C., Rest A., 2001, AJ, 121, 2928 Tran K.-V., Simard L., Illingworth G., Franx M., 2003, ApJ, in press, astro-ph/0302292 Treu T., Koopmans L. V. E., 2002, MNRAS, 337, 6 van Dokkum P. G., Stanford S. A., 2001, ApJ, 562, L35 Vanderriest C., Wlerick G., Lelievre G., Schneider J., Sol H., Horville D., Renard L., Servan B., 1986, A&A, 158, L5 Wo´zniak P. R., Alard C., Udalski A., Szyma´nski M., Kubiak M., Pietrzy´nski G., Zebru´n K., 2000, ApJ, 529, 88 Yee H. K. C., Bechtold J., 1996, AJ, 111, 1007 APPENDIX A: GALAXIES IN CL 1358+62 All the galaxies for which Kelson et al. (2000) provides sur- face brightness profile information are employed in our cal- culation of certain statistical properties, like the high magni- fication cross sections in §4.2. For completeness, we present here the surface brightness profile parameters for all the galaxies. Discs in early-type lens galaxies 17 ←− Bulge only −→ ←− Bulge and disc No. Type < Ih > rh Σb Re < Ih > rh Σb < Ihd > rd Σd Re fe 212 242 256 303 360 375 409 412 531 534 536 233 269 309 353 381 493 095 135 182 211 215 236 298 300 343 359 408 410 463 481 110 129 142 164 292 335 366 369 397 423 440 454 523 209 328 356 368 371 372 465 549 234 E E E E E E E E E E E E/S0 E/S0 E/S0 E/S0 E/S0 E/S0 S0 S0 S0 S0 S0 S0 S0 S0 S0 S0 S0 S0 S0 S0 S0/a S0/a S0/a S0/a S0/a S0/a S0/a S0/a S0/a S0/a S0/a S0/a S0/a Sa Sa Sa Sab Sa Sa Sa Sab Sb 20.780 21.920 20.880 20.590 20.220 22.860 21.270 21.390 21.260 21.210 21.180 20.400 20.390 20.020 21.050 20.340 19.980 20.940 20.450 20.460 20.120 20.660 21.100 19.690 19.840 20.540 20.260 19.820 20.670 20.430 20.250 19.450 19.930 21.400 22.620 20.320 20.860 20.790 21.860 20.500 21.610 22.450 20.420 21.640 20.650 20.720 20.200 21.100 21.040 21.920 21.010 21.920 21.860 0.683 1.529 1.380 0.638 0.341 4.979 0.498 0.767 1.549 0.620 1.266 0.784 0.962 0.502 1.120 0.512 0.193 0.704 0.418 0.355 0.448 0.457 0.776 0.554 0.453 0.404 0.574 0.382 0.488 0.639 0.260 0.350 0.489 0.834 2.251 0.486 0.539 0.630 1.036 0.409 1.021 1.108 0.595 1.044 0.625 0.666 0.604 0.542 0.956 1.435 0.609 0.866 0.996 0.488 0.171 0.445 0.581 0.817 0.072 0.310 0.278 0.313 0.328 0.337 0.692 0.698 0.982 0.380 0.731 1.019 0.421 0.661 0.655 0.895 0.545 0.363 1.330 1.159 0.608 0.787 1.180 0.540 0.673 0.794 1.660 1.067 0.275 0.090 0.745 0.453 0.483 0.180 0.631 0.227 0.105 0.679 0.221 0.550 0.515 0.832 0.363 0.384 0.171 0.394 0.171 0.180 1.220 1.360 2.330 1.270 2.176 2.650 2.176 2.176 2.090 0.860 1.790 1.740 2.150 1.370 1.710 2.176 0.540 1.150 0.900 0.760 1.160 2.176 1.150 1.810 2.176 0.830 1.380 1.160 0.930 2.176 2.176 1.300 1.400 1.030 1.250 1.130 0.920 1.120 0.960 0.860 1.110 0.690 1.310 1.110 1.200 1.230 1.500 2.176 1.470 1.280 0.950 0.770 0.920 21.170 22.420 21.200 21.110 21.660 23.010 21.490 22.320 21.710 22.170 21.520 20.940 20.620 20.210 21.240 20.500 21.660 21.840 20.870 20.770 20.410 21.770 21.630 20.840 20.530 21.100 21.050 20.660 21.130 20.740 20.760 18.360 20.950 20.470 22.920 20.390 20.900 21.160 21.980 18.960 22.460 22.980 22.270 21.960 22.060 21.530 19.780 22.390 20.640 22.720 21.220 19.900 17.800 0.781 1.825 1.551 0.774 0.594 5.267 0.542 1.091 1.830 0.878 1.433 0.952 1.045 0.540 1.205 0.550 0.374 0.986 0.490 0.392 0.458 0.700 0.944 0.835 0.584 0.492 0.769 0.523 0.575 0.714 0.157 0.102 0.726 0.318 2.541 0.297 0.541 0.718 1.085 0.111 1.397 1.328 0.990 1.171 0.960 0.862 0.392 0.627 0.468 1.905 0.660 0.142 0.035 0.340 0.108 0.331 0.360 0.217 0.063 0.254 0.118 0.207 0.136 0.247 0.421 0.565 0.824 0.319 0.631 0.217 0.184 0.449 0.492 0.685 0.196 0.223 0.461 0.614 0.363 0.380 0.545 0.353 0.506 0.497 4.529 0.417 0.649 0.068 0.698 0.437 0.344 0.161 2.606 0.104 0.064 0.124 0.164 0.150 0.244 1.225 0.111 0.555 0.082 0.325 1.096 7.586 18.030 20.720 18.000 18.890 17.620 19.220 22.600 17.660 19.300 16.680 18.570 18.740 18.710 18.910 19.970 20.020 18.740 17.750 18.010 21.910 21.680 16.950 17.870 19.530 18.660 20.930 19.050 18.860 20.840 21.180 20.650 20.120 16.680 22.440 17.770 21.100 20.610 20.460 20.280 21.450 17.950 18.800 20.610 21.080 20.930 21.300 22.070 21.560 21.920 21.990 21.100 22.500 22.030 0.031 0.156 0.053 0.054 0.030 0.096 0.132 0.033 0.108 0.019 0.057 0.067 0.055 0.036 0.079 0.052 0.035 0.038 0.023 0.130 0.256 0.022 0.033 0.142 0.055 0.098 0.072 0.057 0.093 0.136 0.166 0.248 0.028 0.648 0.036 0.332 0.035 0.083 0.047 0.320 0.043 0.036 0.273 0.117 0.253 0.232 0.533 0.305 0.693 0.351 0.079 0.609 0.621 6.138 0.515 6.310 2.780 8.954 2.051 0.091 8.630 1.905 21.281 3.733 3.192 3.281 2.729 1.028 0.982 3.192 7.943 6.252 0.172 0.213 6.596 7.112 1.542 3.436 0.425 2.399 2.858 0.461 0.337 0.550 0.895 21.281 0.106 7.798 0.363 0.570 0.655 0.773 0.263 6.607 3.020 0.570 0.370 0.425 0.302 0.149 0.238 0.171 0.160 0.363 0.100 0.154 1.160 1.270 2.260 1.210 1.000 2.870 0.660 0.870 1.980 0.770 1.720 1.640 2.100 1.000 1.680 1.000 0.480 1.040 0.870 0.740 1.120 0.790 1.080 1.630 1.280 0.780 1.270 1.080 0.880 1.350 1.000 1.000 1.300 0.870 1.230 1.000 0.910 1.070 0.950 1.000 1.000 0.640 1.100 1.070 1.040 1.130 1.420 1.000 1.260 1.150 1.000 0.560 0.520 0.808 0.806 0.836 0.787 0.565 0.800 0.959 0.502 0.772 0.489 0.807 0.796 0.905 0.913 0.902 0.924 0.549 0.583 0.817 0.941 0.921 0.524 0.711 0.735 0.789 0.879 0.748 0.755 0.886 0.926 0.521 0.722 0.647 0.934 0.616 0.803 0.963 0.882 0.903 0.841 0.515 0.572 0.646 0.882 0.755 0.862 0.939 0.713 0.890 0.877 0.925 0.938 0.903 −→ ft 0.960 0.950 0.967 0.947 0.867 0.984 0.969 0.908 0.954 0.901 0.965 0.947 0.976 0.979 0.979 0.979 0.838 0.912 0.956 0.945 0.873 0.889 0.946 0.873 0.932 0.935 0.923 0.915 0.951 0.965 0.351 0.363 0.899 0.496 0.966 0.506 0.992 0.963 0.987 0.443 0.918 0.950 0.655 0.968 0.772 0.881 0.749 0.568 0.497 0.909 0.977 0.284 0.094 Table A1. Galaxies in CL 1358+62. The values for < Ib >, rb and < Ihd > are taken directly from Kelson et al. (2000). The disc scale length rd is related to the half-light radius as given in Kelson et al. by rd = 1.688rhd. In a cosmology with ΩΛ = 0.7, ΩM = 0.3 and H0 = 50 km s−1 Mpc−1, 1 arcsec corresponds to 6.6 kpc at a redshift of z = 0.33.
0802.2070
1
0802
2008-02-14T18:12:22
Ages and metallicities of circumnuclear star formation regions from Gemini IFU observations
[ "astro-ph" ]
Aims: We derive the age and metallicity of circumnuclear star formation regions (CNSFRs) located in the spiral galaxies NGC6951 and NGC1097, and investigate the cause of the very low equivalent widths of emission lines found for these regions. Methods: We used optical two-dimensional spectroscopic data obtained with Gemini GMOS-IFUs and a grid of photoionization models to derive the the metallicities and ages of CNSFRs. Results: We find star formation rates in the range 0.002-0.14 Msun/yr and oxygen abundance of 12+log(O/H)~8.8 dex, similar to those of most metal-rich nebulae located in the inner region of galactic disks. Conclusions: We conclude that the very low emission-line equivalent widths observed in CNSFRs are caused by the ``contamination'' of the continuum by (1) contribution of the underlying bulge continuum combined with (2) contribution from previous episodes of star formation at the CNSFRs.
astro-ph
astro-ph
Astronomy&Astrophysicsmanuscript no. dors December 5, 2018 c(cid:13) ESO 2018 8 0 0 2 b e F 4 1 ] h p - o r t s a [ 1 v 0 7 0 2 . 2 0 8 0 : v i X r a Ages and metallicities of circumnuclear star formation regions from Gemini IFU observations O. L. Dors Jr., T. Storchi-Bergmann, Rogemar A. Riffel, A. A Schimdt Universidade Federal do Rio Grande do Sul, IF, CP 15051, Porto Alegre 91501-970, RS, Brazil. Received 29 octubre 2007 / Accepted 28 january 2008 ABSTRACT Aims. We derive the age and metallicity of circumnuclear star formation regions (CNSFRs) located in the spiral galaxies NGC 6951 and NGC 1097, and investigate the cause of the very low equivalent widths of emission lines found for these regions. Methods. We used optical two-dimensional spectroscopic data obtained with Gemini GMOS-IFUs and a grid of photoionization models to derive the the metallicities and ages of CNSFRs. Results.We find star formation rates in the range 0.002-0.14 M⊙ yr−1 and oxygen abundance of 12+log(O/H)≈8.8 dex, similar to those of most metal-rich nebulae located in the inner region of galactic disks. Conclusions. We conclude that the very low emission-line equivalent widths observed in CNSFRs are caused by the "contamination" of the continuum by (1) contribution of the underlying bulge continuum combined with (2) contribution from previous episodes of star formation at the CNSFRs. Key words. Galaxy: abundances -- Galaxies: individual: NGC 6951, NGC 1097 -- ISM: abundances -- ISM: HII regions 1. Introduction The pioneer works by Morgan (1958) and S´ersic & Pastoriza (1967) have shown that many spiral galaxies host circumnu- clear star formation regions (CNSFRs) and several works have studied these regions. In one of the most recent works, Knapen (2005) uses Hα images of a sample of spiral galaxies to study CNSFRs, as well as nuclear star formation regions. They found that CNSFRs are present in about 20% of spiral galaxies and al- most always occur within a barred host. In this paper we adopt the definition of Knapen (2005) to distinguish CNSFRs from nuclear star formation regions: CNSFRs are the ones observed within circular areas with diameters of a few kpc around the galactic nucleus, while nuclear star formation regions are ob- served in more internal areas a few hundred parsec across. In particular, the formation of CNSFRs in barred galaxies seems to be due to the radial inflow of gas along bars to the galac- tic center (Roberts et al. 1979; Friedli & Benz 1995; Sakamoto et al. 1999; Sheth et al. 2005; Jogge et al 2005). This gas is accumulated in a ring structure between two inner Lindblad resonances (Knapen et al. 1995). Gas inflow is also necessary to feed the nuclear massive black hole in active galactic nuclei (AGNs) and surround- ing nuclear star formation regions, probably yielding the so- called AGN-starburst connection (Norman & Scoville 1988; Send offprint requests to: Oli L. Dors Jr., e-mail: [email protected] Terlevich et al. 1990; Heckman et al. 1997; Gonz´alez Delgado et al. 1998; Ferrarese & Merritt 2000; Cid Fernandes et al. 2001; Storchi-Bergmann 2001; Heckman 2004). Recent obser- vational studies have shown evidence of gas inflow: Storchi- Bergmann et al. (2007) and Fathi et al. (2006) report the dis- covery of gas streaming motions along spiral arms towards the LINER nuclei of the galaxies NGC 6951 and NGC 1097, while Fathi et al. (2007) found gas inflow to the nuclear star formation region of the galaxy M 83. Moreover, several works have found star formation around active nuclei (e.g. Riffel et al. 2007; Kauffmann et al. 2003) In CNSFRs, the line ratio [O iii]λ5007/Hβ (hereafter [O iii]/Hβ) is usually weak or not observable and the emis- sion lines have lower equivalent widths than those of disk H ii regions (e.g. Diaz et al. 2007; Bresolin et al. 1999; Bresolin & Kennicutt 1997; Kennicutt et al. 1989). The low value of [O iii]/Hβ has been attributed to over-solar abundances (e.g. Boer & Schulz 1993). At their location in the innermost parts of galactic disks, these objects are expected to be the most metal-rich star-forming regions. In fact, D´ıaz et al. (2007) ob- tained long-slit observations for 12 CNSFRs and, using a semi- empirical method to derivate of oxygen abundances, found that some of them have 12+log(O/H)=8.85, a value consistent with the maximum oxygen abundance value derived for central parts of spiral galaxies (Pilyugin et al. 2007). Regarding the equivalent width of emission lines, Kennicutt et al. (1989) found that the equivalent widths of Hα 2 O. L. Dors Jr., T. Storchi-Bergmann, Rogemar A. Riffel, A. A Schimdt: Circumnuclear star formation regions 96 pc for the former and 68 pc for the latter galaxy. The reader is referred to the papers above for a full description of the data. Figure 1 shows spectra of CNSFRs identified by the let- ters F and J in Figs. 2 and 3 located in the circumnuclear rings of NGC 6951 and NGC 1097, respectively. Using the observed spectra we built maps for flux distributions in the Hα emis- sion line and for the line ratios [N ii]λ6583/Hα, ([S ii]λ6717 + λ6731)/Hα, as well as for the EW(Hα). These maps are shown in Figs. 2 and 3. Throughout the paper we identify the line ra- tios above as [N ii]/Hα and [S ii]/Hα. For NGC 6951 the ob- served field completely covers the circumnuclear ring. The data of the ring were separated from those of the inner region by fit- ting two ellipses to the Hα image of NGC 6951: one to the inner border of the ring and the other to the outer border (see Storchi- Bergmann et al. 2007). In the NGC 6951 Hα map, we clearly see several regions of star formation. We have identified nine of them as indicated in Fig. 2. Since for NGC 1097 the observa- tion field covered only part of the ring, we could only identify two regions in this galaxy (see Fig. 3). For these measurements, we assumed that each star forma- tion region has a circular symmetry, whose center corresponds to the peak in the Hα intensity and the radius defined as the one where the flux reaches 50 % of the peak value. Nevertheless, we can see from Fig. 2 that the CNSFRs have irregular morpholo- gies and the assumption above can result in some uncertainties in the Hα flux and in line ratio measurements. To estimate how much flux we may be missing, for NGC 6951, we compared the sum of the Hα flux of all regions (obtained from photom- etry within circular apertures) with the flux integrated over the whole ring. We found that the diffuse gas contribution (from regions not covered by the circles) is about 12 %. A similar re- sult was found for individual H ii regions by Oey & Kennicutt (1997). In relation to line ratios, we measured the median val- ues considering different values for the radius of each region and differences no greater than 10% were found. We thus es- timate that uncertainties in both fluxes and line ratios due to uncertainties in the geometry of the CNSFRs are about 10%. In Table 1 we show the adopted area of each region, the lumi- nosity of Hα, and the median values of EW(Hα), [N ii]/Hα, and [S ii]/Hα. 3. Photoionization models To derive the metallicity, ionization parameter, and the age of the CNSFRs, we employed the photoionization code Cloudy/95.03 (Ferland 2002) to build a grid of models cov- ering a large space of nebular parameters. The models were built as in Dors & Copetti (2006), with metallicities Z=1.1, 1.0, 0.7, 0.6 Z⊙, and ionization parameter logarithm (log U) of −2.0, −2.5, and −3.0 dex. In each model a stellar cluster was assumed as the ionizing source with the stellar energy distributions ob- tained from the synthesis code S T ARBURS T 99 (Leitherer et al. 1999). We built models with stellar clusters formed by in- stantaneous and continuous bursts with stellar upper mass lim- its of Mup = 20, 30 and 100 M⊙, Z= 1.0 and 0.4 Z⊙, and ages ranging from 0.01 to 10 Myr with a step of 0.5 Myr. Clusters older than about 10 Myr and formed instantaneously have no more O stars (Leitherer & Heckman 1995), so were not consid- Fig. 1. Sample spectra: star-forming regions F in NGC 6951 (top) and J in NGC 1097 (bottom). EW(Hα) in CNSFRs are about 7 times lower than the ones in disk H ii regions. They proposed several scenarios to ex- plain this behavior, such as (i) deficiency of high-mass stars in the initial mass function (IMF), (ii) a long timescale for star formation, and (iii) high dust abundance. However, nei- ther of these mechanisms turned out to be the dominant effect. Unfortunately, spectroscopic data and abundance estimates are only available for a small number of CNSFRs. With the goal of increasing the number of CNSFRs with determined gas abundances and star formation rates, we have combined two-dimensional integral field unit (IFU) data of CNSFRs located in the galaxies NGC 1097 and NGC 6951 with a grid of photoionization models to derive their nebular gas properties. Using high-quality two-dimensional data, we could also investigate the cause of lower equivalent widths of emis- sion lines observed in CNSFRs with respect to those of inner disk HII regions. In Section 2 we give details about the obser- vational data we used, in Section 3 we present a description of the methodology used to derive the physical parameters, while in Section 4 we present our results. The discussion of the out- come and our final conclusions are given in Sections 5 and 6, respectively. 2. Observational data The observational spectroscopic data on NGC 1097 and NGC 6951 were drawn from the works of Fathi et al. (2006) and Storchi-Bergmann et al. (2007). Basically these data were obtained with IFUs of the Gemini Multi-Object Spectrographs (GMOS-IFU) on the Gemini South (for NGC 1097) and North (for NGC 6951) telescopes using the R400-G5325 grating and r-G0326 filter, covering the spectral range 5600-7000 Å with a resolution of R ≈ 2300. Three IFU fields were observed in each galaxy, covering an angular field of 15′′× 7′′. The reduc- tion procedure resulted in 50 × 70 spectra for each IFU field, each corresponding to an angular coverage of 0.1′′× 0.1′′. We adopted distances for NGC 6951 and NGC 1097 of 24 Mpc and 17 Mpc (Tully 1988), respectively, such that 1′′corresponds to O. L. Dors Jr., T. Storchi-Bergmann, Rogemar A. Riffel, A. A Schimdt: Circumnuclear star formation regions 3 Fig. 2. Star-forming ring in NGC 6951. From left to right: map of the Hα integrated flux (ergs cm−2 per pixel); line ratio maps [N ii]/Hα [S ii]/Hα and EW(Hα). The circles and the letters identify each region as defined in Section 2. ered in our models. The metallicity of the nebula was matched to the closest available metallicity of the stellar cluster. The so- lar composition refers to Asplund et al. (2005) and corresponds to log(O/H) = −3.30. 4. Results Since the observational data for NGC 6951 and NGC 1097 cov- ers a restricted wavelength range, few emission line ratios can be used in our analysis. Storchi-Bergmann et al. (1994) and Raimann et al. (2000) showed that [N ii]/Hα can be used to ob- tain the oxygen abundance of star-forming galaxies. However, this line ratio also dependents on the ionization parameter U (Kewley & Dopita 2002), thus the combination with other line ratios is desirable in order to decrease the uncertainties in the O/H estimates. Thus we used our photoionization model grid to build an [N ii]/Hα vs. [S ii]/ Hα diagram to derive Z and U of the CNSFRs located in NGC 6951 and NGC 1097. Recently, Viironen et al. (2007) have shown that diagnostic diagrams us- ing [N ii]/Hα and [S ii]/ Hα are a powerful tool for estimating Z and U of star-forming regions. Moreover, these line ratios are not dependent on the Mup and age assumed in the models (see Fig. 5). We point out that a cautionary note has been put forth by Mazzuca et al. (2006) regarding the use of diagnostic diagrams including the line ratio [N ii]/Hα: they can yield de- generate values for the metallicity, as star-forming regions with low Z and U have [N ii]/Hα values similar to regions with high Z and U. However, CNSFRs are known to have high abun- dances (near solar) (see Sect. 5.1), ruling out the possibility of low abundances. This is confirmed by the work of Storchi- Bergmann et al. (1996), which shows an increasing gas abun- dance towards the nucleus, including a few regions in circum- nuclear rings. The results of our models compared to observational data for NGC 6951 and NGC 1097 are shown in Fig. 4, in which we considered an instantaneous burst with age of 2.5 Myr and Mup = 100 M⊙. We can see that the observational data are well represented by models having metallicities Z ≈ Z⊙ and log U ≈ −2.3, with uncertainties of about 0.03 and 0.1 dex, due to the uncertainties in the line ratios discussed above. In Table 1 we show the Z and log U values for each region, obtained by linear interpolation from the model grid. In this table we also present the values of electron density Ne computed from the observed line ratio [S ii]λ6717/λ6731 and adopting Te = 104 K (Osterbrock 1989). The Ne values cover the range 50-600 cm−3, which is typical of H ii regions (Copetti et al. 2000). To obtain the ages of the CNSFRs, we used the observed values of equivalent widths and [O iii]/Hβ, which decrease al- most monotonically as a function of time (Magris et al. 2003, Copetti et al. 1986). In Fig. 5 we plot the values predicted by the models for EW(Brγ), EW(Hβ), EW(Hα), as well as [N ii]/Hα, [S ii]/Hα and [O iii]/Hβ versus age. We considered instanta- neous and continuous star formation bursts with Mup = 20, 30 and 100 M⊙, log U = −2.5 and Z = Z⊙. The values of the line ratios [N ii]/Hα and [S ii]/Hα were found to be approximately constant for the range of ages considered. However, [O iii]/Hβ shows a chaotic behavior between 2 and 6 Myr, mainly for Mup=100 M⊙ and instantaneous burst. This happens due to the increase in the number of ionizing photons by the presence of Wolf Rayet stars in this phase (Leitherer & Heckman 1995). This behavior was also noted by Dopita et al. (2006). As no feature of Wolf Rayet stars was observed in our data and be- cause [O iii]/Hβ is very dependent on many nebular parame- ters, it will not be considered as an age tracer in this paper. The use of the equivalent width of emission lines seems to be more reliable for deriving ages of star-forming regions, since they basically depend on Mup and on the slope of the IMF (Copetti et al. 1986). Using the EW(Hα) values shown in Table 1, the CNSFR ages predicted by instantaneous and continuous models are around 7 Myr and over 10 Myr, respectively, independent of the Mup assumed. Table 1 also presents the values for the star formation rate (SFR) computed using the luminosity of the re- 4 O. L. Dors Jr., T. Storchi-Bergmann, Rogemar A. Riffel, A. A Schimdt: Circumnuclear star formation regions Fig. 3. Left: map of the Hα-integrated flux (ergs cm−2 per pixel) showing the field covered by the IFU in NGC 1097, which includes the nucleus and part of the ring (at the bottom). Right: line ratio maps [N ii]/Hα [S ii]/Hα and EW(Hα) for the partial ring. The circles and the letters identify each region defined in Section 2. Fig. 4. Diagnostic diagram [N ii]/Hα vs. [S ii]/Hα showing the grid of photoionization models for an ionizing cluster with age 2.5 Myr and Mup = 100 M⊙. Solid lines connect curves of iso- Z, while dotted lines connect curves of iso-U. The values of log U and Z are indicated. Squares are the median values mea- sured for the CNSFRs in NGC 6951 and triangles are those for NGC 1097. Fig. 5. Evolution of the EW(Brγ), EW(Hβ), EW(Hα), and rel- evant line ratios as predicted by our photoionization models considering different Mup values and star formation regimes as indicated, for Z = Z⊙ and log U = −2.5. gions obtained in this paper and the relation SFR(M⊙ year−1)= 7.9 × L(Hα) (ergs s−1) given by Kennicutt (1998). Acording to our models, the CNSFRs of NGC 6951 and NGC 1097 have solar or slightly higher solar metallicities. 5. Discussion 5.1. Metallicities O. L. Dors Jr., T. Storchi-Bergmann, Rogemar A. Riffel, A. A Schimdt: Circumnuclear star formation regions 5 Table 1. Physical properties of the CNSFRs in NGC 6951 and NGC 1097 Region A (arcsec2) LHα (1038 erg s−1) [N II]/Hα [S II]/Hα SFR (M⊙ yr−1) EW(Hα) Z/Z⊙ log U Ne(cm−3) A B C D E F G H I J K 3.0 1.5 1.4 0.6 1.4 2.9 0.9 0.8 3.0 8.9 1.8 121 68.2 23.1 25.0 59.7 94.5 16.5 12.2 47.6 45.8 2.06 0.32±0.03 0.35±0.03 0.35±0.03 0.31±0.02 0.32±0.02 0.29±0.01 0.31±0.01 0.34±0.02 0.34±0.02 0.38±0.02 0.40±0.03 0.15±0.01 0.16±0.01 0.18±0.01 0.15±0.01 0.13±0.01 0.11±0.02 -- 0.22±0.02 0.20±0.01 0.16±0.01 0.18±0.01 0.09 0.05 0.02 0.02 0.05 0.07 0.01 0.01 0.04 0.03 0.002 28 31 16 39 45 26 32 31 38 12 7 1.04 1.05 1.04 1.03 1.05 1.04 -- 0.94 0.99 1.08 1.09 −2.3 −2.2 −2.4 −2.3 −2.1 −2.0 -- −2.6 −2.6 −2.1 −2.2 208±59 287±151 297±134 525±103 354±24 148±40 273±36 68±10 138±29 292±45 223±149 Considering an uncertainty on the order of 0.1 dex due to the method used to estimate abundances (see Evans 1986, Dors & Copetti 2005, 2006), our abundance values are comparable to the maximum oxygen abundance derived for central parts of spiral galaxies of 12+log(O/H) ≈ 8.87 (Pilyugin et al. 2007). Similarly, Mazzuca et al. (2006) present an emission-line diag- nostic analysis of the nuclear starburst ring of NGC 7742 and find roughly solar abundances. Recently, Sarzi et al. (2007) presented a study of CNSFRs located in eight galaxies which include the regions B, E and F of our sample. They also used empirical methods to derive the metallicities of these regions and find abundances similar to the ones derived in this paper. D´ıaz et al. (2007) use a semi- empirical abundance calibration to find that the CNSFRs in NGC 2903, NGC 3351, and NGC 3504 have metallicities com- parable to the ones derived in this paper. Until now, only a case of very low abundance on the order of 0.2-0.4 Z⊙ in the CNSFRs located in NGC 3310 was reported by Pastoriza et al. (1993). This low abundance can be understood as due to the fact that this galaxy has an interaction history (Elmegreen et al. 2002), through which neutral gas may have been accreted to the ring. Due to the fact that CNSFRs have very low excitation, with [O iii]/Hβ < 1 (e.g. Hagale et al. 2007, P´erez et al. 2000) and to their central location in the galactic disk, it is expected that these objects have somewhat higher abundance than is de- rived for inner disk H ii regions. To check this, we have plot- ted the [N ii]/Hα in Fig. 6 (lower panel) against oxygen abun- dance for some disk H ii regions located in 13 spiral galax- ies. This sample includes the most metal-rich nebulae in which temperature sensitive emission lines have been detected. The data were taken from Bresolin (2007), Bresolin et al. (2005, 2004), Kennicutt et al. (2003), Castellanos et al. (2002), and Skillman et al. (1996). With the exception of the oxygen abun- dances from Skillman et al. (1996), which were recalculated by us using the P-method (Pilyugin 2001), all O/H values were obtained using direct derivation of the electron temperature. Also included in Fig. 6 are the data of our sample. We can see that the CNSFRs show the highest oxygen abundances, to- gether with some disk nebulae, a result also obtained by D´ıaz et al. (2007). Since abundance determinations via photoioniza- tion models are overestimated by a factor of 0.1-0.4 dex when compared to direct determinations (Dors & Copetti 2005), our abundance estimates should be interpreted as upper limits. If CNSFRs are included in the computation of abundance gra- dients in spirals, we would expect therefore a plateau at very small galactocentric distances. 5.2. Starformationhistory The Hα luminosity values found for the CNSFRs in this paper are in the range 2 × 1038 − 1.2 × 1040 erg s−1, compatible with values of regions in circumnuclear rings of other spiral galaxies (e.g. D´ıaz et al. 2007). These luminosity values yield star for- mation rates of 0.002-0.1 M⊙ yr−1, characterizing a moderate star-forming regime. Averaging the SFR values in Table 1, we find 0.04 M⊙ yr−1, a similar value to the one found by Ho et al. (1997) for a sample of nuclear star-forming regions and by Shi et al. (2006), who studied a sample of 385 circumnuclear star-forming regions in galaxies with different Hubble types. We cannot conclude from our data alone whether the star formation has been continuous or instantaneous in the rings, although the former scenario has not been favored by previous studies. For example, Sarzi et al. (2007) use absorption-line in- dex diagrams to show that the hypothesis of a constant star for- mation in their sample can be ruled out. Allard et al. (2006) find that the star formation in the ring of M 100 have occurred in a series of short bursts for the past 500 Myr or so. These results can be understood within the following scenario. Once mas- sive stars are formed, they cause turbulence and heating in the molecular cloud, inhibiting further star formation (Hartmann et al. 2001, Blitz & Shu 1980). Since no systematic age sequence is observed along the NGC 6951 ring, the star formation mode seems to be as predicted by the popcorn model (Boker et al. 2008), in which the entire ring begins to form stars at about the same time or at random times. A similar result was also derived by Mazzuca et al. (2008), who used photometric Hα imaging of 22 nuclear rings to find that only a very small fraction of them show age sequences along the ring. 5.3. Ageestimates Our modeling gives ages of about 7.0 Myr for an instanta- neous burst and more than 10 Myr for a continuous burst, with 6 O. L. Dors Jr., T. Storchi-Bergmann, Rogemar A. Riffel, A. A Schimdt: Circumnuclear star formation regions the CNSFRs in NGC 1097 slightly older than in NGC 6951. To compare our age estimates with those of other CNSFRs, we collected equivalent widths of more CNSFRs from the literature and estimated their ages using Fig. 5. The data were extracted from D´ıaz et al. (2007) for NGC 2903 and NGC 3505, Hagele et al. (2007) for NGC 3351, Allard et al. (2006) for M 100, Reunanen et al. (2000) for NGC 7771, and Wakamatsu & Nishida (1980) for NGC 4314, which give EW(Hβ), EW(Hα), and EW(Brγ) in the range 1-17, 10-40, and 5-60 Å, respectively. The predicted ages for these EW values are of 5-8 Myr and more than 10 Myr for instantaneous and continuous bursts, respectively, thus similar to the ones we de- rived for the CNSFRs in the present paper. Somewhat higher EW values have been obtained by Mazzuca et al. (2008) for HII regions in nuclear rings of more than 20 galaxies. On the other hand, studies also utilizing photoionization models give ages of 1.5-3 Myr for H ii regions located in galactic disks and for H II galaxies (Dors & Copetti 2006; Stasinska & Izotov 2003; Bresolin et al. 1999; Copetti et al. 1985). We have also compared the ages of our sample with those of disk nebulae of similar metallicity. In Fig. 6 we show log(EW(Hβ)) and 12+log(O/H) plotted against log([N ii]/Hα) for these disk H ii regions, together with our data for the CNSFRs. We converted the EW(Hα) values shown in Table 1 in EW(Hβ) using the re- lation EW(Hβ)≈ 0.15 EW(Hα) obtained from our models. We can see that the EW(Hβ) of the CNSFRs are lower by a factor of about 30 than those of their counterpart disk H ii regions. Assuming a EW(Hβ)= 100 Å for the innermost disk regions and using Fig. 5, we find that these regions are about 3 Myr younger than the CNSFRs. 5.4. ThelowEWvalues The age differences found above may not be real, since the very low equivalent widths in CNSFRs stem from other effects than the age. Kennicutt et al. (1989) pointed out possible mecha- nisms such as (i) deficiency of high-mass stars in the initial mass function, (ii) very high dust abundance, (iii) contamina- tion of the continuum by contribution from the bulge or other underlying stellar populations, and (iv) continuous star forma- tion. These hypotheses are discussed below. Regarding the hypothesis of the deficiency of high-mass stars, since Mup of an ionizing cluster is dependent on the metallicity (Larson & Starrfield 1971; Kahn 1974; Shields & Tinsley 1976; Stasi´nska 1980; Vilchez & Pagel 1988; Campbell 1988; Bresolin et al. 1999; Dors & Copetti 2003, 2005, 2006), CNSFRs and innermost disk nebulae should have about same Mup. It is unlikely that the discrepancy in EWs is only due to variation in Mup. Even assuming that the ionizing cluster of disk nebulae have Mup = 100 M⊙ and the ones of CNSFRs have Mup = 20 M⊙, the age difference remains. Also a varia- tion in the IMF can be discarded, since stellar clusters appear to form with a universal IMF slope (Kroupa 2007, 2002, Chabrier 2003). The hypothesis of very high dust abundance in CNSFRs is not favored because they have similar Balmer decrements to those observed for disk H ii regions, suggesting that this is not the main source of the low EWs in CNSFRs (Kennicutt et al. 1989). Concerning the contamination of the continuum, several works have found the presence of underlying stellar popula- tions from previous generations of stars in CNSFRs, which can decrease the equivalent widths and yield no real age estimates from Fig. 5. For example, Buta et al. (2000) found a spread in ages of about 5 to 200 Myr for clusters in the circumnuclear ring of NGC 1326, with the largest number of clusters younger than 20 Myr. In the case of NGC 6951 and with measurements of the equivalent width of calcium triplet at ≈8500 Å, P´erez et al. (2000) found evidence of a population of red supergiant stars with ages of 10 to 20 Myr in the ring. An underlying stellar population can contribute in two ways to the low EWs by causing a Balmer absorption and yielding an extra contribution to the continuum. To estimate the effect of an underlying Balmer absortion, we constructed models of stellar populations (using the Padova 1994 tracks) with ages in the range 10-1000 Myr. We normalized the continuum of each model to the observed one and subtracted them from the ob- served spectra of the CNSFRs of NGC 6951. For all cases, the increase in the EW values due to subtracting of an underlying absorption was at most 10%, much less than needed to com- pensate for the low observed EW in CNSFRs. Determining of the continuum contribution from underly- ing stellar populations is not an easy task, since two types of stellar populations can be contributing: the stellar population from the bulge and stars formed in previous episodes of star formation in the ring. To evaluate the contribution of the un- derlying bulge, we performed aperture photometry on the con- tinuum images (not shown) measuring both the CNSFR con- tinuum and that of the bulge in the surrounding regions. We concluded that the bulge contributes with ≈ 75 % of the total flux in the case of NGC 6951 and with 50 % in NGC 1097. The corresponding corrections are an increase by a factor of 2 and 4 in EW(Hα) for NGC 1097 and NGC 6951, respectively. These factors increase the EW(Hα) in NGC 6951 from average values of ≈30 Å to ≈120 Å and from ≈10 Å to ≈20 Å in NGC 1097. The corresponding changes in age are from ≈7 Myr to ≈6 Myr in NGC 6951 and no change for NGC 1097 (≈7 Myr). Thus the contribution from the bulge to the continuum also does not ex- plain the low EW's in CNSFRs. Stauffer (1981) also studied the bulge contamination using aperture photometry of galaxy nuclei to measure the background contribution to the contin- uum flux, and Kennicutt et al. (1989) used those results to find a correction factor similar to the one found by us. Several generations of stars may coexist in CNSFRs be- cause the tidal effects are not strong enough to disrupt the clus- ter at the radius of the ring, allowing several generations of stars to coexist (Buta et al. 2000). Thus the continuum may have a contribution from previous bursts of star formation in the rings. To investigate the effect of this continuum, we built photoion- ization models to represent a scenario where three bursts of star formation have occurred, with ages of 0.01, 10, and 20 Myr for two values of Mup: 100 and 30 M⊙. This is illustrated in Fig. 7, which shows the evolution of EW(Hβ) as a function of time. It can be seen that the EW values decrease considerably when compared with those of a single burst (Fig. 5). If we consider a O. L. Dors Jr., T. Storchi-Bergmann, Rogemar A. Riffel, A. A Schimdt: Circumnuclear star formation regions 7 Fig. 6. Oxygen abundance vs. [N ii]/Hα (lower panel) and EW(Hβ) vs. [N ii]/Hα (upper panel). The squares represent data collected from the literature, while the circles represent the data of the CNSFRs analyzed in this paper. Fig. 7. Evolution of the EW(Hβ) as predicted by photoioniza- tion models considering three episodes of star formation as in- dicated. The ages correspond to the youngest burst. typical EW(Hβ) for CNSFRs of 10 Å (not corrected for the un- derlying bulge contribution) and compute the age from Fig. 7, we find ages of 6-7 Myr, similar to the values derived above. We conclude that none of the mechanisms discussed above alone yield the low EW values (and resulting larger age) observed in CNSFRs when compared to disk H ii regions. However, it could be that a combination of more than one mechanism can explain the lower EWs in CNSFRs. For exam- ple, we can consider the effects of both the dilution by the bulge and by previous bursts of star formation. A typical EW(Hβ)= 10 Å in NGC 6951 will increase to 40 Å if we correct for con- tamination by the bulge. Then, we can consider the effect of previous generation of stars in the ring using Fig. 7 instead of Fig. 5 to obtain the age. We obtain an age of 3-4 Myr, a value in consonance with ages of the innermost disk regions. In this way we eliminate the apparent difference in age between CNSFRs and inner-disk H ii regions. 6. Conclusions We have used optical IFU spectroscopic data of CNSFRs in NGC 6951 and NGC 1097, which compared to a grid of pho- toionization models, in order to obtain physical parameters of these regions. We find that the star formation rates are in the range 0.002-0.14 M⊙ yr−1, thus covering a range from very low to moderate star-forming regime. We also find that the CNSFRs have oxygen abundances of 12+log(O/H)≈8.8, similar to those of the most metal-rich nebulae located in inner parts the disks of spiral galaxies. We investigated the cause of the very low equivalent width of emission lines found in CNSFRs, which suggests an older age for these objects than for disk H ii regions. Among the sev- eral scenarios invoked to explain this fact, we have concluded that the contamination of the continua of CNSFRs by under- lying contributions from both old bulge stars and stars formed in the ring in previous episodes of star formation (10-20 Myr) yield the observed low equivalent widths. Correcting for these contributions, there are not any more significant differences in ages between the CNSFRs and inner disk H ii regions. Acknowledgements. We thank the anonymous referee for valuable suggestions which helped to improve the paper. This work was based on observations obtained at the Gemini Observatory, which is oper- ated 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 SECYT (Argentina) References Allard, E. L., Knapen, J. H., Peletier, R. F., Sarzi, M. 2006, MNRAS, 371, 1087 Asplund, M., Grevesse, N., Sauval, A. J. 2005, Cosmic Abundances as Records of Stellar Evolution and Nucleosynthesis, Vol. 336 of Astronomical Society of the Pacific Conference Series, The Solar Chemical Composition. pp. 25 Boer, B., & Schulz, H. 1993, A&A, 277, 397 Boker, T., Falc´on-Barroso, Schinnerer, E., Knapen, J. H., Ryder, S. 2008, AJ, 135, 479 Blitz, L., Shu, F. H. 1980, ApJ, 238, 148 Bresolin, F. 2007, ApJ, 656, 186 Bresolin, F., Schaerer, D., Gonz´alez Delgado, R. M., Stas´ınska, G. 2005, A&A, 441, 981 Bresolin, F., Garnett, D. R., Kennicutt, R. C. 2004, ApJ, 615, 228 Bresolin, F., Kennicutt, R. C., Garnett, D. R. 1999, ApJ, 510, 104 Bresolin, F., Kennicutt, R. C 1997, AJ, 113, 3 Buta, R., Treuthardt, P. M., Byrd, G. G., Crocker, D. A. 2000, AJ, 120, 1289 Campbell, A. 1988, ApJ, 335, 644 Castellanos, M., D´ıaz, A. I., Terlevich, E. 2002, MNRAS, 329, 315 Chabrier, G. 2003, PASP, 115, 763 8 O. L. Dors Jr., T. Storchi-Bergmann, Rogemar A. Riffel, A. A Schimdt: Circumnuclear star formation regions Copetti, M. V. F., Mallmann, J., Schmidt, A. A., Castaneda, H. O. 2000, A&A, 357, 621 Copetti, M. V. F., Pastoriza, M. G., Dottori, H. A. 1986, A&A, 156, 111 P´erez, E., M´arquez, I. M., Durret, F. et al. 2000, A&A, 353, 893 Pilyugin, L. S., Thuan, T. X., V´ılchez, J. M. 2007, MNRAS, 376, 353 Pilyugin, L. S. 2001, A&A, 369, 594 Terlevich, E., Diaz, A. I., Terlevich, R. 1990, MNRAS, 242, 271 Copetti, M. V. F., Pastoriza, M. G., Dottori, H. A. 1985, A&A, 152, 0707.0651 427 Cid Fernandes, R., Heckman, T, Schimitt, H., Gonz´ales Delgado, R., Storchi-Bergmann, T. 2001, MNRAS, 558, 81 D´ıaz, A. I., Terlevich, E., Castellanos, M., Hagele, G. 2007, MNRAS, 382, 251 Dors, O. L., & Copetti, M. V. F. 2006, A&A, 452, 437 Dors, O. L., & Copetti, M. V. F. 2005, A&A, 437, 837 Dors, O. L., & Copetti, M. V. F. 2003, A&A, 404, 969 Dopita, M. A., Fischera, J., Sutherland, R. S. et al. 2006, ApJS, 167, 177 Elmegreen, D. M., Chromey, F., McGrath, E., Ostenson, J. 2002, AJ, 123, 1388 Evans, I. N. 1986, ApJ, 309, 544 Hagele, G. F., D´ıaz, A. I., Cardaci, M. C., Terlevich, E., Terlevich, R. 2007, MNRAS, 378, 163 Hartmann, L., Ballesteros-Paredes, J.; Bergin, E. A. 2001, ApJ, 562, 852 Heckman, T. M. 2004, cbhg.symp, 358 Heckman, T. M., Gonz´alez Delgado, R. M., Leitherer, C. et al. 1997, ApJ, 482, 114 Ho, L. C., Filippenko, A. V.; Sargent, W. L. W. 1997, 487, 591 Fathi, K., Beckman, J. E., Lundgren, A. A. et al. 2007, [arXiv:astro- ph/0712.0793] Fathi, K., Storchi-Bergmann, T., Riffel, R. A. et al. 2006, ApJ, 641, L25 Ferland, G. J. 2002, Hazy, a brief introduction to Cloudy 96.03, Univ. Kentucky, Dept. Phys., Astron. internal report Ferrarese, L., & Merritt, D. 2000, ApJ, 539, L9 Friedli, D., & Benz 1995, A&A, 301, 649 Gonz´alez Delgado, R. M., Heckman, T., Leitherer, C. et al. 1998, ApJ, 505,174 Jogge, S., Scoville, N., Kenney, J. D. P. 2005, ApJ, 630, 837 Kahn, F. D. 1974, A&A, 37, 149 Kauffmann, G., Heckman, T. M., Tremonti, C. et al. 2003, MNRAS, 346, 1055 Kennicutt, R. C., Bresolin, F., Garnett, D. R. 2003, ApJ, 591, Kennicutt, R. C. 1998, ARA&A, 36, 189 Kennicutt, R. C. Keel, W., Blaha, C. A. 1989, ApJ, 97, 1022 Kewley, L. J., & Dopita, M. A. 2002, ApJSS,142, 35 Knapen, J. H., Beckman, J. E., Heller, C. H., Shlosman, I., de Jong, R. S. 1995, ApJ, 454, 623 Knapen, J. H., 2005, A&A, 429, 141 Kroupa, P. 2007, [arXiv:astro-ph/0703282] Kroupa, P. 2002, Science, 295, 82 Larson, R. B., Starrfield, S. 1971, A&A, 13, 190L Leitherer, C., Schaerer, D., Goldader, J. D. et al. 1999, ApJS, 123, 3 Leitherer, C., Heckman, T. M. 1995, ApJS, 96, 9 Magris C. G., Binette, L., Bruzual A. G. 2003, ApJS, 149, 313 Mazzuca, L. M., Knapen, J. H., Veilleux, S., Regan, M. W. 2008, ApJS, 174, 337 Mazzuca, L. M., Sarzi, M., Knapen, J. H., Veilleux, S., Swaters, R. 2006, ApJ, 649, L79 Morgan, W. W. 1958, PASP, 70, 364 Norman, C., & Scoville, N. 1988, ApJ, 332, 124 Oey, M. S., & Kennicutt, R. C. 1997, MNRAS, 291, 827 Osterbrock, D. E. 1989, Astrophysics of Gaseous Nebulae and Active Galactic Nuclei, (Mill Valley: Univ. Science Books) Pastoriza, M. G., Dottori, H. A., Terlevich, E., Terlevich, R. 1993, MNRAS, 260, 177 Tully, R. B. 1988, Nearby Catalogue, Cambridge University Press Raimann, D., Storchi-Bergmann, T., Bica, E., Melnick, J. Schmitt, H. 2000, MNRAS, 316, 559 Reunanen, J., Kotilainen, J. K., Laine, S., Ryder, S. D. 2000, ApJ, 529, 853 Riffel, R., Pastoriza, G., Rodr´ıgues-Ardila, A., Maraston, C. 2007, ApJ, 659, L103 Roberts, W. W., Huntley, J. M., van Albada, G. D. 1979, ApJ, 233, 67 Riffel, R. A., Storchi-Bergmann, T., Winge, C., Barbosa, F. K. 2006, MNRAS, 373, 2 Sakamoto, K., Okumura, S. K., Ishizuki, S., Scoville, N. Z. 1999, ApJ, 525, 691 Sarzi, M., Allard, E. L., Knapen, J. H., Mazzuca, L. M. 2007, MNRAS, 380, 949 Stauffer, J. P. 1981, Ph.D thesis, University of California, Berkeley Stasi´nska, G. 1980, A&A, 84, 320 S´ersic, J. L., Pastoriza, M. 1967, PASP, 79, 427 Sheth, K., Vogel, S. N., Regan, M. W., et al. 2005, ApJ, 632, 217 Shields, G. A., Tinsley, B. M. 1976, ApJ, 203, 66 Shi, L., Gu, Q. S., Peng, Z. X. 2006, A&A, 450. 15 Skillman E. D., Kennicutt, R. C., Shields, G. A., Zaritsky, D. 1996, ApJ, 462, 147 Stasi´nska, G., Izotov, I. 2003, A&A, 397, 71 Storchi-Bergmann, T., Dors, O. L., Riffel, R. A. et al. 2007, ApJ, 670, 959 Storchi-Bergmann, T., Gonz´alez Delgado, R. M. et al. 2001, ApJ, 559,147 Storchi-Bergmann, T., Rodriguez-Ardila, A., Schmitt, H. R., Wilson, A. S., Baldwin, J. A. 1996, ApJ, 472, 83 Storchi-Bergmann, T., Calzetti, D., Kinney, A. 1994, ApJ, 572, 581 Viironen, K., Delgado-Inglada, G., Mampaso, A., Magrini, L., Corradi, R. L. M. 2007, MNRAS, 381, 1719 Vilchez, J. M., & Pagel, B. E. J. 1988, MNRAS, 231, 257 Wakamatsu, K., & Nishida, M. T. 1980, PASJ, 32, 389 List of Objects 'NGC 6951' on page 1 'NGC 1097' on page 1 'NGC 6951' on page 1 'NGC 1097' on page 1 'NGC 6951' on page 1 'NGC 1097' on page 1 'M 83' on page 1 'NGC 6951' on page 2 'NGC 1097' on page 2 'NGC 1097' on page 2 'NGC 6951' on page 2 'NGC 1097' on page 2 'NGC 6951' on page 2 'NGC 1097' on page 2 'NGC 6951' on page 2 'NGC 6951' on page 2 'NGC 1097' on page 2 'NGC 6951' on page 2 'NGC 1097' on page 2 O. L. Dors Jr., T. Storchi-Bergmann, Rogemar A. Riffel, A. A Schimdt: Circumnuclear star formation regions 9 'NGC 6951' on page 2 'NGC 6951' on page 2 'NGC 6951' on page 2 'NGC 1097' on page 2 'NGC 6951' on page 2 'NGC 6951' on page 3 'NGC 1097' on page 3 'NGC 6951' on page 3 'NGC 1097' on page 3 'NGC 6951' on page 3 'NGC 1097' on page 3 'NGC 6951' on page 4 'NGC 1097' on page 4 'NGC 6951' on page 4 'NGC 1097' on page 4 'NGC 6951' on page 4 'NGC 1097' on page 4 'NGC 7742' on page 4 'NGC 6951' on page 5 'NGC 1097' on page 5 'NGC 2903' on page 5 'NGC 3351' on page 5 'NGC 3504' on page 5 'NGC 3310' on page 5 'M 100' on page 5 'NGC 6951' on page 5 'NGC 1097' on page 5 'NGC 6951' on page 5 'NGC 2903' on page 5 'NGC 3505' on page 5 'NGC 3351' on page 5 'M 100' on page 5 'NGC 7771' on page 5 'NGC 4314' on page 5 'NGC 1326' on page 6 'NGC 6951' on page 6 'NGC 6951' on page 6 'NGC 6951' on page 6 'NGC 1097' on page 6 'NGC 1097' on page 6 'NGC 6951' on page 6 'NGC 6951' on page 6 'NGC 1097' on page 6 'NGC 6951' on page 6 'NGC 1097' on page 6 'NGC 6951' on page 7 'NGC 6951' on page 7 'NGC 1097' on page 7
astro-ph/0412693
1
0412
2004-12-30T11:31:30
IVS Products for Precise Global Reference Frames
[ "astro-ph" ]
VLBI plays a unique and fundamental role in the maintenance of global reference frames which are required for precise positioning in many research areas related e.g to the understanding and monitoring of global changes, to geodesy, to space missions etc. The International VLBI Service for Geodesy and Astrometry coordinates internationally the VLBI components and resources and is tasked by IAG and IAU for the provision of products describing the Celestial Reference Frame through positions of quasars and their changes with time and also for products describing the rotation of the Earth in space. This paper summarises today's status of the products, achieved over the last years by evolving the observing programs. It points out the activities to improve further on the product quality to meet future service requirements, which will come up with the need for highly precise global reference frames.
astro-ph
astro-ph
Proceedings of the 7th European VLBI Network Symposium Bachiller, R. Colomer, F., Desmurs, J.F., de Vicente, P. (eds.) October 12th-15th 2004, Toledo, Spain IVS Products for Precise Global Reference Frames Wolfgang Schluter1 and Nancy Vandenberg2 1 Bundesamt fur Kartographie und Geodaesie, Fundamentalstation Wettzell, D 93444 Kotzting, Germany 2 NVI, Inc./ Goddard Space Flight Center, Greenbelt, USA Abstract. VLBI plays a unique and fundamental role in the maintenance of global reference frames which are required for precise positioning in many research areas related e.g to the understanding and monitoring of global changes, to geodesy, to space missions etc. The International VLBI Service for Geodesy and Astrometry coordinates internationally the VLBI components and resources and is tasked by IAG and IAU for the provision of products describing the Celestial Reference Frame through positions of quasars and their changes with time and also for products describing the rotation of the Earth in space. This paper summarises today's status of the products, achieved over the last years by evolving the observing programs. It points out the activities to improve further on the product quality to meet future service requirements, which will come up with the need for highly precise global reference frames. 1. Introduction 2. Products and improvements The International VLBI Service for Geodesy and Astrometry (IVS) is a service of the International Association of Geodesy (IAG), the International Astronomical Union (IAU) and the Federation of Astronomical and Geophysical Data Analysis Services (FAGS). The main task of the IVS is the coordina- tion of VLBI components in order to guarantee the provision of the products and parameters to realize the Celestial Reference Frame (CRF), the Terrestrial Reference Frame (TRF) and the Earth Rotation with its orientation of the rotation axis in both reference frames and its angular velocity through the Earth Orientation Parameter (EOP). The EOP enables the transfor- mation between TRF and CRF. VLBI is fundamental and unique for the realization of the CRF through a catalogue of quasar positions. The IAU tasked IVS to maintain the CRF and released a resolution in August 2001, during the General Assembly in Birmingham. VLBI contributes strongly to the TRF by the determination of station positions, in particular of baseline lengths between the stations. Due to the long in- tercontinental baselines VLBI is strongly supporting the scale of the TRF. Time series of station positions give informa- tion about their movements (plate motions). As VLBI pro- vides the complete set of EOP and uniquely the UT1-UTC (DUT1) parameter and the CRF, VLBI is the key technique for the monitoring of global reference frames. The interna- tional collaboration is based on a Call for Participation in 1998 with respect to the IVS Terms of Reference (ToR). Proposals were made to operate more than 73 permanent com- ponents by 37 institutions in 17 countries. IVS has more than 250 Associate Members. Annual Reports and Meeting Proceedings are published (Vandenberg 1999, Vandenberg, Baver 2001, Vandenberg, Baver 2002 , Vandenberg, Baver 2003 , Vandenberg, Baver 2004 , Vandenberg, Baver 2000, Vandenberg, Baver 2002b , Vandenberg, Baver 2004b). When IVS started in March 1999, the demand for continuity in maintaining the reference frames forced us to employ the exist- ing observing programs set up by the US Naval Observatory as NEOS, or by NASA as CORE. In 2001 a working group (IVS WG2) was established to review the products and the existing observing programs. The WG2 report was the basis for improv- ing products and evolving observing programs to meet service requirements. The IVS products can be defined in terms of their accuracy, reliability, frequency of observing sessions, tempo- ral resolution of the estimated parameters, time delay from ob- serving to final product, and frequency of solutions. The situ- ation before 2002 and the goals for the follow on years with IVS products are described in detailed tables in the WG2 re- port (Schuh et al. 2002). The main IVS products, their current accuracies and the goals are summarized in table 1. The VLBI technique allow us to provide additional products and IVS in- tends to set up the extended products summarized in table 2. As of late 2001, IVS products were generated from 3 days/week observing with 6-station networks. The time delay ranged from several days up to 4 months, with an overall av- erage value of 60 days. Over the next years, the goals of IVS with respect to its products were the following (specific goals for each product are listed in the WG2 report tables): -- improve the accuracies of all EOP and TRF products by a factor of 2 to 4 and improve the sky distribution of the CRF, -- decrease the average time delay from 60 to 30 days, and designate 2 days per week as rapid turnaround sessions with a maximum delay of 3-4 days, -- increase the frequency of observing sessions from 3 to 7 days per week, -- deliver all products on a regular, timely schedule. 310 Wolfgang Schluter and Nancy Vandenberg: IVS Products for Precise Global Reference Frames Fig. 1. Overview of IVS components 3. Evolving observing programs To meet its product goals, beginning with the 2002 observ- ing year, IVS designed an observing program coordinated with the international community. The 2002 observing program in- cluded the following sessions: -- EOP: Two rapid turnaround sessions each week (IVS R1 and IVS R2), initially with 6 stations, increasing to 8. These networks were designed with the goal of having compara- ble xp and yp results. One-baseline 1-hr INTENSIVE ses- sions four times per week, with at least one parallel session. -- TRF: Monthly TRF sessions with 8 stations including a core network of 4 to 5 stations and using all other stations three to four times per year. -- CRF: Bi-monthly RDV sessions using the Very Long Baseline Array (VLBA) and 10 geodetic stations, plus quarterly astrometric sessions to observe mostly southern sky sources. -- Monthly R&D sessions to investigate instrumental effects, research the network offset problem, and study ways for technique and product improvement. -- Annual, or semi-annual if resources are available, 14-day continuous sessions to demonstrate the best results that VLBI can offer, aiming for the highest sustained accuracy. Although certain sessions have primary goals, such as CRF, all sessions are scheduled so that they contribute to all geodetic and astrometric products. Sessions in the observing program that are recorded and correlated using S2 or K4 technology will have the same accuracy and timeliness goals as those us- ing Mark 4/Mark 5. The new IVS observing program began in January 2002. The observing program and product delivery was accomplished by making some changes and improvements in IVS observing program resources (station days, correlator time, and magnetic media), by improving and strengthening analysis procedures, and by a vigorous technology develop- ment program. 4. Status and experiences of the new IVS Program two year after its implementation The number of station observing days increased by about 10compared to 2001, with an additional 12campaign. Not counting CONT02, the number of observing days increased Wolfgang Schluter and Nancy Vandenberg: IVS Products for Precise Global Reference Frames 311 Table 1. Summary of IVS main products, status and goal specifications Products Specifications Status UT1 - UTC DUT1 Celestial Pole dǫ, δψ Polar Motion xp, yp accuracy latency resolution frequency of solution accuracy latency resolution accuracy latency resolution frequency of solution accuracy accurary frequency of solution latency 1 improved distribution TRF(x,y,z) CRF (α; δ) xp 100µas, yp 200µas 1 -- 4 weeks . . . 4 months 1 day +3 days/week 5µs . . . 20µs 1 week 1 day 100µas . . . 400µas 1 -- 4 weeks . . . 4 months 1 day 3 days/week 5 mm -- 20 mm 0.2µas -- 3mas 1 year 3 -- 6 months xp, yp: 50µas 4 -- 3 days 1 day . . . 1 h Goals (2002 -- 2005) . . . 25µas . . . 1 day . . . 10 min . . . 7 days/week 3µs . . . 2µs 4 -- 3 days 1 day 50µas 4 -- 3 days 1 day 5mm 0.25µas1 1 year 3 months . . . 1 day . . . 10 min . . . 25µas . . . 1 day . . . 7 days/week . . . 2 mm . . . 1 month Table 2. Extended products derived by VLBI and provided or intended to be provided by IVS Earth Orientation Parameter additions Terrestrial Reference Frame (TRF) Celestial Reference Frame (CRF) Geodynamical Parameter Physical Parameter • dUT1/dt (length of day) • dxp/dt; dyp/dt (pole rates) • x-, y-, z- time series • Episodic events • Annual solutions • Non linear changes • Source structure • Flux density • Solid Earth tides (Love numbers h,l) • Ocean Loading (amplitudes and phases Ai, Φi) • Atmospheric loading (site-dependent coefficients) • Tropospheric parameters • Ionospheric mapping • Light deflection parameter γ by another 12will continuously increase such that by 2005 the top dozen geodetic stations will need to be observing up to 4 days per week - an ambitious goal. Increased station reliabil- ity and unattended operations can improve temporal coverage by VLBI and also allow substantial savings in operating costs. Higher data rate sessions can yield more accurate results, and therefore nearly all geodetic stations were upgraded to Mark 5 and K5 technology. A deployment plan for the Mark 5 system was proposed. As of the end of 2003 the correlator and most of the observing stations were equipped with Mark 5 digital recording systems. All correlators were committed to handling the IVS data with priority processing for meeting timely prod- uct delivery requirements. High capacity disks were purchased and organized in a common pool to replace magnetic tapes and to obtain additional recording media capacity. The progress in communication technologies supported the breakthrough for e- VLBI. Several tests have been conducted on national, continen- tal and global levels. Some station are already connected to fast internet links and regular applications for e-VLBI (real time or near real time) will be established. The 1-hr INTENSIVE ob- servation sessions are routinely transferred electronically to a correlator and will soon be operational. The increased amount of VLBI data to be produced under the new observing program required Analysis Centers to handle a larger load. Partially au- tomated analysis procedures helped to improve the timeliness of product delivery. As the official IVS product, a complete set of Earth Orientation Parameter is regularly submitted to the International Earth Rotation Service (IERS). The set is ob- tained as a combination of the individual solutions of the six IVS Analysis Centers (Nothnagel, Steinforth 2002). Up to the end of the year 2001, the parameters were derived from the NEOS observations, while since January 2002 the IVS R1 and IVS R4 were used. It should be noted that up to the end of 2001 NEOS was the rapid turn around program which since January 2002 now includes the IVS R1 and R4 sessions. The objective of the rapid turn around observation sessions was to minimize the delay between the observations and the availability of the results. For the NEOS the delay was approximately 2 weeks, which has not changed by its transition to the IVS R4 as all the routine procedures were already established. For the IVS R4, as new stations were added the data shipping procedures needed to be set up to be routine. Initially this caused unex- pected delays, which were overcome with time. For the IVS R1 processing, experience needed to be gained at the correlators at 312 Wolfgang Schluter and Nancy Vandenberg: IVS Products for Precise Global Reference Frames Nothnagel, A., Ch. Steinforth 2002, IVS Analysis Coordination, IAG CSTG Bulletin no.17, Progress Report 2001 Schuh H., Boehm, J. 2002, IVS- Zenith Wet path delay , 2002 General Meeting Proceedings, NASA/CP-2002- 210002, Greenbelt, MD Bonn, Haystack and Washington during the first months. The delay between the observation to the results is approximately two weeks since April 2002. This should be regarded as signif- icant and real progress, even though the WG2 goal of only 4 days has not been achieved. Improvements for data transmis- sion and a higher throughput at the correlator were achieved, after the implementation of the newly developed Mark 5 digi- tal data recording system, which has e-VLBI capabilities, al- lowing data transmission via high speed Internet links. The determination of DUT1 from the near-daily 1-hour observa- tions known as "Intensives" have been carried out since 1983 via the baseline Wettzell-Westford, since 1994 via the baseline Wettzell-Green Bank and since 2000 via the baseline Wettzell- Kokee Park. These baselines now employ Mark 5 systems . In the year 2002 a time series observed on the baseline Wettzell- Tsukba/Japan with the Japanese K4 system was set up. The regular application of fast Internet links for the INTENSIVE's will start soon, which will allow rapid availability of DUT1. The VLBI observations from the IVS R1 and R4 allow determi- nation of tropospheric parameters, in particular the wet zenith path delay. Since July 2003 the zenith wet path delay is an of- ficial IVS product. The University of Vienna is combining the solutions of five Analysis Centers (Schuh, Boehm 2002). 5. Plans A "Vision Paper 2010" for geodetic VLBI is under develop- ment by IVS. Considering our increasing requirements (e.g. the IGGOS project, the increase in radio frequency interference, our aging antennas) general refinements and upgrades of VLBI technology are needed in the future. A Working Group, WG3, was established with the objective to develop future visions. Goals include unattended observation, improved global cover- age of the network, employment of the new data transmission technologies and provision of near real time correlation and products. In collaboration with radio astronomers some guide- lines for future developments will be derived. References Vandenberg, N.R. (editor) 1999, Annual Report 1999, NASA/TP- 1999-209243, Greenbelt, MD Vandenberg, N. R., Baver K. D.(editors) 2001, Annual Report 2000, NASA/TP-2001-209979, Greenbelt, MD Vandenberg, N. R., Baver K. D.(editors) 2002, Annual Report 2001, NASA/TP-2002-210001, Greenbelt, MD Vandenberg, N. R., Baver K. D.(editors) 2003, Annual Report 2002, NASA/TP-2003-211619, Greenbelt, MD Vandenberg, N. R., Baver K. D.(editors) 2004, Annual Report 2003, NASA/TP-2004-212254, Greenbelt, MD; Vandenberg, N. R., Baver K. D.(editors) 2000, 2000 General Meeting Proceedings, NASA/CP-2000-209893, Greenbelt, MD Vandenberg, N. R., Baver K. D.(editors) 2002b, 2002 General Meeting Proceedings, NASA/CP-2002-210002, Greenbelt, MD Vandenberg, N. R., Baver K. D.(editors) 2004b, 2004 General Meeting Proceedings, NASA/CP-200-212255, Greenbelt, MD Schuh, H. et al. 2002, IVS Working Group 2 for Product Specification and Observing programs, Final Report, Annual Report 2001, page 13 - 45, NASA/TP-2002-210001, Greenbelt, MD
astro-ph/9808315
1
9808
1998-08-27T18:49:36
Spectroscopy of Outlying H II Regions in Spiral Galaxies: Abundances and Radial Gradients
[ "astro-ph" ]
We present the results of low dispersion optical spectroscopy of 186 H II regions spanning a range of radius in 13 spiral galaxies. Abundances for several elements (oxygen, nitrogen, neon, sulfur, and argon) were determined for 185 of the H II regions. As expected, low metallicities were found for the outlying H II regions of these spiral galaxies. Radial abundance gradients were derived for the 11 primary galaxies; similar to results for other spiral galaxies, the derived abundance gradients are typically -0.04 to -0.07 dex/kpc.
astro-ph
astro-ph
Accepted: Astronomical Journal Spectroscopy of Outlying H II Regions in Spiral Galaxies: Abundances and Radial Gradients National Radio Astronomy Observatory,2 PO Box 0, Socorro, NM 87801 Liese van Zee1 [email protected] John J. Salzer3 Astronomy Department, Wesleyan University, Middletown, CT 06459 -- 0123 [email protected] Martha P. Haynes Center for Radiophysics and Space Research and National Astronomy and Ionosphere Center4 Cornell University, Ithaca, NY 14853 [email protected] Physics Department, St. Lawrence University, Canton, NY 13617 Aileen A. O'Donoghue [email protected] Thomas J. Balonek Department of Physics and Astronomy, Colgate University 13 Oak Drive, Hamilton, NY 13346 [email protected] 8 9 9 1 g u A 7 2 1 v 5 1 3 8 0 8 9 / h p - o r t s a : v i X r a 1Jansky Fellow 2The National Radio Astronomy Observatory is a facility of the National Science Foundation, operated under a cooperative agreement by Associated Universities Inc. 3NSF Presidential Faculty Fellow 4The National Astronomy and Ionosphere Center is operated by Cornell University under a cooperative agreement with the National Science Foundation. -- 2 -- ABSTRACT We present the results of low dispersion optical spectroscopy of 186 H II regions spanning a range of radius in 13 spiral galaxies. Abundances for several elements (oxygen, nitrogen, neon, sulfur, and argon) were determined for 185 of the H II regions. As expected, low metallicities were found for the outlying H II regions of these spiral galaxies. Radial abundance gradients were derived for the 11 primary galaxies; similar to results for other spiral galaxies, the derived abundance gradients are typically -- 0.04 to -- 0.07 dex kpc−1. Subject headings: galaxies: abundances -- galaxies: ISM -- galaxies: spiral 1. Introduction Radial abundance gradients, in the sense that the inner H II regions have higher abundances than the outer ones, are found in almost all large spiral galaxies. For instance, a gradient of approximately -- 0.07 dex kpc−1 is observed in both the stellar and gaseous components of the Milky Way (Shaver et al. 1983; Smartt & Rolleston 1997). The physical basis for the presence of an abundance gradient is still a matter of debate; successful chemical evolution models typically include a combination of (1) a radial dependence on the star formation activity, generally taking the form of a Schmidt law (e.g., Phillipps & Edmunds 1991); (2) radial gas flows (e.g., Edmunds 1990; Gotz & Koppen 1992); and (3) a radial dependence on stellar yields due to IMF variations (e.g., Gusten & Mezger 1982). The relative importance of these processes can be constrained by investigating the observed gradients for a large sample of galaxies which span a range of morphological type (e.g., Vila -- Costas & Edmunds 1992) or environment (e.g., Skillman et al. 1996). Extensive studies of individual galaxies either via spectroscopy (e.g., NGC 5457, Kennicutt & Garnett 1996; NGC 2403, Garnett et al. 1997) or spectrophotometric imaging (e.g., NGC 628, NGC 6946, Belley & Roy 1992; NGC 925, NGC 1073, Martin & Roy 1994; NGC 2366, NGC 4395, Roy et al. 1996; NGC 1313, Walsh & Roy 1997) indicate that large numbers of H II region abundances are necessary to determine accurate abundance gradients. While spectrophotometric imaging observations have the advantage that large samples are easily obtained, they are not as accurate as spectroscopic studies for individual objects. Recently, several large spectroscopic surveys and literature compilations have appeared (e.g., McCall et al. 1985; Vila -- Costas & Edmunds 1992; Zaritsky et al. 1994). These studies indicate that the oxygen abundance gradients may be a function of Hubble type (e.g., Vila -- Costas & Edmunds 1992). Furthermore, the mass surface density appears to play a critical role in supporting an abundance gradient (e.g., Vila -- Costas & Edmunds 1992; Garnett et al. 1997). -- 3 -- If the abundance gradients are constant across the full radial extent of the galaxy, the outermost H II regions in spiral galaxies should have low abundances, similar to those found in low luminosity dwarf galaxies. Thus, the outlying H II regions of spiral galaxies provide a new environment in which to investigate metallicity effects on elemental yields (particularly of nitrogen) and other aspects of the star formation process. In this paper, we present the results of new spectroscopic observations of a large number of outlying H II regions in 11 spiral galaxies. In the few instances where the target galaxies did not have previous abundance measurements reported in the literature, spectra of inner H II regions were obtained as well. The observations presented in this paper complement previous spectroscopic observations by extending the radial coverage. In a previous paper (van Zee et al. 1998, hereafter Paper I), these H II region abundances were used to investigate the origin of nitrogen in low metallicity environments. Here, we present the complete data set and discuss the derived radial abundance gradients. This paper is organized as follows. In Section 2, the galaxy sample is defined and the imaging and spectroscopic observations are described. The observed H II region line strengths and diagnostics are presented in Section 3. The abundance derivations are described in Section 4; the radial abundance gradients for all 11 spiral galaxies are discussed in Section 5. The conclusions are summarized in Section 6. 2. Observations 2.1. Sample Selection To facilitate identification of outlying H II regions and to minimize internal extinction, a sample of 11 nearly face -- on spiral galaxies were selected for imaging and spectroscopic observations. In addition, two galaxies from the NGC 2805 group (NGC 2820 and IC 2458) were also observed; NGC 2820 is a nearly edge -- on spiral and IC 2458 is a dwarf galaxy. Table 6 lists the physical properties of the sample galaxies. The morphological type, inclination, position angle, and isophotal radius (at the 25 mag arcsec−2 surface brightness isophote) for each system have been taken from the RC3 (de Vaucouleurs et al. 1991). The inclination was derived by assuming an intrinsic axial ratio, q, of 0.2. With the exception of NGC 1068 and NGC 2903, all of the primary targets are classified as late -- type (Sc or Sd) spirals. Furthermore, all the galaxies in the sample have low inclination angles and subtend a large angular size. The distances listed in Table 6 were derived either from Cepheid variables (NGC 2403, Freedman & Madore 1988; NGC 925, Silbermann et al. 1996; NGC 5457, Kelson et al. 1996), or were estimated from the recessional velocity based on a Virgocentric infall model and an assumed H0 of 75 km s−1 Mpc−1 (NGC 628, NGC 1068, NGC 2903, NGC 3184: Zaritsky et al. 1994; NGC 4395: Wevers et al. 1986; NGC 1232, NGC 1637, NGC 2805, NGC 2820, IC 2458: this paper). Absolute magnitudes were computed from the apparent magnitudes listed in the RC3, corrected for internal and Galactic extinction, and the adopted distances. All of the primary targets have -- 4 -- B -- band apparent magnitudes brighter than 11.2, corresponding to absolute magnitudes ranging between -- 17.7 and -- 21.3. Of the 11 primary targets, three (NGC 1232, NGC 1637, and NGC 2805) had no previous abundance measurements reported in the literature. The other 8 were known to have moderate to steep abundance gradients (e.g., Zaritsky et al. 1994), but lacked abundance measurements of the outermost H II regions. The total number of new H II region abundances for each galaxy are tabulated in Table 6. Multiple H II regions were observed in all galaxies except for NGC 1068, where the outlying H II regions were discovered to be too faint for the planned abundance measurements. 2.2. Optical Imaging In preparation for the spectroscopic observations, wide field images of the selected galaxies were obtained with the Burrell Schmidt telescope at KPNO5 in 1995 April and October and 1996 January. The S2KA CCD had a read noise of 3 e− and a pixel scale of 2.028′′ pix−1. A gain of 2.5 DN/e− was selected for the 1995 April observing run; a gain of 3.7 DN/e− was selected for all subsequent observing runs. While the typical seeing was less than 2′′, the spatial resolution of these images is dominated by the large pixel size. R -- band and Hα narrow band images were obtained for all program galaxies. During the 1995 April run, three to four 450 second observations of each program galaxy were obtained with the R -- band filter; during the subsequent observing runs, one 600 second exposure was obtained for each galaxy. During all the observing runs, the Hα imaging consisted of several sets of ON -- OFF -- ON observations with individual exposure times of 600 seconds each. A total of three to five Hα sets were obtained for each program galaxy (corresponding to total on -- line integration times of 3600 to 6000 seconds). The Hα filter (KP 1468) had a central wavelength of 6571 A and a width of 84 A; the OFF images were obtained through a matched off -- line filter (KP 808) centered at 6411 A, with a width of 88A. Data reduction was performed within the IRAF6 package and followed standard practice. After bias subtraction and flat fielding, images with multiple exposures were aligned and the sky background was subtracted. The images were then scaled to the level of the "most photometric" (either the frame with the lowest airmass or the one taken under the best weather conditions) and combined with a median filter. At the same time, the off -- line Hα images were scaled to the level of the on -- line images, based on the observed counts for at least six to ten stars in each frame. Because the majority of these images were taken under non -- photometric conditions, they were used primarily as finding charts for the outlying (faint) H II regions. The relevant R -- band 5Observations made with the Burrell Schmidt of the Warner and Swasey Observatory, Case Western Reserve University. 6IRAF is distributed by the National Optical Astronomy Observatories. -- 5 -- and continuum subtracted Hα images are presented in Figure 1. The images are oriented with north to the top and east to the left. The 2′ slits on the Hα images indicate the spatial scale of each pair of images. Optical scale lengths (Rd), tabulated in Table 6, were derived from the R -- band images. The IRAF task ELLIPSE was used to fit ellipses of constant surface brightness. We note that several of the galaxies in the present sample are asymmetric, making such fits difficult to interpret. In particular, satisfactory fits were obtained only for NGC 628, NGC 925, NGC 1232, NGC 2403, and NGC 2903. The scale length tabulated in Table 6 for NGC 1068 was derived from a fit to only the outer isophotes, those with r > 50′′. By far the worst fit was obtained for the severely asymmetric spiral NGC 1637; the scale length presented here should only be taken as a rough estimate. The derived scaled lengths for the remaining galaxies, NGC 2805, NGC 3184, NGC 4395, and NGC 5457, are reasonable, although we caution that the ellipse fitting was not ideal due to their modest asymmetries. 2.3. Optical Spectroscopy Low resolution optical spectra of the selected H II regions were obtained with the Double Spectrograph on the 5m Palomar7 telescope during observing runs in 1996 May and November and 1997 January. The long slit (2′) was set at a 2′′ aperture during all observing runs; the seeing was generally better than 2′′. During all observing runs, a 5500 A dichroic was used to split the light to the two sides (blue and red), providing complete spectral coverage from 3500 -- 7600 A. The blue spectra were acquired with the 300 lines/mm diffraction grating (blazed at 3990 A). The red spectra were acquired with the 316 lines/mm diffraction grating (blazed at 7500 A). A thinned 800×800 TI CCD with a read noise of 8 e−, a gain of 1.5, and 9.2 A effective resolution (2.19 A/pix) was used on the blue side; a thinned 1024×1024 TEK CCD with a read noise of 7.5, gain of 2.0, and an effective resolution of 7.8A (2.46 A/pix) was used on the red side. The spatial scale of the long slit was 0.8 arcsec/pix on the blue and 0.49 arcsec/pix on the red. The CCD on the blue side had moderate focussing problems across the full wavelength range. During all observing runs, the focus was optimized for Hδ and [O III] λ4363. The observations were conducted via blind offsets from nearby stars. Stellar and H II region coordinates were derived from the Schmidt narrow band Hα images. Astrometric plate solutions were calculated using the coordinates of the bright stars in the HST Guide Star Catalog, yielding positions accurate to within 1′′. The telescope pointing was verified each run by using the same offsetting technique to move between several stars in a field. Offsets between stars and H II regions were typically less than 20′. Previous observations (e.g., van Zee et al. 1997) indicated that the 7Observations at the Palomar Observatory were made as part of a continuing cooperative agreement between Cornell University and the California Institute of Technology. -- 6 -- 5m telescope could point and track accurately; nonetheless, the newly commissioned offset guider was used to improve the tracking during long exposures. To reduce light losses due to atmospheric refraction, the slit position angle was set close to the parallactic angle during the observations. The slit positions are illustrated on the Hα images in Figure 1 and tabulated in Table 6. The slit numbers listed in Table 6 are ordered in increasing Right Ascension for each galaxy and each slit is labelled by the offset (arcseconds east -- west and north -- south) from the galaxy center (tabulated in Table 6). The offsets do not necessarily correspond to the location of an H II region since the telescope was often pointed between two (or more) H II regions in order to maximize the number of H II regions observed. The astrometric pointing centers are also given in Table 6 along with the position angle (PA) of the slit. Finally, the observing run and the total integration times are listed. Most observations consisted of two or three 1200 second exposures; in a few instances, primarily due to time constraints, the integration time was only 300 seconds. Flux calibration was obtained by observations of at least 5 standards per night (Stone 1977; Oke & Gunn 1983; Massey et al. 1988), interspersed with the H II region observations. Wavelength calibration was obtained by observations of arc lamps taken before and after each H II region observation. A Hollow Cathode (Fe and Ar) lamp was used to calibrate the blue spectra; a combination of He, Ne, and Ar lamps were used to calibrate the red spectra. The spectra were reduced and analyzed with the IRAF 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 the arc lamp observations 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 each H II region were then extracted from the rectified images. Throughout this paper, the H II region nomenclature is based on east -- west and north -- south offsets from the galaxy center. These offsets were derived either from the pointing center (if only one H II region was in the slit), or were computed for each H II region from the pointing center, the slit position angle, and the spatial scale of the image. These offsets, therefore, are generally only accurate to 1 -- 2′′. The 1D spectra were corrected for atmospheric extinction and flux calibrated. Since the night -- to -- night variation in the calculated sensitivity function was small, the sensitivity function for each run was created using the standard stars from all nights. As a final step, the individual 20 minute exposures were combined. In general, there was excellent agreement in the continuum level of the final blue and red spectra for each H II region, indicating that the flux calibration and extraction regions were well matched. A few representative spectra are shown in Figure 2. The three H II regions in Figure 2 are ordered in increasing distance from the center of NGC 1232. Figure 2(a) illustrates a typical spectrum from the inner galaxy; absorption features are present and the low ionization lines, such as [O II] and [N II], are strong while the high ionization lines, such as [O III], are extremely weak -- 7 -- or absent. Figure 2(b) illustrates a typical spectrum from the outer galaxy; the emission lines dominate the spectrum, with both the low and high ionization lines at moderate strength. Finally, Figure 2(c) illustrates the low abundance nature of the outermost H II regions. Here, the [O III] lines are very strong while [N II] and [S II] are weak, indicating that this is a high temperature (low abundance) H II region. Furthermore, [O III] λ4363 is clearly detected in this spectrum, permitting an accurate temperature estimate (see Section 4.3). 3. Line Ratios and Diagnostic Diagrams 3.1. Extinction The reddening along the line of sight to each H II region was computed from the observed strengths of the Balmer emission lines. The intrinsic line ratios of Hα/Hβ and Hγ/Hβ were interpolated from the values of Hummer & Storey (1987) for case B recombination, assuming that Ne = 100 cm−3 and Te = 10000 K. For the few H II regions where the [O III] λ4363 line was detected, the electron temperature as derived from the [O III] lines (see Section 4.3) was used instead of the nominal 10000 K. The reddening function, f (λ), normalized at Hβ from the galactic reddening law of Seaton (1979) as parameterized by Howarth (1983), was used, assuming a value of R = AV /EB−V = 3.1. When necessary to bring the two Balmer line ratios into agreement, an underlying Balmer absorption with an equivalent width of 2 A was assumed. The reddening coefficients, cHβ, derived for each H II region are listed in Table 6 and are shown as a function of radius in Figure 3. While there is a fair amount of scatter in the reddening plots, the general trend is for the outer H II regions to have lower extinction than the inner ones. Similar extinction gradients have been seen in photometric studies of the dust content in spiral galaxies (e.g., Peletier et al. 1995). 3.2. Line Intensities The reddening corrected line intensities relative to Hβ are given in Table 6. In this and subsequent tables, each H II region is identified by its east -- west and north -- south offsets from the galaxy center (in arcsec, north and east are positive) and by its slit number (Table 6). Furthermore, the H II regions for each galaxy are ordered by increasing radial distance. The error associated with each relative line intensity was determined by taking into account the Poisson noise in the line, the error associated with the sensitivity function, the contributions of the Poisson noise in the continuum, read noise, sky noise, and flat fielding or flux calibration errors, the error in setting the continuum level (assumed to be 10% of the continuum level), and the error in the reddening coefficient. For a few of the high abundance H II regions, the [O III] lines were so weak that [O III] λ4959 was not detected. In those instances, the tabulated [O III] line intensities were derived by assuming that the [O III] λ5007 line intensity was 2.88 times the intensity of [O III] λ4959. -- 8 -- 3.3. Diagnostic Diagrams Since the majority of the H II regions in this sample are from the outer regions of spiral galaxies, they should form well defined H II region sequences in diagnostic diagrams (e.g., Osterbrock 1989). A few galactic nuclei (NGC 925+002 -- 002, NGC 1637 -- 001 -- 000, and NGC 4395 -- 003 -- 003) were also observed, however, so it is worthwhile to examine such plots to determine if these are stellar photoionized emission regions or AGN. Several of the line ratios commonly used as H II region diagnostics are tabulated in Table 6 and are graphically shown in Figures 4 and 5. In these and subsequent Figures, the H II regions in each galaxy are coded by common symbols. The well defined H II region sequence formed by the [N II]/Hα and [O III]/Hβ line ratios is illustrated in Figure 4. A theoretical curve from the H II region model of Baldwin et al. (1981) is superposed on the data in Figure 4(a), illustrating that the majority are indeed H II regions. Panel (b) shows the same diagnostic, but with the sum of [O II]/Hβ and [O III]/Hβ (R23) on one axis instead of [O III]/Hβ (after McCall et al. 1985). In both of these plots, NGC 4395 -- 003 -- 003 (marked by an arrow) deviates from the overall H II region sequence. This supports the suggestion by Ho et al. (1995) that the nucleus of NGC 4395 harbors a dwarf Seyfert. Similarly, while the diagnostic diagrams from [S II]/Hα are less definitive, the nucleus of NGC 4395 again falls outside of the H II region sequence in Figure 5. Due to the unknown ionization source for NGC4395 -- 003 -- 003, it has been excluded from the following abundance analysis. 4. Abundance Determinations Several assumptions are required to convert observed line ratios into nebular abundances. In sections 4.1 -- 4.3, we describe the methodology used to obtain the electron density and electron temperature for each H II region. In section 4.4, we describe the ionization correction factors which were used to convert the derived ionic abundances to nebular abundances. Finally, in sections 4.5 and 4.6 we discuss the derived oxygen abundances and the relative enrichments of nitrogen, neon, sulfur, and argon. 4.1. Electron Density The density sensitive line ratio [S II] λ6717/6731 is tabulated for each H II region in Table 6 and graphically illustrated in Figure 6. The maximum value of this line ratio in the low density limit is marked on Figure 6. The majority of the H II regions are within the low density limit (I(λ6717)/I(λ6731) > 1.35). In the following abundance analysis, an electron density of 100 cm−3 was assumed for the majority of the H II regions; in the few instances where the [S II] line ratios -- 9 -- were below the low density limit, a version of the FIVEL program of De Robertis et al. (1987) was used to calculate the electron density from the observed [S II] line ratios. 4.2. Semi -- Empirical Oxygen Abundances The optical oxygen lines ([O II] λ3727 and [O III] λλ4959, 5007) are sensitive to both the oxygen abundance and electron temperature. At low metallicities, the dominant coolant is the collisionally excited Lyman series, and thus the total oxygen line intensity (R23 ≡ ([O II]+[O III])/Hβ) increases as the abundance increases. As the metallicity increases, however, the infrared fine structure lines, such as the 52µ and 88µ oxygen lines and the C+ 158µ transition, begin to dominate the cooling. At an oxygen abundance of approximately 1/3 solar, the intensity of the optical oxygen lines reaches a maximum and then declines as the oxygen abundance increases. The behavior of the optical oxygen line intensities as a function of oxygen abundance has been examined via semi -- empirical methods (e.g., Searle 1971; Pagel et al. 1979; Alloin et al. 1979; Edmunds & Pagel 1984; Dopita & Evans 1986; Skillman 1989) and derived from H II region ionization models (e.g., McGaugh 1991; Olofsson 1997). The theoretical models of McGaugh (1991) are shown in Figure 7. The surface is double valued due to the ambiguity between high and low abundances. Furthermore, the geometry of the H II region (represented by the average ionization parameter, ¯U, the ratio of ionizing photon density to particle density) introduces a further spread in the estimated abundance for a given R23. In addition, aging of the H II regions can also introduce further scatter as both the ionization parameter and the shape of the ionizing spectrum evolve (e.g., Stasi´nska & Leitherer 1996; Olofsson 1997); we do not include such effects in the present analysis. The observed line strengths for the H II regions are plotted in Figure 7. The lowest values of R23 occur for the innermost H II regions observed, and therefore presumably populate the high abundance side of the R23 surface (see below). At these high oxygen abundances, the ionization parameter ( ¯U) is approximately 0.001 for all of the H II regions. In contrast, the H II regions with low oxygen abundances span a wide range of ¯U, indicating that it is important to include such ionization effects when using the strong line method to estimate oxygen abundances. The process of determining empirical oxygen abundances from the strong line ratios requires several steps. First, abundances for both the high and low abundance side of the surface were determined. Further information is required, however, to resolve the ambiguity between these two values. The degeneracy can be broken by looking at the line strengths of other species. For instance, the [N II]/[O II] line ratio forms a single parameter sequence from high to low abundance (Figure 8). In general, H II regions with log ([N II]/[O II]) < -- 1.0 are believed to have low oxygen abundances, while those with log ([N II]/[O II]) greater than -- 1.0 are on the high abundance side. For those H II regions where the [N II]/[O II] diagnostic was conclusive (those with log ([N II]/[O II]) > -- 0.8 or log ([N II]/[O II]) < -- 1.05), the abundance from the appropriate side of the surface was assigned. -- 10 -- A large fraction of the H II regions in this sample fall in the ambiguous region of the R23 relation (the turnover region) where the [N II]/[O II] diagnostic is inconclusive. Thus, it was necessary to identify an alternative means of determining the oxygen abundance for these H II regions. Like the [N II]/[O II] diagnostic, the [N II]/Hα ratio increases with increasing oxygen abundance (Figure 9). The relation between the [N II]/Hα line ratio and the oxygen abundance was determined by a weighted least squares fit to the low abundance H II regions from van Zee et al. (1997) and the high abundance H II regions (8.8 < 12 + log (O/H) < 9.1) in the present sample: 12 + log(O/H) = 1.02 log([NII]/Hα) + 9.36. (1) Note that the [N II]/Hα line ratio is only valid as a metallicity estimator for 12 + log (O/H) < 9.1 and in the absence of shock excitation. Furthermore, with typical errors of 0.2 dex or more, it is not a particularly accurate abundance estimator. However, it does provide an additional method to break the ambiguity of the R23 abundances. For those H II regions where the oxygen abundance could not be resolved based on the [N II]/[O II] line ratio, the [N II]/Hα relation was used to calculated an empirical oxygen abundance. This empirical oxygen abundance was then compared with the two (high and low abundance) estimates from the R23 relation. Whichever R23 abundance was closest to the empirical abundance was then assigned, unless both the R23 abundances were significantly different from the empirical abundance. In the instances where neither the upper nor the lower branch R23 abundances were within 0.2 dex of the empirical abundance, the empirical abundance was assigned. This primarily occurred when the empirical abundance estimate was between 8.2 and 8.5, corresponding to the turnover region of the R23 relation, a region where typical errors are several tenths of a dex. 4.3. Electron Temperature The preferred method of obtaining an estimate of the electron temperature of the ionized gas is to derive it from the corrected line strengths of the [O III] lines. For the 16 H II regions where [O III] λ4363 was detected (9% of the sample), the [O III] ratio λλ4959, 5007/λ4363 is tabulated in Table 6. A version of the FIVEL program of De Robertis et al. (1987) was used to compute Te from this reddening -- corrected line ratio. Not unexpectedly, the H II regions with detectable [O III] λ4363 correspond to those with high Te; the electron temperatures were typically in the range of 9000 to 14000 K, with errors of 300 to 600 K. In most cases, however, [O III] λ4363 was not detected, or was contaminated by the nearby In these instances, the electron temperature was estimated from Hg λ4358 night sky line. ionization models. Because oxygen is one of the dominant coolants, the electron temperature depends strongly on the oxygen abundance. Thus, if the oxygen abundance is known, the electron temperature can be estimated. The anticorrelation between oxygen abundance and electron temperature (Pagel et al. 1979; Shaver et al. 1983) was used in a self -- consistent manner to -- 11 -- estimate Te. The errors in the derived temperatures are quite large. For the H II regions where the electron temperature was derived from the assumed oxygen abundance, the error in the electron temperature was set to yield the appropriate error (as determined by the calibration process) in the computed oxygen abundance. For the abundances estimated from R23 (71% of the sample), this corresponds to 0.1 dex, or ∼500 K for a Te of 5000 to 7000 K. For the oxygen abundances estimated from the [N II]/Hα relation (20% of the sample), errors of 0.2 dex in the oxygen abundance were adopted, corresponding to temperature uncertainties of ∼ 2000 K for a Te of 10000 K. As a check of self -- consistency, the same method was used to estimate the electron temperatures for those H II regions where [O III] λ4363 was detected; in general, the two methods yielded similar results, with absolute differences less than the estimated errors for the empirical method. The electron temperatures adopted for the subsequent abundance analysis are tabulated in Table 6. Numerical models of H II regions have shown that, particularly at low abundances, there are significant differences between the temperatures in the high -- and low -- ionization zones (Stasi´nska 1980). This difference must be taken into account to derive accurate abundances. We therefore adopt the approach taken by Pagel et al. (1992), who use the H II region models of Stasi´nska (1990) to derive an approximation for the electron temperature of the O+ zone: Te(O+) = 2[T −1 e (O++) + 0.8]−1, (2) where Te is the electron temperature in units of 104 K. 4.4. Ionization Correction Factors For all atoms other than oxygen, the derivation of atomic abundances requires the use of ionization correction factors (ICFs) to account for the fraction of each atomic species which is in an unobserved ionization state. To estimate the nitrogen abundance, we have assumed that N/O = N+/O+ (Peimbert & Costero 1969). To estimate the neon abundance, we have assumed that Ne/O = Ne++/O++ (Peimbert & Costero 1969). To determine the sulfur and argon abundances, we are forced to adopt an ICF from published H II region models to correct for the unobserved S+3 and Ar+3 states. We have adopted analytical forms of the ICFs as given by Thuan et al. (1995). In H II regions where S++ λ6312 was not detected, an ICF was assumed based on the average value of the ICF in all other H II regions. The use of an "average" ICF assumes that the ionization characteristics of the H II regions with non -- detections of the S++ λ6312 line are similar to those where λ6312 was detected. This is not necessarily the case, but it appears to work reasonably well. We have assumed an error of 50% in this "average" ICF of 5.8±2.9. -- 12 -- 4.5. Oxygen Abundances The oxygen abundances tabulated in Table 6 were derived from the observed line strengths (Table 6) and the relevant emissivity coefficients (based on the electron temperature (Table 6) and electron density) as calculated by a version of the FIVEL program of De Robertis et al. (1987). The errors in the derived abundances are dominated by the errors inherent in the electron temperature estimates. As expected, the outermost H II regions in the spiral galaxies have low oxygen abundances. The new observations confirm the low abundance nature of H681 (+010+885) in NGC 5457 (Garnett & Kennicutt 1994), which has the lowest oxygen abundance (12 + log (O/H) = 7.92 ± 0.03) of all the spiral galaxy H II regions in this sample (the H II regions in the dwarf galaxy IC 2458 also have similarly low oxygen abundances). The highest oxygen abundances are found in the asymmetric spiral NGC 1637, with typical abundances of 1.5 to 2 times the solar value. The majority of the oxygen abundances were ultimately derived from the strong line method (either from the McGaugh (1991) R23 calibration or the modified version of this R23 relation as described in Section 4.2). Other calibrations of the R23 relation yield slightly different results. In particular, the R23 calibration of Zaritsky et al. (1994) (ZKH) will yield very similar abundances for H II regions with 12 + log (O/H) > 8.5 (Table 6). Note, however, that since the ZKH analytical form does not explicitly take into account the turnover in the R23 relation, it can result in artificially high abundances for the outlying H II regions. A comparison of the derived abundances from the modified McGaugh calibration (this paper) with those derived from the Zaritsky et al. (1994) analytical formula is shown in Figure 10. Also shown in Figure 10 and tabulated in Table 6 are the R23 abundance estimates from the calibration of Edmunds & Pagel (1984). It is clear from this plot that the EP84 calibration will yield steeper abundance gradients than either the ZKH or the modified McGaugh calibrations used here. 4.6. Nitrogen, Neon, Sulfur, and Argon Abundances The abundances of nitrogen, neon, sulfur, and argon were computed from the observed line strengths (Table 6), the emissivity coefficients, and the ionization correction factors described in Section 4.4. Their abundances relative to oxygen are tabulated in Table 6. The original purpose of the observations described here was to investigate the N/O ratio in the outlying, low abundance, H II regions of spiral galaxies. Detailed analysis of the derived N/O ratios is presented in a separate paper (van Zee et al. 1998, Paper I). Briefly, as seen in other high metallicity H II regions, the N/O ratio increases linearly with the oxygen abundance at high abundance. However, the N/O ratio plateaus for H II regions with metallicities less than 1/3 solar, indicating that there is a primary origin for nitrogen in low metallicity environments. The relative abundances of the α elements (Ne, S, and Ar) are expected to be constant with -- 13 -- increasing oxygen abundance. Their abundances relative to oxygen are shown in Figure 11. Also shown in Figure 11 are the solar values (Anders & Grevesse 1989). As expected for primary elements, the relative abundance ratios are essentially constant: the mean log (Ne/O) is -- 0.65 ± 0.18; the mean log (S/O) is -- 1.55 ± 0.15; and the mean log (Ar/O) is -- 2.28 ± 0.11. These mean ratios are similar to those derived for low metallicity H II regions in dwarf galaxies (e.g., Thuan et al. 1995; van Zee et al. 1997). 5. Radial Abundance Gradients The oxygen abundances presented in this paper increase the total number of H II regions observed in these spiral galaxies. Furthermore, a large fraction of the H II regions in this sample are located at radii greater than half of the isophotal radius. These outlying H II regions provide a significant lever -- arm for the computation of radial abundance gradients (see also Ferguson et al. 1998). The spiral galaxy abundance gradients are illustrated in Figure 12. All deprojected radii have been computed from the H II region offsets and the assumed inclination and position angles (Table 6); the deprojected radii are tabulated (in arcsec) in Table 6. In Figure 12, the deprojected radii have been normalized by the isophotal radius (R25), as listed in the RC3 (de Vaucouleurs et al. 1991) and tabulated in Table 6. The filled symbols in Figure 12 represent H II regions from the present study. The open circles represent data from the literature: NGC 628 (McCall et al. 1985), NGC 925 (Zaritsky et al. 1994), NGC 1068 (Evans & Dopita 1987; Oey & Kennicutt 1993), NGC 2403 (McCall et al. 1985; Fierro et al. 1986; Garnett et al. 1997), NGC 2903 (McCall et al. 1985; Zaritsky et al. 1994), NGC 3184 (Zaritsky et al. 1994), NGC 4395 (McCall et al. 1985), and NGC 5457 (Kennicutt & Garnett 1996). To place the literature abundances on a common scale with the abundances derived in this paper, the analytical form of the R23 calibration presented by Zaritsky et al. (1994) was used to compute the oxygen abundances. As illustrated in Table 6 and Figure 10, this analytical form of the R23 calibration yields similar abundance estimates for high abundance H II regions. The only exceptions to this recalibration of the literature data were the few instances where abundances were derived from an electron temperature which was directly measured, either through the [O III] or [O II] line ratios (e.g., NGC 5457, Garnett & Kennicutt 1994; NGC 2403, Garnett et al. 1997). The oxygen abundance gradients obtained from weighted least -- squares fits to both the literature and new data points are tabulated in Table 6. For those galaxies with previous abundance gradients reported in the literature, the newly derived abundance gradients are generally in agreement with older results (see, e.g., the compilations by Zaritsky et al. 1994; Garnett et al. 1997). The few exceptions are NGC 925, NGC 1068, and NGC 4395. For NGC 1068, the one additional H II region in the present study is at such a large radius compared to the previous observations that it determines the gradient. Further observations of the other outlying H II regions in NGC 1068 will be necessary to confirm the derived gradient. For NGC 925 and NGC 4395, the new observations also result in steeper abundance gradients than previously reported -- 14 -- in the literature. The majority of the H II regions in both of these galaxies are at, or near, the R23 turnover, so the derived abundances have large errors. The previous spectroscopic studies of NGC 925 (Zaritsky et al. 1994) and NGC 4395 (McCall et al. 1986) were limited by the small number of H II regions observed; in these studies, the abundance gradients were swamped by the intrinsic errors of the abundance measurements. Spectrophotometric imaging observations of a large number of H II regions in NGC 925 (Martin & Roy 1994) suggested that a steeper abundance gradient existed; the newly derived abundance gradient from the spectroscopic observations is in good agreement with the spectrophotometric imaging results. Spectrophotometric imaging observations of NGC 4395 indicated that the abundance gradient was flat (or nonexistent) in this low luminosity galaxy (Roy et al. 1996); the new spectroscopic results suggest that a shallow gradient exists, but the errors in the derived slope are quite large. A large scatter in the oxygen abundance for a given radius can indicate azimuthal variations of the abundance gradient An apparent branching of the abundance gradient of NGC 5457 at radii between R/R25 ∼0.2 -- 0.5 was noticed by Kennicutt & Garnett (1996). The two branches are spatially separate, with the higher abundance H II regions lying on the northwest half of the galaxy and the lower abundance H II regions lying on the southeast side (near NGC 5461). As discussed by Kennicutt & Garnett (1996), these abundance variations could arise from the inherent asymmetry of the galaxy, from tidal interactions with its companions NGC 5474 and/or NGC 5477, or from accretion of high -- velocity gas. Other galaxies, such as NGC 925 and NGC 2403, also show large variations in the abundance at a given radius. The H II regions in both of these galaxies lie near the turnover region of the R23 relation, however. Thus, the large scatter in oxygen abundance is probably due to the intrinsic errors in the abundance measurements. New observations with solid detections of the temperature sensitive lines will be necessary to address the validity of these abundance variations. Observations of the Milky Way suggest that the abundance gradient may flatten in the outer disk (e.g., V´ılchez & Esteban 1996). The outermost points of NGC 5457 suggest that the abundance gradient may flatten in this galaxy as well. However, previous claims of flattening in the inner regions of NGC 5457 (e.g. Zaritsky 1992; Scowen et al. 1992; Vila -- Costas & Edmunds 1992) have not been substantiated (Henry & Howard 1995; Kennicutt & Garnett 1996). To determine if flattening of the abundance gradient is a common phenomena of the outer disk, observations of H II regions at even larger radii will be necessary. Such observations are challenging due to the faintness of the outermost H II regions, but not impossible (e.g., Ferguson et al. 1998). 6. Conclusions We have presented the results of high signal -- to -- noise spectroscopy of 186 H II regions spanning a range of radius in 13 spiral galaxies. Abundances for several elements (oxygen, nitrogen, neon, sulfur, and argon) were derived for all except the nucleus of NGC 4395. Below, we summarize the main conclusions. -- 15 -- (1) The H II region diagnostic diagrams indicate that the nucleus of NGC 4395 harbors a dwarf Seyfert. The remaining 185 H II regions form a well defined H II region sequence in the Osterbrock diagnostic diagrams. Furthermore, the majority of the H II regions are low density, with an assumed Ne of 100 cm−3. (2) A modified version of the R23 calibration of McGaugh (1991) was used to derive the majority of the oxygen abundances. In this method, the [N II]/Hα line ratios are used to break the degeneracy between the upper and lower branches; furthermore, this line ratio is used to estimate the oxygen abundance for H II regions which fall in the R23 turnover region (8.2 < 12 + log (O/H) < 8.5). (3) The enrichment of the α elements (neon, sulfur, and argon) relative to oxygen is constant for all H II regions in this sample. The mean α -- to -- oxygen ratios are consistent with those derived for low metallicity H II regions (e.g., Thuan et al. 1995; van Zee et al. 1997). The mean abundance ratios are: -- 0.65 ± 0.18, -- 1.55 ± 0.15, and -- 2.28 ± 0.11 for log (Ne/O), log (S/O), and log (Ar/O), respectively. (4) The outlying H II regions presented here extend the radial range of H II region abundances measured in spiral galaxies. For those galaxies with previous abundance measurements, the newly derived radial gradients are generally consistent with those previously reported. For the three galaxies which had no abundance measurements reported in the literature (NGC 1232, NGC 1637, and NGC 2805), each now has abundances measured in at least 15 H II regions. (5) As expected, the outermost H II regions in the spiral galaxies have abundances similar to those of low luminosity dwarf galaxies. These outlying H II regions provide a new environment in which to investigate metallicity effects on stellar yields and the star formation process (see, e.g., van Zee et al. 1998). We thank Elizabeth Barrett for processing several of the Hα images and Stacy McGaugh for providing us with the ionization models. We thank the anonymous referee for comments which improved the presentation of this paper. We thank Scott Lacey, Stacey Davis, Jennifer Heldmann, Lai Man Lee, Alison Schirmer, and Matt Pickard for their assistance in the observations at the Burrell Schmidt. They and TJB acknowledge travel support from the W.K. Keck Foundation for their support of astronomy through the Keck Northeast Astronomy Consortium. AOD acknowledges travel support by the Judge Francis Bergen Career Development Award in Astrophysics, from the Dudley Observatory, Schenectady, New York. We acknowledge the financial support by NSF grants AST95 -- 53020 to JJS and AST90 -- 23450 and AST95-28860 to MPH. 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. -- 16 -- REFERENCES Alloin, D., Collin -- Soufrin, S., Joly, M., & Vigroux, L. 1979, A&A, 78, 200 Anders, E., & Grevesse, N. 1989, Geochim. Cosmochim. Acta, 53, 197 Baldwin, J. A., Phillips, M. M., & Terlevich, R. 1981, PASP, 93, 5 Belley, J., & Roy, J. -- R. 1992, ApJS, 78, 61 De Robertis, M. M., Dufour, R. J., & Hunt, R. W. 1987, JRASC, 81, 195 de Vaucouleurs, G., de Vaucouleurs, A., Corwin, H. G., Buta, R., Paturel, G., & Fouqu´e, P. 1991, Third Reference Catalogue of Bright Galaxies (Springer, New York) (RC3) Dopita, M. A., & Evans, I. N. 1986, ApJ, 307, 431 Edmunds, M. G. 1990, MNRAS, 246, 678 Edmunds, M. G., & Pagel, B. E. J. 1984, MNRAS, 211, 507 (EP84) Evans, I. N., & Dopita, M. A., 1987, ApJ, 319, 662 Ferguson, A. M. N., Gallagher, J. S., & Wyse, R. F. G. 1998, AJ, in press Fierro, J., Torres -- Peimbert, S., & Peimbert, M. 1986, PASP, 98, 1032 Freedman, W. L., & Madore, B. F. 1988, ApJ, 332, L63 Garnett, D. R., & Kennicutt, R. C. 1994, 426, 123 Garnett, D. R., Shields, G. A., Skillman, E. D., Sagan, S. P., & Dufour, R. J. 1997, ApJ, 489, 63 Gotz, M., & Koppen, J. 1992, A&A, 262, 455 Gusten, R., & Mezger, P. G. 1982, Vistas, Astro., 26, 159 Henry, R. B. C., & Howard, J. W. 1995, ApJ, 438, 170 Ho, L. C., Filippenko, A. V., & Sargent, W. L. W. 1995, ApJS, 98, 477 Kelson, D. D., Illingworth, G. D., Freedman, W. F., Graham, J. A., Hill, R., Madore, B. F., Saha, A., Stetson, P. B., Kennicutt, R. C., Mould, J. R., Hughes, S. M., Ferrarese, L., Phelps, R., Turner, A., Cook, K. H., Ford, H., Hoessel, J. G., & Huchra, J. 1996, ApJ, 463, 26 Kennicutt, R. C., & Garnett, D. R. 1996, ApJ, 456, 504 Kobulnicky, H. A., & Skillman, E. D. 1996, 471, 211 Kobulnicky, H. A., & Skillman, E. D. 1997, 489, 636 Martin, P., & Roy, J. -- R. 1994, ApJ, 424, 599 McCall, M. L., Rybski, P. M., & Shields, G. A. 1985, ApJS, 57, 1 (MRS) McGaugh, S. S. 1991, ApJ, 380, 140 Oey, M. S., & Kennicutt, R. C. 1993, ApJ, 411, 137 Olofsson, K. 1997, A&A, 321, 29 -- 17 -- Osterbrock, D. E. 1989, Astrophysics of Gaseous Nebulae and Active Galactic Nuclei (University Science Books, Mill Valley) Pagel, B. E. J., Edmunds, M. G., Blackwell, D. E., Chun, M. S., & Smith, G. 1979, MNRAS, 189, 95 Pagel, B. E. J., Simonson, E. A., Terlevich, R. J., & Edmunds, M. G. 1992, MNRAS, 255, 325 Peimbert, M., & Costero, R. 1969, Bol. Obs. Tonantzintla y Tacubaya, 5, 3 Peletier, R. F., Valentijn, E. A., Moorwood, A. F. M., Freudling, W., Knapen, J. H., & Beckman, J. E. 1995, A&A, 300, L1 Phillipps, S., & Edmunds, M. G. 1991, MNRAS, 251, 84 Roy, J. -- R., Belley, J., Dutil, Y., & Martin, P. 1996, ApJ, 460, 284 Scowen, P. A., Dufour, R. J., & Hester, J. J. 1992, AJ, 104, 92 Searle, L. 1971, ApJ, 168, 327 Shaver, P. A., McGee, R. X., Newton, L. M., Danks, A. C., Pottasch, S. R. 1983, MNRAS, 204, 53 Silbermann, N. A., Harding, P., Madore, B. F., Kennicutt, R. C., Saha, A., Stetson, P. B., Freedman, W. L., Mould, J. R., Graham, J. A., Hill, R. J., Turner, A., Bresolin, R., Ferrarese, L., Ford, H., Hoessel, J. G., Han, M., Huchra, J., Hughes, S. M. G., Illingworth, G. D., Phelps, R., & Sakai, S. 1996, ApJ, 470, 1 Skillman, E. D. 1989, ApJ, 347, 883 Skillman, E. D., & Kennicutt, R. C. 1993, ApJ, 411, 655 Skillman, E. D., Kennicutt, R. C., & Shields, G. A., & Zaritsky, D. 1996, ApJ, 462, 147 Smartt, S. J. & Rolleston, W. R. J. 1997, ApJ, 481, L47 Stasi´nska, G. 1980, A&A, 84, 320 Stasi´nska, G. 1990, A&AS, 83, 501 Stasi´nska, G., & Leitherer, C. 1996, ApJS, 107, 661 Thuan, T. X., Izotov, Y. I., & Lipovetsky, V. A. 1995, ApJ, 445, 108 van Zee, L., Haynes, M. P., & Salzer, J. J. 1997, AJ, 114, 2479 van Zee, L., Salzer, J. J., & Haynes, M. P. 1998, ApJ, 497, L1 (Paper I) Vila -- Costas, M. B., & Edmunds, M. G. 1992, MNRAS, 259, 121 V´ılchez, J. M. & Esteban, C. 1996, MNRAS, 280, 720 Walsh, J. R., & Roy, J. -- R. 1997, MNRAS, 288, 715 Wevers, B. M. H. R., van der Kruit, P. C., & Allen, R. J. 1986, A&AS, 66, 505 Zaritsky, D. 1992, ApJ, 390, L73 -- 18 -- Zaritsky, D., Kennicutt, R. C., & Huchra, J. P. 1994, ApJ, 420, 87 (ZKH) This preprint was prepared with the AAS LATEX macros v4.0. -- 19 -- Table 1. Galaxy Properties Galaxy NGC 0628 NGC 0925 NGC 1068 NGC 1232 NGC 1637 NGC 2403 NGC 2805 IC 2458 NGC 2820 NGC 2903 NGC 3184 NGC 4395 NGC 5457 RA (2000) Dec (2000) Morph.a Distanceb Type [Mpc] c MB ia p.a.a [arcsec] [arcsec] a R25 Rd No. HII Regions 01:36:41.6 02:27:16.8 02:42:40.6 03:09:45.3 04:41:28.0 07:36:51.0 09:20:20.3 09:21:30.0 09:21:45.4 09:32:10.0 10:18:16.8 12:25:48.9 14:03:12.5 15:47:03 33:34:46 .SAS5.. .SXS7.. -- 00:00:46 RSAT3.. .SXT5.. -- 20:34:41 -- 02:51:27 .SXT5.. .SXS6.. 65:36:04 .SXT7.. 64:06:11 .I.0.P* 64:14:20 64:15:25 .SBS5P/ .SXT4.. 21:30:18 .SXT6.. 41:25:28 .SAS9.. 33:32:52 54:20:55 .SXT6.. 9.7 9.29 14.4 21.5 8.6 3.25 23.5 23.5 23.5 6.3 8.7 4.5 7.4 -- 20.1 -- 19.8 -- 21.3 -- 21.2 -- 18.5 -- 19.1 -- 20.7 -- 16.9 -- 20.1 -- 19.9 -- 19.3 -- 17.7 -- 21.2 25 58 32 30 36 60 42 · · · · · · 64 21 18 18 25 102 70 90 15 126 125 · · · · · · 17 135 147 37 314. 314. 212. 222. 120. 656. 189. · · · · · · 378. 222. 395. 865. 71.8 86.2 111.4 60.2 20.6 117.3 74.8 · · · · · · 85.9 55.8 149.9 152.9 19 44 1 16 15 17 17 3 4 9 17 11 13 aMorphological type, inclination, position angle, and isophotal radius from de Vaucouleurs et al. 1991 (RC3). bDistance references: NGC 628, NGC 1068, NGC 2903, NGC 3184 -- Zaritsky et al. 1994; NGC 925 -- Silbermann et al. 1996; NGC 1232, NGC 1637, NGC 2805, IC 2458, NGC 2820 -- Virgocentric infall model; NGC 2403 -- Freedman & Madore 1988; NGC 4395 -- Wevers et al. 1986; NGC 5457 -- Kelson et al. 1996. cAbsolute magnitude calculated from the apparent magnitude listed in the RC3 and the adopted distance to the system. -- 20 -- Table . Spectroscopy Observing Log Slit RA Dec PA T int No. Name () () [deg] Run [sec] NGC   { +   .   .  Jan  (cid:2)  { +   .   .  Jan  (cid:2)  + {   .    .  Jan  (cid:2)  + {   .    .   Jan  (cid:2)  + {    .   . Nov  (cid:2)  +  {   .   .  Jan  (cid:2) NGC   { +   .   .  Nov  (cid:2)  { +   .   .  Nov  (cid:2)  {  {   .   .  Nov  (cid:2)  { +   .   .  Nov  (cid:2)  { {   .   .  Jan  (cid:2)  { +   .   .  Jan  (cid:2)  {  +   .   . Nov  (cid:2)  { +   .   . Nov  (cid:2) { {   .    .  Nov  (cid:2)  { +   .   .   Jan  (cid:2)  + +   .    .  Nov  (cid:2)  + {   .   .  Nov  (cid:2)  + +   .   .   Nov  (cid:2)  + {   .   . Nov  (cid:2)  + {   .   . Nov  (cid:2) NGC   + {   . {  .  Nov  (cid:2) NGC   { {  .  {  . Nov  (cid:2)  + {  . {  . Nov  (cid:2)  + +   . {  . Nov  (cid:2)  + +   . {  .  Nov  (cid:2)  + {  . {  . Nov  (cid:2)  + {  . {  .  Nov  (cid:2)  + +  . {  .  Nov  (cid:2)  + {  . {  . Nov  (cid:2) NGC   { {   . {  . Nov  (cid:2)  { {   . {  .  Nov  (cid:2)  { +   .  {  . Nov  (cid:2)  + {   . {  .  Nov  (cid:2)  + +    . {  . Nov  (cid:2)  + {   . {  . Nov  (cid:2)  + +   . {  .  Nov  (cid:2)  + {   . {  . Nov  (cid:2) NGC   { {   .   .  Jan  (cid:2)  { +   .   .  Jan  (cid:2)  { {   .   .  Jan  (cid:2)  + {    .   .  Jan  (cid:2)  + {    .   .  Jan  (cid:2)  + {   .   .  Jan  (cid:2) -- 21 -- Table . (continued) Slit RA Dec PA T int No. Name () () [deg] Run [sec] NGC   { +  .   .  Nov  (cid:2)  { {  .   .   Nov  (cid:2)  { {  .   .  Nov  (cid:2)  { {   .   .  Nov  (cid:2)  + {  .    .   Nov  (cid:2)  + {   .    .  Nov  (cid:2)  + {   .   .  Nov  (cid:2)  { {  .   .  Nov  (cid:2) NGC  IC   { {  .   .  Nov  (cid:2) NGC    { {  .   .  Jan  (cid:2)  + +  .   . Jan  (cid:2) NGC   { {   .   .  Jan  (cid:2)  { {   .   .  Jan  (cid:2)  + +   .   .  Jan  (cid:2)  + +   .   .   Jan  (cid:2)  +  {    .    .   Jan  (cid:2) NGC    { +   .   . May  (cid:2)  { +   .    . May  (cid:2)  { {    .    . May  (cid:2)  { {   .    . May  (cid:2)  { +   .   . May  (cid:2)  { {   .    . May  (cid:2)  + {    .   . May  (cid:2)  + {   .   . May  (cid:2) + {   .    . May  (cid:2)  + +   .   . May  (cid:2)  +  {   .   . May  (cid:2)  + +   .   .  May  (cid:2) NGC   { +   .   . May  (cid:2)  { {    .    . May  (cid:2)  {  {   .    .  May  (cid:2)  { +   .   . May  (cid:2)  {  +    .   .  May  (cid:2)  { {   .   . May  (cid:2)  { +   .   . May  (cid:2)  { {   .   . May  (cid:2) { {   .   . May  (cid:2) + +   .   .   May  (cid:2)  + {    .   . May  (cid:2) Table . HII Region Line Strengths O(cid:11)sets Slit [OII] [NeIII] [OIII] [OI] [SIII] H(cid:11) [NII] [SII] [ArIII] EW E-W N-S No. +    +    + +  c (H(cid:12) ) H (cid:12) NGC  + {  . (cid:6). . (cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). . (cid:6).  + {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). . (cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  a + {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  + {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). . (cid:6). .(cid:6).  + {  .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). . (cid:6). .(cid:6). .(cid:6).  + {  . (cid:6). . (cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6).  { +  . (cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). . (cid:6).  { +  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). . (cid:6). .(cid:6).  + {  .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  { +  .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  { +  .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  + {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  + {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). . (cid:6). .(cid:6). .(cid:6).  { +  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  + {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). . (cid:6). . (cid:6). .(cid:6).  + {  .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  + {  . (cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  { +  .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).   -- 2 2 -- +  {  .(cid:6). . (cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). . (cid:6).  NGC  + {  . (cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  { +  .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6). .(cid:6). .(cid:6). . (cid:6). . (cid:6). .(cid:6). .(cid:6).  { +  .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  + {  . (cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). .(cid:6). . (cid:6). .  (cid:6). . (cid:6). .(cid:6). .(cid:6).  + {  . (cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  + {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) . (cid:6). .  (cid:6). .(cid:6). .(cid:6). .(cid:6).  + {  .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  + {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  + {  .(cid:6). .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). . (cid:6). .(cid:6). .(cid:6).  { { .(cid:6). .(cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6). . (cid:6). . (cid:6). .(cid:6). .(cid:6).  { { .(cid:6). .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). . (cid:6). .(cid:6). .(cid:6).  + {  .(cid:6). .(cid:6). .(cid:6). . (cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). . (cid:6). .(cid:6). . (cid:6).  { +  .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). . (cid:6). . (cid:6). .(cid:6). .(cid:6).  + +  .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  Table . (continued) O(cid:11)sets Slit [OII] [NeIII] [OIII] [OI] [SIII] H(cid:11) [NII] [SII] [ArIII] EW E-W N-S No. +    +    + +  c (H(cid:12) ) H (cid:12) { +  .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  { {  .(cid:6). .(cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  { +  .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6). .(cid:6). .(cid:6). . (cid:6). .(cid:6). .(cid:6). . (cid:6).  { +  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). . (cid:6). .(cid:6). . (cid:6).  + +  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  + +  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  + +  .(cid:6). . (cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). . (cid:6).  { {  .(cid:6). . (cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). . (cid:6). .(cid:6).  {  {  .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). . (cid:6). .(cid:6).  {  {  .(cid:6). .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  {  {  . (cid:6). .(cid:6). .(cid:6). . (cid:6). .(cid:6). . (cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6).  + {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  { {  .(cid:6). . (cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  + {  .(cid:6). .(cid:6). . (cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  + {  .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  { +  . (cid:6). . (cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6).  + {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  + +  . (cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). . (cid:6). .(cid:6).  { +  . (cid:6). . (cid:6). . (cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  + {  .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6).  + +  . (cid:6). .(cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). . (cid:6). .(cid:6).  { +  .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). . (cid:6). .(cid:6). .(cid:6).  { +  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  { +  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  { +  .(cid:6). .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6).  { +  .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  { +  . (cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  { +  . (cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). . (cid:6). .(cid:6).  { +  . (cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  { +  .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  + {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6).  a NGC  NGC  + +  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6). . (cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  + {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  a -- 2 3 -- Table . (continued) O(cid:11)sets Slit [OII] [NeIII] [OIII] [OI] [SIII] H(cid:11) [NII] [SII] [ArIII] EW E-W N-S No. +    +    + +  c (H(cid:12) ) H (cid:12) + {  . (cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  a + +  . (cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). .(cid:6). .(cid:6). . (cid:6). .(cid:6). . (cid:6). .(cid:6).  { {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  { +  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  + +   .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). . (cid:6).  + {  .(cid:6). (cid:1) (cid:1) (cid:1) .  (cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).   + +   .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6). .(cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6). . (cid:6).  + +  .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  + {  . (cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6). .(cid:6). . (cid:6). .(cid:6).  + +  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  + {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). . (cid:6).  +  {  . (cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  + {  .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). . (cid:6).  + +  . (cid:6). . (cid:6). .(cid:6). .(cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6). . (cid:6). .(cid:6).  { {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6).  .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). + {  .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  a { {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  a { {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). a a NGC  { +  .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). . (cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  + {  .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  a + +  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). . (cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  a + +  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  a { {  .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  a + +  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). . (cid:6). (cid:1) (cid:1) (cid:1) . (cid:6).  a + {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  a { {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6).  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  a + {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  + +  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  a { {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). . (cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  a + {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6).  + {  .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). . (cid:6). .(cid:6).  + {  .(cid:6). .(cid:6). .  (cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  + {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) . (cid:6). . (cid:6). .(cid:6). .(cid:6). . (cid:6).  a NGC  -- 2 4 -- Table . (continued) O(cid:11)sets Slit [OII] [NeIII] [OIII] [OI] [SIII] H(cid:11) [NII] [SII] [ArIII] EW E-W N-S No. +    +    + +  c (H(cid:12) ) H (cid:12) + {  . (cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). . (cid:6). .(cid:6). . (cid:6).  + {   .(cid:6). . (cid:6). .(cid:6).  (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  + {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). . (cid:6). .(cid:6). .(cid:6).  + {  .(cid:6). .(cid:6). . (cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).   + {  .(cid:6). .(cid:6). .(cid:6).  .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  { +  .(cid:6). .(cid:6). . (cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  { {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  { {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  { +  . (cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  + {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). . (cid:6). .(cid:6). .(cid:6).  { {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6).  { {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  { {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6).  { {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  NGC  { {  . (cid:6). .(cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  + {  . (cid:6). .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). . (cid:6). .(cid:6).  + {  .(cid:6). .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  { {  . (cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  + {  .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6). .(cid:6). . (cid:6). .(cid:6).  -- 2 5 -- { {  .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  { {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  { {  .(cid:6). .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). . (cid:6).  + {  .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  + {  .(cid:6). .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6). . (cid:6). .(cid:6). .(cid:6).  { {  .(cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  + {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6). . (cid:6). .(cid:6). .(cid:6).  + {  .(cid:6). .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  + {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). . (cid:6). .(cid:6). . (cid:6). .(cid:6). .(cid:6).  + {   .(cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  { +  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  .(cid:6). .(cid:6). . (cid:6). . (cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  { +  . (cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). . (cid:6).  IC  + {  .(cid:6). .(cid:6). .(cid:6).  .(cid:6). .(cid:6). . (cid:6). .(cid:6). . (cid:6). .(cid:6). .(cid:6).  { {  .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) . (cid:6). . (cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  Table . (continued) O(cid:11)sets Slit [OII] [NeIII] [OIII] [OI] [SIII] H(cid:11) [NII] [SII] [ArIII] EW E-W N-S No. +    +    + +  c (H(cid:12) ) H (cid:12) { {  .(cid:6). .(cid:6). .(cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  + +  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6).  (cid:1) (cid:1) (cid:1) .(cid:6).  a NGC  + +  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). .(cid:6). . (cid:6). .(cid:6).  { {  .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  { {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  NGC   { {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). . (cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  { {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). .(cid:6).  (cid:1) (cid:1) (cid:1) .(cid:6).  a { {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  { {  . (cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  a + +   .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  + +  . (cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  + +  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .  (cid:6). .(cid:6). .(cid:6). .(cid:6).  + +  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). . (cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  a + +  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6).  NGC  { {  . (cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  { {  .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). . (cid:6). . (cid:6).  a { {  . (cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  + {  . (cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  a + +  .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  a + {  .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). . (cid:6). .(cid:6).  + +  .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). . (cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  a +  {   .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  + +  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  { +  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). . (cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  a { +  . (cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  + {  . (cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6).  (cid:1) (cid:1) (cid:1) .(cid:6).  a { {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6).  a { {  .(cid:6). . (cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  { {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6).  { {  . (cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). . (cid:6). .(cid:6).  {  {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  NGC   -- 2 6 -- Table . (continued) O(cid:11)sets Slit [OII] [NeIII] [OIII] [OI] [SIII] H(cid:11) [NII] [SII] [ArIII] EW E-W N-S No. +    +    + +  c (H(cid:12) ) H (cid:12) { {  .(cid:6). . (cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  + {  .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  + {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). . (cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).   + {  .(cid:6). .(cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). . (cid:6). . (cid:6). .(cid:6).  { {  .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6).  + { .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6). . (cid:6). .(cid:6). .(cid:6).  { {   .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). . (cid:6). .(cid:6).  + +  .(cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6).  +  {  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  . (cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). . (cid:6). .(cid:6).  + +  . (cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). . (cid:6). .(cid:6). .(cid:6).  { +  .(cid:6). .(cid:6). . (cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6).  { +  .(cid:6). . (cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6). . (cid:6). .(cid:6). .(cid:6). . (cid:6).  { +  .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  a NGC  { +  . (cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6).  { +  . (cid:6). .(cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6). . (cid:6). .(cid:6). . (cid:6). .(cid:6).  { {  .(cid:6). .(cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6).  + {   .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). . (cid:6).  -- 2 7 -- { {  . (cid:6). . (cid:6). . (cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  { {  .(cid:6). (cid:1) (cid:1) (cid:1) . (cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). . (cid:6). .(cid:6). .(cid:6).  { +  . (cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  {  {  .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). . (cid:6). .(cid:6).  {  +  .(cid:6). .(cid:6). . (cid:6). (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6). . (cid:6). .(cid:6). .(cid:6).  { +  .(cid:6). .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  + +  .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6). .(cid:6).  + +  .(cid:6). .(cid:6). .(cid:6). .(cid:6). (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). .(cid:6). .(cid:6). .(cid:6).  a Only (cid:21) detected{ sum derived assuming (cid:21)/(cid:21)  = . Table . HII Region Ratios O(cid:11)sets r Slit [OIII] [SII] Log Log Log Log Log E-W N-S [arcsec] No. ratio ratio ([OII]+[OIII]) ([OIII]/[OII]) ([NII]/H(cid:11)) ([NII]/[OII]) ([SII]/[OII]) NGC  + { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). {.(cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). + { .  (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). + {  .  (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). {. (cid:6). { + .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {. (cid:6). { + .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). { + .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {. (cid:6). {.(cid:6). {. (cid:6). {. (cid:6). { +  .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). + { .  (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). {. (cid:6). { + .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). {. (cid:6). + {  .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {. (cid:6). + {  .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). + {  .  (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). { +  .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {. (cid:6). {. (cid:6). +  { .  (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). NGC  + { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). { + .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). { + .  (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). {. (cid:6). {.(cid:6). {.(cid:6). {.(cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). {. (cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {. (cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). {.(cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {. (cid:6). {. (cid:6). {. (cid:6). { { . (cid:6) . (cid:6). .(cid:6). .(cid:6). {. (cid:6). {. (cid:6). {.(cid:6). { { . (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {. (cid:6). {. (cid:6). {.(cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). { + .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). + +  .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). {. (cid:6). {. (cid:6). {.(cid:6). -- 2 8 -- Table . (continued) O(cid:11)sets r Slit [OIII] [SII] Log Log Log Log Log E-W N-S [arcsec] No. ratio ratio ([OII]+[OIII]) ([OIII]/[OII]) ([NII]/H(cid:11)) ([NII]/[OII]) ([SII]/[OII]) { + .  (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {. (cid:6). { { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). {. (cid:6). {. (cid:6). {.(cid:6). { + .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {. (cid:6). {.(cid:6). {. (cid:6). {.(cid:6). { + .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). + + .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). + +  .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). + + .  (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {. (cid:6). { { .  (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). .(cid:6). {. (cid:6). {. (cid:6). {.(cid:6). {  { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). . (cid:6). {. (cid:6). {.(cid:6). {. (cid:6). {  { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). {. (cid:6). {.(cid:6). {. (cid:6). {  { .  (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). .(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {. (cid:6). {. (cid:6). {.(cid:6). {. (cid:6). { { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). .(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). {. (cid:6). {.(cid:6). {.(cid:6). {.(cid:6). { + .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). {.(cid:6). + {  .  (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). .(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). + +  .  (cid:1) (cid:1) (cid:1) . (cid:6). . (cid:6). .(cid:6). {. (cid:6). {. (cid:6). {. (cid:6). { + .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {. (cid:6). {. (cid:6). {.(cid:6). + + .  (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). { +  .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). { + .  (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). {.(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). { +  .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {. (cid:6). {. (cid:6). {.(cid:6). {.(cid:6). { + .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). {.  (cid:6). {.(cid:6). {. (cid:6). { +  .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). { +  .  (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {. (cid:6). { + .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). { + .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). {. (cid:6). { +  .  (cid:6)  .(cid:6). .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {. (cid:6). NGC  + {  .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {. (cid:6). {. (cid:6). {.(cid:6). {.(cid:6). NGC  + + .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {. (cid:6). {. (cid:6). {.(cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {. (cid:6). {.(cid:6). {. (cid:6). {.(cid:6). -- 2 9 -- Table . (continued) O(cid:11)sets r Slit [OIII] [SII] Log Log Log Log Log E-W N-S [arcsec] No. ratio ratio ([OII]+[OIII]) ([OIII]/[OII]) ([NII]/H(cid:11)) ([NII]/[OII]) ([SII]/[OII]) + { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). + + .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {. (cid:6). {. (cid:6). {.(cid:6). {.(cid:6). { { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {. (cid:6). {. (cid:6). { + .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). + +  .  (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). {.(cid:6). {. (cid:6). {. (cid:6). {.(cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). {.(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). + +  .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). + + .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). + + .  (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). {.(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). +  { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {. (cid:6). {.(cid:6). {. (cid:6). {. (cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). + + .   (cid:6) .(cid:6). .(cid:6). .(cid:6). {. (cid:6). {. (cid:6). {.(cid:6). NGC  { { .  (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). {.(cid:6). {. (cid:6). .(cid:6). {.(cid:6). + {  .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). { { .  (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). { { .  (cid:1) (cid:1) (cid:1) .(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). . (cid:6). .(cid:6). { + .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {. (cid:6). {. (cid:6). .(cid:6). {.(cid:6). + {  .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {. (cid:6). + + .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). + + .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). { { .  (cid:1) (cid:1) (cid:1) .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). .(cid:6). {.(cid:6). + + .  (cid:1) (cid:1) (cid:1) .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). . (cid:6). .(cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). {.(cid:6). {.(cid:6). {. (cid:6). .(cid:6). .(cid:6). { { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {. (cid:6). {.(cid:6). .(cid:6). {.(cid:6). + +  .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). .(cid:6). {.(cid:6). { { .  (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). {.(cid:6). {.(cid:6). . (cid:6). {.(cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {. (cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). .(cid:6). {.(cid:6). {.(cid:6). {. (cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). .(cid:6). {. (cid:6). {. (cid:6). {. (cid:6). + {  .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). {.(cid:6). NGC  -- 3 0 -- Table . (continued) O(cid:11)sets r Slit [OIII] [SII] Log Log Log Log Log E-W N-S [arcsec] No. ratio ratio ([OII]+[OIII]) ([OIII]/[OII]) ([NII]/H(cid:11)) ([NII]/[OII]) ([SII]/[OII]) + { .  (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {. (cid:6). + {  .  (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). .(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {. (cid:6). {.  (cid:6). {.(cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {. (cid:6). {. (cid:6). {. (cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). { + .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {. (cid:6). {. (cid:6). { { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {. (cid:6). {. (cid:6). { { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). {. (cid:6). { + .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). {. (cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). { { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {. (cid:6). {. (cid:6). {. (cid:6). { { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {. (cid:6). {. (cid:6). {.(cid:6). {. (cid:6). { {  .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). { {  .  (cid:1) (cid:1) (cid:1) . (cid:6). . (cid:6). {.(cid:6). {. (cid:6). {.(cid:6). {. (cid:6). NGC  { { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). {. (cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {. (cid:6). {. (cid:6). {.(cid:6). {.(cid:6). { { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {. (cid:6). { { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). { { .  (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). { { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {. (cid:6). {.(cid:6). {. (cid:6). {.(cid:6). + { .  (cid:6) .(cid:6). .(cid:6). .(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). { { .  (cid:6) .(cid:6). . (cid:6). . (cid:6). {. (cid:6). {.(cid:6). {.(cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). {.(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). .(cid:6). {. (cid:6). {. (cid:6). {.(cid:6). + { .  (cid:6) .(cid:6). . (cid:6). .(cid:6). {.(cid:6). {. (cid:6). {. (cid:6). + {  .  (cid:6) .(cid:6). . (cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). { + .  (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). {. (cid:6). {.(cid:6). {.(cid:6). {.(cid:6). { + .  (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). .(cid:6). {.(cid:6). {. (cid:6). {. (cid:6). IC  + { (cid:1) (cid:1) (cid:1)  (cid:6) .(cid:6). . (cid:6). .(cid:6). {.(cid:6). {.(cid:6). {. (cid:6). { { (cid:1) (cid:1) (cid:1)  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). . (cid:6). {.(cid:6). {. (cid:6). {.(cid:6). -- 3 1 -- Table . (continued) O(cid:11)sets r Slit [OIII] [SII] Log Log Log Log Log E-W N-S [arcsec] No. ratio ratio ([OII]+[OIII]) ([OIII]/[OII]) ([NII]/H(cid:11)) ([NII]/[OII]) ([SII]/[OII]) { { (cid:1) (cid:1) (cid:1)  (cid:6) . (cid:6). . (cid:6). .(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). + + (cid:1) (cid:1) (cid:1)  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {. (cid:6). + + (cid:1) (cid:1) (cid:1)  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). { { (cid:1) (cid:1) (cid:1)  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). { { (cid:1) (cid:1) (cid:1)  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). NGC   NGC  { { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). { { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). { { .  (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). {.(cid:6). { { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). + +  .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). {.(cid:6). + + .  (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). + + .  (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). {. (cid:6). {. (cid:6). {.(cid:6). {.(cid:6). + + .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). {.(cid:6). + + .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). NGC  { { .  (cid:1) (cid:1) (cid:1) .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). . (cid:6). . (cid:6). { { .  (cid:1) (cid:1) (cid:1) .(cid:6). {.(cid:6). {.(cid:6). {. (cid:6). .(cid:6). {.(cid:6). { { .  (cid:1) (cid:1) (cid:1) . (cid:6). {. (cid:6). {. (cid:6). {.(cid:6). .(cid:6). {. (cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). .(cid:6). {.(cid:6). + + .  (cid:1) (cid:1) (cid:1) .(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). .(cid:6). {.(cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). .(cid:6). {. (cid:6). + + .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). .(cid:6). {.(cid:6). +  {  .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). {.(cid:6). + + .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {. (cid:6). {. (cid:6). {.(cid:6). {.(cid:6). { + .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). {.(cid:6). { + .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). { { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). {. (cid:6). { { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). { { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {. (cid:6). {. (cid:6). {.(cid:6). {.(cid:6). { { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {  { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). NGC   -- 3 2 -- Table . (continued) O(cid:11)sets r Slit [OIII] [SII] Log Log Log Log Log E-W N-S [arcsec] No. ratio ratio ([OII]+[OIII]) ([OIII]/[OII]) ([NII]/H(cid:11)) ([NII]/[OII]) ([SII]/[OII]) { { .  (cid:6) .(cid:6). . (cid:6). .(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). + { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). + { .  (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {. (cid:6). { { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). + { . (cid:6) .(cid:6). .(cid:6). . (cid:6). {.(cid:6). {.(cid:6). {.(cid:6). { {  .  (cid:1) (cid:1) (cid:1) . (cid:6). .  (cid:6). .(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). + + .  (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6). {.(cid:6). {. (cid:6). {. (cid:6). {. (cid:6). +  { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). + + .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). { + .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). .(cid:6). {. (cid:6). {.(cid:6). {. (cid:6). { + .  (cid:6) .(cid:6). . (cid:6). .(cid:6). {.(cid:6). {. (cid:6). {. (cid:6). NGC  { + .  (cid:1) (cid:1) (cid:1) .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). .(cid:6). .(cid:6). { + .  (cid:1) (cid:1) (cid:1) . (cid:6). .(cid:6). {. (cid:6). {.(cid:6). .(cid:6). {.(cid:6). { + .  (cid:6) .(cid:6). .  (cid:6). . (cid:6). {. (cid:6). {.(cid:6). {.(cid:6). { { .  (cid:6) .(cid:6). .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {. (cid:6). + {   .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). -- 3 3 -- { { .  (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). {.(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). { { .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {. (cid:6). {. (cid:6). {.(cid:6). { + .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). {  {  .  (cid:6) .(cid:6). .(cid:6). .(cid:6). {. (cid:6). {.(cid:6). {. (cid:6). {  +  .  (cid:1) (cid:1) (cid:1) .(cid:6). . (cid:6). {.(cid:6). {. (cid:6). {.(cid:6). {.(cid:6). { + .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {. (cid:6). {.(cid:6). {. (cid:6). + + .  (cid:6) .(cid:6). .(cid:6). .(cid:6). {. (cid:6). {. (cid:6). {. (cid:6). + + .  (cid:1) (cid:1) (cid:1) .(cid:6). .(cid:6). {.(cid:6). {.(cid:6). {.(cid:6). {. (cid:6). -- 34 -- Table 5. HII Region Abundances T(O++) (O/H) (O/H) 12+Log(O/H) Log(N/O) Log(Ne/O) Log(S/O) Log(Ar/O) ZKH EP84 NGC 0628 Offsets Slit E-W N-S [arcsec] No. r0 +081 -- 140 +062 -- 158 +047 -- 172 +044 -- 175 +180 -- 008 +178 -- 052 -- 086 +186 -- 075 +200 +203 -- 041 -- 073 +203 -- 069 +208 +254 -- 043 +256 -- 041 -- 232 +112 +262 -- 041 +264 -- 037 +265 -- 041 -- 227 +148 +292 -- 020 +002 -- 002 -- 005 +000 -- 008 +000 +010 -- 004 +015 -- 005 +030 -- 008 +036 -- 010 +042 -- 011 +087 -- 031 -- 012 -- 066 -- 047 -- 058 +135 -- 016 -- 109 +062 +137 +011 -- 137 +056 -- 145 -- 006 -- 080 +087 -- 114 +087 +161 +015 +174 +018 +182 +019 -- 187 -- 017 -- 192 -- 018 -- 192 -- 018 -- 198 -- 013 +209 -- 006 -- 200 -- 020 +217 -- 006 +156 -- 114 172.6 178.8 185.9 187.7 195.5 203.3 216.9 224.6 226.4 226.5 229.6 281.3 283.0 283.0 289.4 290.7 292.6 297.2 318.0 3.6 5.2 8.4 11.1 16.0 31.1 37.5 43.5 94.1 120.5 124.2 137.2 137.6 149.4 153.1 154.8 156.3 170.2 177.0 192.5 201.5 205.3 211.1 211.1 214.1 216.5 220.8 224.9 226.4 5700 ± 500 4800 ± 500 5500 ± 500 6200 ± 500 5100 ± 500 5300 ± 500 6400 ± 500 6700 ± 500 6000 ± 500 6200 ± 500 6900 ± 500 7000 ± 700 8500 ± 800 6400 ± 500 7500 ± 500 7400 ± 500 8300 ± 500 7400 ± 500 3 3 3 3 4 4 2 2 5 2 2 6 5 1 5 6 5 1 6 12000 ± 1500 5100 ± 500 5100 ± 500 5500 ± 500 5500 ± 500 5100 ± 500 5400 ± 500 5800 ± 500 5800 ± 500 6900 ± 500 8890 ± 520 7400 ± 500 7900 ± 500 7900 ± 1500 8900 ± 2000 7000 ± 500 10 10 10 10 10 10 10 10 12 9 9 12 6 13 6 5 10300 ± 2000 8300 ± 800 7 8500 ± 800 7 8700 ± 2000 13 8900 ± 1000 13 13 12500 ± 1500 7800 ± 500 8300 ± 800 8500 ± 800 5 3 5 3 10000 ± 2000 8600 ± 1600 7500 ± 500 15 5 15 11200 ± 1500 14 7700 ± 600 8.88±0.12 8.95±0.19 8.90±0.12 8.82±0.10 8.94±0.14 8.95±0.14 8.79±0.10 8.69±0.10 8.85±0.11 8.80±0.10 8.73±0.10 8.69±0.11 8.48±0.10 8.70±0.10 8.60±0.10 8.65±0.10 8.55±0.10 8.64±0.10 8.10±0.10 8.98±0.15 8.94±0.14 8.89±0.12 8.87±0.12 8.96±0.15 8.86±0.13 8.82±0.11 8.84±0.11 8.71±0.10 8.48±0.05 8.60±0.10 8.58±0.10 8.48±0.23 8.38±0.25 8.68±0.10 8.23±0.17 8.48±0.10 8.51±0.10 8.38±0.24 8.28±0.11 8.01±0.10 8.61±0.10 8.56±0.10 8.53±0.10 8.35±0.18 8.39±0.19 8.65±0.10 8.19±0.10 8.59±0.10 · · · · · · · · · · · · · · · · · · · · · -- 1.15±0.17 -- 0.68±0.34 -- 1.83±0.26 -- 2.16±0.21 -- 1.52±0.30 -- 1.14±0.24 -- 1.65±0.26 -- 2.22±0.22 -- 1.17±0.18 -- 1.81±0.25 -- 2.11±0.18 -- 1.12±0.15 -- 1.15±0.21 -- 1.61±0.27 -- 2.13±0.25 -- 1.14±0.20 -- 0.60±0.40 -- 1.80±0.27 -- 2.16±0.24 -- 1.19±0.14 -- 0.81±0.26 -- 1.26±0.25 -- 2.20±0.17 -- 1.23±0.13 -- 1.63±0.24 -- 2.18±0.16 -- 1.61±0.25 -- 2.03±0.19 -- 0.97±0.16 -- 1.18±0.15 -- 1.67±0.24 -- 2.20±0.17 -- 1.15±0.12 -- 0.73±0.23 -- 1.51±0.21 -- 2.20±0.15 -- 1.16±0.16 -- 1.50±0.25 -- 2.31±0.20 -- 1.38±0.24 -- 2.20±0.16 -- 1.20±0.13 -- 1.56±0.25 -- 1.47±0.14 -- 1.43±0.23 -- 2.14±0.14 -- 1.18±0.11 -- 1.55±0.24 -- 2.24±0.14 -- 1.26±0.11 -- 1.22±0.14 -- 1.62±0.24 -- 2.28±0.18 -- 1.39±0.11 -- 1.80±0.24 -- 2.37±0.14 -- 1.39±0.13 -- 0.70±0.25 -- 1.87±0.24 -- 2.29±0.16 · · · · · · · · · · · · · · · · · · · · · · · · · · · NGC 0925 · · · · · · · · · · · · · · · · · · · · · · · · · · · -- 1.63±0.28 -- 2.09±0.26 -- 1.09±0.21 -- 1.45±0.27 -- 2.25±0.25 -- 1.14±0.21 -- 1.49±0.26 -- 2.23±0.22 -- 1.18±0.18 -- 1.44±0.26 -- 2.24±0.22 -- 1.16±0.18 -- 1.52±0.28 -- 2.09±0.26 -- 1.03±0.21 -- 1.45±0.26 -- 2.14±0.23 -- 1.19±0.19 -- 1.49±0.25 -- 2.13±0.20 -- 1.23±0.17 -- 1.18±0.17 -- 1.56±0.25 -- 2.23±0.20 -- 1.30±0.13 -- 0.28±0.23 -- 1.61±0.24 -- 2.18±0.15 -- 1.34±0.08 -- 0.61±0.13 -- 1.62±0.12 -- 2.11±0.10 -- 1.37±0.11 -- 0.25±0.20 -- 1.42±0.24 -- 2.27±0.13 -- 1.42±0.13 -- 0.25±0.25 -- 1.56±0.24 -- 2.30±0.16 -- 1.25±0.30 -- 0.50±0.69 -- 1.72±0.34 -- 2.14±0.37 -- 1.43±0.35 -- 2.41±0.43 -- 1.37±0.33 -- 1.35±0.12 -- 1.79±0.24 -- 2.27±0.15 -- 1.28±0.23 -- 0.70±0.52 -- 1.60±0.43 -- 2.17±0.29 -- 1.42±0.14 -- 0.43±0.26 -- 1.53±0.22 -- 2.26±0.17 -- 1.44±0.24 -- 2.35±0.16 -- 1.45±0.13 -- 1.55±0.36 -- 2.36±0.43 -- 1.46±0.35 -- 1.64±0.16 -- 1.39±0.25 -- 1.43±0.12 -- 0.89±0.23 -- 1.74±0.19 -- 2.22±0.16 -- 1.41±0.10 -- 0.67±0.18 -- 1.66±0.23 -- 2.34±0.12 -- 1.49±0.14 -- 0.68±0.26 -- 1.72±0.24 -- 2.36±0.17 -- 1.44±0.13 -- 0.68±0.25 -- 1.51±0.21 -- 2.37±0.16 -- 1.40±0.25 -- 0.72±0.56 -- 1.67±0.42 -- 2.32±0.31 -- 1.52±0.26 -- 1.55±0.30 -- 2.18±0.33 -- 1.34±0.11 -- 0.49±0.19 -- 1.68±0.23 -- 2.28±0.13 -- 1.48±0.15 -- 0.48±0.29 -- 1.52±0.25 -- 2.45±0.19 -- 1.46±0.12 -- 1.38±0.20 -- 2.33±0.14 · · · · · · · · · · · · · · · · · · · · · 8.94 9.06 8.95 8.84 9.02 9.02 8.78 8.72 8.87 8.82 8.71 8.67 8.52 8.77 8.62 8.62 8.50 8.63 8.45 9.05 9.01 8.93 8.93 9.07 8.94 8.87 8.89 8.70 8.47 8.65 8.57 8.74 8.60 8.68 8.57 8.57 8.49 8.67 8.70 8.54 8.58 8.50 8.51 8.44 8.63 8.60 8.43 8.60 8.80 9.01 8.81 8.66 8.94 8.93 8.60 8.54 8.70 8.64 8.53 8.49 8.36 8.59 8.44 8.44 8.35 8.45 7.81 9.00 8.92 8.78 8.79 9.03 8.79 8.70 8.73 8.51 8.32 8.47 8.40 8.56 8.43 8.50 7.70 8.40 8.34 8.48 8.51 7.72 8.41 8.34 8.35 7.81 8.45 8.42 7.82 8.42 -- 35 -- Table 5 -- Continued T(O++) (O/H) (O/H) 12+Log(O/H) Log(N/O) Log(Ne/O) Log(S/O) Log(Ar/O) ZKH EP84 Offsets Slit E-W N-S [arcsec] No. r0 -- 220 +004 +221 -- 006 +052 +130 -- 250 +019 +206 -- 114 +019 +143 -- 174 +140 -- 262 +011 -- 272 +016 -- 274 +010 -- 285 +023 -- 159 +162 -- 308 +035 -- 149 +177 -- 022 +227 228.7 229.0 249.1 255.6 256.8 258.7 269.3 270.2 279.2 283.1 291.1 294.3 313.2 314.0 396.8 3 7100 ± 500 15 11200 ± 1500 11 11000 ± 1500 9300 ± 2000 7400 ± 500 2 14 11 10000 ± 1000 4 11000 ± 1500 9800 ± 1000 1 1 9700 ± 1000 2 10300 ± 2000 9700 ± 1000 1 7800 ± 700 4 8100 ± 800 1 8600 ± 1600 4 8 9440 ± 1380 8.67±0.10 8.27±0.11 8.20±0.10 8.31±0.23 8.65±0.10 8.39±0.10 8.19±0.10 8.30±0.10 8.29±0.10 8.29±0.16 8.30±0.10 8.49±0.10 8.50±0.10 8.38±0.18 8.39±0.09 · · · · · · -- 1.50±0.12 -- 0.71±0.22 -- 1.41±0.20 -- 2.33±0.14 -- 1.27±0.15 -- 1.58±0.25 -- 2.35±0.19 -- 1.57±0.15 -- 0.68±0.30 -- 1.68±0.25 -- 2.21±0.19 -- 1.27±0.29 -- 0.77±0.70 -- 1.31±0.62 -- 2.15±0.38 -- 1.45±0.11 -- 1.62±0.24 -- 2.29±0.14 -- 1.47±0.12 -- 0.52±0.23 -- 1.71±0.21 -- 2.32±0.15 -- 1.43±0.15 -- 0.65±0.30 -- 1.62±0.24 -- 2.31±0.19 -- 1.45±0.24 -- 1.52±0.14 -- 1.55±0.15 -- 1.34±0.24 -- 2.44±0.18 -- 1.41±0.23 -- 0.36±0.51 -- 1.50±0.40 -- 2.40±0.29 -- 1.58±0.15 -- 0.51±0.25 -- 1.25±0.23 -- 2.39±0.17 -- 1.28±0.24 -- 1.51±0.14 -- 1.43±0.14 -- 1.34±0.24 -- 2.10±0.17 -- 1.51±0.26 -- 0.63±0.40 -- 1.60±0.47 -- 2.40±0.32 -- 1.53±0.13 -- 0.87±0.24 -- 1.93±0.24 -- 2.23±0.16 · · · · · · · · · · · · · · · · · · 8.67 8.31 8.44 8.66 8.62 8.32 8.51 8.50 8.56 8.48 8.56 8.63 8.61 8.69 8.46 8.49 7.91 7.81 8.48 8.45 7.90 7.76 7.76 7.71 7.78 7.70 8.45 8.43 8.50 7.79 +111 -- 088 159.5 1 4700 ± 300 9.04±0.13 -- 0.82±0.15 · · · -- 1.56±0.26 · · · 9.11 9.13 NGC 1232 NGC 1068 +062 +004 +075 -- 001 +075 -- 012 +059 +058 -- 103 -- 021 -- 103 +022 +021 +091 +004 -- 101 +057 +098 +056 +106 +147 -- 030 +117 -- 084 +099 -- 101 +093 -- 110 +147 -- 065 +135 +114 -- 001 -- 000 +016 -- 001 -- 019 -- 009 -- 020 -- 000 -- 001 +026 +009 -- 026 +022 +018 +026 +018 -- 017 -- 031 +051 +012 -- 055 -- 011 +047 -- 031 +047 +046 -- 055 -- 032 +058 -- 054 62.2 75.0 76.2 88.9 105.8 106.0 106.5 115.9 126.0 133.9 151.0 151.6 152.4 156.7 164.8 187.9 1.2 19.5 23.9 24.2 26.6 29.6 30.7 34.8 35.8 61.7 66.3 67.7 69.9 71.0 93.3 4200 ± 200 4 4900 ± 300 5 5200 ± 300 5 4300 ± 200 4 5100 ± 400 1 5900 ± 400 1 5800 ± 400 3 5900 ± 400 2 7000 ± 700 4 6600 ± 500 4 5800 ± 500 8 6500 ± 500 7 6700 ± 500 6 7300 ± 800 6 6400 ± 500 8 7 10670 ± 650 3 4 2 2 3 4 4 5 2 7 1 6 7 1 8 4000 ± 200 4900 ± 300 5200 ± 300 3300 ± 200 4300 ± 200 4100 ± 200 3900 ± 200 4100 ± 200 3500 ± 200 3550 ± 200 4900 ± 300 3900 ± 200 4000 ± 200 4500 ± 200 6200 ± 400 9.07±0.10 8.94±0.11 8.87±0.10 9.08±0.10 8.94±0.14 8.83±0.10 8.85±0.10 8.80±0.10 8.67±0.10 8.74±0.10 8.85±0.10 8.77±0.10 8.71±0.10 8.64±0.11 8.75±0.10 8.20±0.05 9.18±0.11 8.91±0.13 8.94±0.11 9.30±0.21 9.10±0.10 9.10±0.10 9.12±0.11 9.11±0.11 9.26±0.16 9.24±0.17 8.95±0.11 9.12±0.10 9.12±0.10 9.04±0.10 8.74±0.10 · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · -- 1.49±0.24 -- 0.95±0.13 -- 1.45±0.25 -- 1.07±0.14 -- 1.42±0.24 -- 1.00±0.13 -- 0.93±0.30 -- 2.38±0.14 -- 0.88±0.12 -- 1.55±0.27 -- 2.34±0.22 -- 1.00±0.26 -- 1.61±0.24 -- 2.45±0.18 -- 1.07±0.13 -- 1.19±0.24 -- 2.24±0.16 -- 1.02±0.14 -- 1.20±0.23 -- 2.27±0.16 -- 1.03±0.13 -- 1.13±0.31 -- 2.30±0.20 -- 1.12±0.16 -- 1.61±0.24 -- 1.21±0.14 -- 1.56±0.25 -- 2.40±0.19 -- 1.05±0.17 -- 1.68±0.24 -- 2.44±0.16 -- 1.18±0.14 -- 1.71±0.24 -- 2.30±0.16 -- 1.16±0.13 -- 1.56±0.25 -- 2.39±0.21 -- 1.16±0.17 -- 1.58±0.24 -- 2.43±0.19 -- 1.22±0.14 -- 1.09±0.08 -- 0.96±0.11 -- 1.74±0.11 -- 2.23±0.10 · · · · · · · · · · · · · · · · · · · · · · · · · · · NGC 1637 -- 0.55±0.14 -- 0.97±0.16 -- 0.82±0.15 -- 0.61±0.20 -- 0.79±0.12 -- 0.90±0.13 -- 0.96±0.14 -- 0.92±0.13 -- 0.79±0.17 -- 0.63±0.17 -- 0.89±0.14 -- 0.88±0.14 -- 0.88±0.14 -- 0.78±0.12 -- 1.14±0.12 · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · -- 1.40±0.25 -- 1.42±0.26 -- 1.61±0.26 -- 1.46±0.31 -- 1.53±0.24 -- 1.46±0.25 -- 1.50±0.25 -- 1.48±0.25 -- 1.52±0.28 -- 1.42±0.28 -- 1.51±0.25 -- 1.56±0.24 -- 2.16±0.17 -- 1.52±0.25 -- 1.44±0.24 -- 1.50±0.24 -- 2.06±0.15 · · · · · · · · · · · · · · · · · · 9.18 9.02 8.98 9.16 9.03 8.88 8.91 8.88 8.70 8.76 8.90 8.80 8.74 8.66 8.78 8.53 9.23 9.01 9.01 9.35 9.18 9.18 9.20 9.18 9.30 9.31 9.06 9.22 9.18 9.15 8.82 9.25 8.92 8.87 9.23 8.95 8.72 8.77 8.72 8.52 8.57 8.74 8.61 8.55 8.48 8.60 7.73 9.34 8.92 8.92 9.52 9.25 9.25 9.29 9.25 9.44 9.46 9.01 9.32 9.26 9.21 8.64 -- 36 -- Table 5 -- Continued T(O++) (O/H) (O/H) 12+Log(O/H) Log(N/O) Log(Ne/O) Log(S/O) Log(Ar/O) ZKH EP84 NGC 2403 Offsets Slit E-W N-S [arcsec] No. r0 +186 -- 177 +178 -- 203 +176 -- 211 +166 -- 229 +356 -- 195 +360 -- 190 +377 -- 163 +376 -- 106 -- 377 +104 -- 105 -- 218 -- 105 -- 224 -- 381 +082 +383 -- 056 -- 104 -- 256 -- 425 -- 002 -- 423 -- 010 -- 421 -- 017 -- 051 -- 009 +055 -- 026 +058 -- 022 -- 021 -- 045 +032 -- 056 -- 001 -- 079 -- 005 -- 080 -- 041 -- 090 +037 -- 115 +078 -- 107 -- 068 -- 079 +080 -- 110 +017 -- 119 +089 -- 107 +116 -- 098 -- 145 +104 -- 151 +102 +002 -- 006 -- 028 -- 007 -- 033 -- 007 +001 +001 -- 018 -- 009 +021 +011 -- 025 -- 013 262.4 286.4 295.2 315.9 414.2 417.5 435.2 446.8 448.9 452.0 461.5 464.5 482.7 511.2 585.8 590.3 594.1 60.8 61.5 63.4 65.3 68.8 97.2 99.8 130.0 136.4 137.6 137.7 141.3 142.3 142.8 152.3 178.4 182.2 · · · · · · · · · · · · · · · · · · · · · -- 062 -- 085 -- 065 -- 073 -- 067 -- 061 -- 060 -- 100 123.2 123.7 124.9 127.8 7600 ± 1400 4 8200 ± 1500 4 6200 ± 500 4 7100 ± 500 4 9100 ± 1000 5 8000 ± 1500 5 5 8500 ± 1500 6 10300 ± 2000 2 8000 ± 1500 7800 ± 1500 3 7400 ± 1400 3 8600 ± 1600 2 9100 ± 2000 6 3 8500 ± 1500 8900 ± 2000 1 9300 ± 2000 1 1 9000 ± 1000 5900 ± 400 2 7300 ± 800 6 6500 ± 500 6 6200 ± 500 3 7300 ± 800 6 6800 ± 500 4 7300 ± 800 4 8000 ± 800 4 9710 ± 490 5 5 7900 ± 500 3 11480 ± 840 6 10000 ± 1000 9700 ± 2000 5 6 9180 ± 2670 6 10970 ± 270 1 9300 ± 1000 1 12000 ± 1500 8.74±0.24 8.40±0.21 8.72±0.11 8.66±0.10 8.44±0.10 8.44±0.18 8.42±0.18 8.23±0.17 8.46±0.20 8.33±0.22 8.44±0.23 8.43±0.19 8.35±0.23 8.34±0.18 8.33±0.24 8.26±0.21 8.21±0.11 8.84±0.10 8.64±0.11 8.75±0.10 8.79±0.10 8.62±0.11 8.75±0.10 8.65±0.11 8.56±0.10 8.33±0.04 8.61±0.10 8.19±0.05 8.38±0.10 8.28±0.20 8.49±0.13 8.22±0.03 8.40±0.10 8.16±0.10 1 13120 ± 460 1 13100 ± 1500 1 13210 ± 1110 8.06±0.03 7.87±0.10 7.98±0.05 1 1 1 1 1 1 1 1 6000 ± 1000 6900 ± 500 6000 ± 400 8000 ± 800 8.95±0.25 8.69±0.10 8.83±0.10 8.60±0.10 6200 ± 500 5500 ± 500 5800 ± 500 5600 ± 500 8.78±0.10 8.79±0.16 8.81±0.11 8.79±0.14 · · · · · · -- 1.16±0.30 -- 0.50±0.69 -- 1.50±0.34 -- 2.26±0.38 -- 1.23±0.27 -- 0.81±0.62 -- 1.87±0.32 -- 2.12±0.35 -- 1.31±0.25 -- 2.27±0.22 -- 1.28±0.15 -- 1.38±0.12 -- 1.52±0.24 -- 2.40±0.14 -- 1.35±0.14 -- 0.80±0.28 -- 1.54±0.25 -- 2.27±0.17 -- 1.36±0.28 -- 0.20±0.67 -- 1.32±0.30 -- 2.39±0.34 -- 1.39±0.25 -- 1.04±0.56 -- 1.69±0.30 -- 2.22±0.31 -- 1.53±0.23 -- 0.72±0.52 -- 1.70±0.46 -- 2.31±0.29 -- 1.40±0.28 -- 0.63±0.67 -- 1.69±0.32 -- 2.22±0.35 -- 1.48±0.31 -- 1.48±0.32 -- 1.47±0.26 -- 1.56±0.31 -- 1.58±0.25 -- 1.58±0.32 -- 1.56±0.29 -- 1.55±0.15 -- 1.70±0.33 -- 1.62±0.33 -- 1.69±0.30 -- 2.24±0.33 -- 1.94±0.34 -- 2.37±0.39 -- 1.68±0.30 -- 2.36±0.31 -- 1.61±0.34 -- 2.35±0.41 -- 1.71±0.32 -- 2.45±0.37 -- 1.49±0.25 -- 2.12±0.19 · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · NGC 2805 · · · · · · · · · -- 1.14±0.13 -- 0.62±0.25 -- 1.21±0.24 -- 2.33±0.16 -- 1.27±0.17 -- 0.50±0.35 -- 1.62±0.25 -- 2.36±0.21 -- 1.25±0.14 -- 0.47±0.26 -- 1.56±0.24 -- 2.31±0.16 -- 1.18±0.15 -- 1.69±0.24 -- 2.23±0.18 -- 1.32±0.17 -- 1.57±0.26 -- 2.51±0.22 -- 1.28±0.13 -- 0.65±0.24 -- 1.31±0.22 -- 2.28±0.15 -- 1.29±0.17 -- 1.47±0.29 -- 2.43±0.20 -- 1.35±0.15 -- 0.43±0.28 -- 1.49±0.24 -- 2.43±0.17 -- 1.34±0.08 -- 0.70±0.11 -- 1.57±0.10 -- 2.32±0.09 -- 1.37±0.13 -- 0.76±0.25 -- 1.61±0.25 -- 2.37±0.17 -- 1.31±0.08 -- 0.62±0.13 -- 1.67±0.11 -- 2.35±0.10 -- 1.38±0.13 -- 1.42±0.24 -- 2.48±0.16 -- 1.30±0.27 -- 0.96±0.62 -- 1.60±0.31 -- 2.24±0.34 -- 1.53±0.19 -- 1.64±0.30 -- 2.42±0.22 -- 1.37±0.06 -- 0.62±0.07 -- 1.64±0.07 -- 2.35±0.08 -- 1.42±0.14 -- 1.32±0.13 -- 0.75±0.25 -- 1.72±0.22 -- 2.32±0.17 IC 2458 -- 1.43±0.07 -- 0.68±0.08 -- 1.68±0.08 -- 2.28±0.08 -- 1.68±0.14 -- 1.48±0.09 -- 0.95±0.13 -- 1.63±0.13 -- 2.39±0.11 -- 1.37±0.24 -- 1.45±0.24 · · · · · · · · · · · · · · · · · · NGC 2820 -- 1.30±0.34 -- 1.22±0.14 -- 1.08±0.13 -- 1.25±0.15 NGC 2903 -- 0.96±0.15 -- 1.16±0.19 -- 1.09±0.17 -- 1.12±0.18 · · · · · · · · · · · · · · · · · · · · · · · · · · · -- 1.54±0.36 -- 1.49±0.24 -- 1.66±0.24 -- 2.31±0.16 -- 1.56±0.25 · · · · · · · · · -- 1.47±0.25 -- 1.31±0.28 -- 1.50±0.25 -- 2.40±0.24 -- 1.34±0.27 · · · · · · 8.41 8.79 8.81 8.66 8.44 8.78 8.65 8.55 8.71 8.93 8.85 8.60 8.61 8.72 8.68 8.70 8.79 8.93 8.65 8.77 8.85 8.64 8.73 8.65 8.56 8.53 8.56 8.41 8.37 8.61 8.36 8.45 8.44 8.32 8.32 8.70 8.48 8.61 8.90 8.69 8.53 8.81 8.90 8.88 8.89 8.28 8.60 8.63 8.48 8.30 8.59 8.47 8.39 8.52 8.78 8.67 8.43 8.44 8.53 8.49 8.52 8.61 8.78 8.47 8.58 8.67 8.46 8.54 8.46 8.40 8.37 8.39 7.84 8.26 8.44 8.25 7.80 8.30 7.91 7.90 7.56 7.78 8.43 8.74 8.50 8.37 8.63 8.75 8.72 8.73 -- 37 -- Table 5 -- Continued (O/H) (O/H) 12+Log(O/H) Log(N/O) Log(Ne/O) Log(S/O) Log(Ar/O) ZKH EP84 8.71±0.10 8.78±0.10 8.70±0.11 8.74±0.12 8.79±0.12 9.27±0.15 9.23±0.15 9.19±0.11 9.18±0.10 9.24±0.14 9.12±0.10 9.11±0.11 9.09±0.10 9.10±0.19 9.05±0.10 9.05±0.10 9.01±0.22 8.97±0.11 8.71±0.10 8.90±0.13 8.87±0.12 8.82±0.10 8.48±0.21 8.57±0.10 8.29±0.20 8.39±0.23 8.41±0.17 8.33±0.17 8.39±0.17 8.31±0.17 8.38±0.22 8.06±0.10 8.33±0.05 9.23±0.17 9.13±0.12 8.41±0.03 8.22±0.04 8.49±0.21 8.30±0.19 8.38±0.24 8.28±0.23 8.01±0.03 8.44±0.19 8.09±0.10 7.92±0.03 7.95±0.10 -- 1.19±0.13 -- 1.19±0.14 -- 1.26±0.14 -- 1.25±0.16 -- 1.24±0.17 NGC 3184 -- 0.70±0.18 -- 0.67±0.16 -- 0.74±0.15 -- 0.78±0.14 -- 0.81±0.16 -- 0.89±0.14 -- 0.79±0.12 -- 0.99±0.13 -- 1.08±0.28 -- 1.06±0.13 -- 0.97±0.12 -- 1.21±0.28 -- 1.07±0.15 -- 1.02±0.16 -- 1.05±0.18 -- 1.06±0.18 -- 1.09±0.15 NGC 4395 · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · -- 1.72±0.24 -- 2.17±0.16 -- 1.79±0.24 -- 2.17±0.17 -- 1.42±0.25 -- 2.02±0.20 -- 1.45±0.26 -- 1.57±0.26 -- 2.02±0.22 · · · · · · · · · · · · · · · · · · -- 1.55±0.27 -- 1.55±0.27 -- 2.16±0.23 -- 1.61±0.25 -- 1.52±0.25 -- 1.52±0.25 -- 1.48±0.24 -- 2.25±0.19 -- 1.46±0.25 -- 1.40±0.24 -- 2.18±0.16 -- 1.70±0.31 -- 2.26±0.33 -- 1.46±0.25 -- 1.56±0.24 -- 1.62±0.33 -- 1.49±0.25 -- 2.24±0.25 -- 1.62±0.25 -- 2.19±0.20 -- 1.84±0.26 -- 2.12±0.22 -- 1.71±0.26 -- 2.23±0.23 -- 1.60±0.25 -- 2.27±0.18 · · · · · · · · · · · · · · · · · · -- 1.51±0.32 -- 2.28±0.37 -- 1.48±0.35 -- 1.50±0.31 -- 1.29±0.25 -- 1.44±0.27 -- 0.84±0.62 -- 1.46±0.56 -- 2.16±0.34 -- 1.60±0.32 -- 0.69±0.80 -- 1.47±0.34 -- 2.40±0.41 -- 1.44±0.27 -- 0.57±0.47 -- 1.61±0.30 -- 2.25±0.31 -- 1.59±0.23 -- 0.58±0.52 -- 1.54±0.29 -- 2.26±0.30 -- 1.41±0.25 -- 0.91±0.56 -- 1.22±0.52 -- 2.19±0.31 -- 1.66±0.25 -- 1.82±0.44 -- 2.32±0.31 -- 1.48±0.31 -- 0.71±0.74 -- 1.59±0.33 -- 2.21±0.39 -- 1.46±0.13 -- 0.83±0.25 -- 1.69±0.24 -- 2.17±0.16 -- 1.55±0.09 -- 0.75±0.14 -- 1.65±0.14 -- 2.28±0.11 · · · NGC 5457 · · · · · · · · · · · · · · · · · · -- 1.53±0.28 -- 1.63±0.25 -- 0.57±0.17 -- 0.80±0.15 -- 1.17±0.06 -- 0.69±0.07 -- 1.38±0.20 -- 2.03±0.07 -- 1.39±0.07 -- 0.68±0.10 -- 1.65±0.21 -- 2.32±0.09 -- 1.49±0.30 -- 1.43±0.63 -- 2.28±0.36 -- 1.47±0.26 -- 0.60±0.40 -- 1.46±0.30 -- 2.36±0.33 -- 1.54±0.33 -- 1.70±0.34 -- 2.54±0.42 -- 1.48±0.31 -- 0.74±0.74 -- 1.58±0.33 -- 2.42±0.39 -- 1.44±0.06 -- 0.77±0.07 -- 1.66±0.07 -- 2.30±0.07 -- 1.55±0.27 -- 0.67±0.60 -- 1.56±0.30 -- 2.44±0.33 -- 1.43±0.14 -- 0.23±0.27 -- 1.47±0.24 -- 2.24±0.17 -- 1.45±0.06 -- 0.75±0.07 -- 1.69±0.39 -- 2.46±0.08 -- 1.51±0.13 -- 0.51±0.24 -- 1.53±0.24 -- 2.44±0.16 8.74 8.79 8.80 8.85 8.89 9.35 9.29 9.28 9.24 9.28 9.21 9.20 9.16 8.94 9.16 9.13 8.78 9.07 8.72 8.96 8.95 8.84 8.71 8.51 8.58 8.57 8.58 8.32 8.71 8.48 8.55 8.52 8.42 9.33 9.21 8.32 8.55 8.74 8.76 8.61 8.72 8.50 8.64 8.56 8.55 8.73 8.55 8.60 8.61 8.68 8.72 9.51 9.43 9.41 9.34 9.42 9.30 9.28 9.22 8.80 9.22 9.18 8.60 9.03 8.53 8.82 8.82 8.66 8.53 8.36 8.41 8.40 8.41 7.90 8.52 8.33 8.39 7.74 8.29 9.48 9.30 8.23 8.38 8.56 8.57 8.43 8.54 7.76 8.46 7.70 7.72 7.53 Offsets Slit E-W N-S [arcsec] No. r0 +171 +196 +171 +226 +171 +232 +171 +236 +171 +243 -- 058 -- 007 -- 064 -- 006 -- 080 -- 005 +085 -- 004 +079 +035 +059 -- 079 +074 +064 +092 -- 093 +005 +135 -- 002 +136 -- 017 +137 +111 -- 102 -- 119 -- 121 -- 113 -- 127 -- 101 -- 137 -- 110 -- 130 -- 095 -- 142 +061 -- 029 +099 -- 029 +002 -- 127 -- 075 -- 117 +088 -- 119 -- 167 -- 093 +118 +206 +195 -- 159 +200 +141 -- 226 +237 -- 272 +186 -- 075 +028 -- 106 +028 -- 347 +276 -- 459 -- 053 +271 -- 393 -- 276 -- 417 -- 037 -- 532 -- 250 +484 -- 398 -- 436 -- 291 +489 -- 499 +300 +010 +885 +005 +887 326.6 336.3 338.7 340.3 343.3 60.9 66.9 83.3 87.8 91.5 98.7 104.5 130.8 140.1 140.6 141.7 150.8 181.3 181.5 181.5 181.8 182.0 68.5 105.3 128.8 144.7 148.0 200.5 246.7 252.8 256.7 328.0 332.0 84.7 114.3 465.2 474.1 498.6 500.2 541.2 566.6 590.6 593.1 610.6 900.6 902.8 T(O++) 6700 ± 500 6400 ± 500 6400 ± 500 6000 ± 500 5700 ± 500 3400 ± 200 3700 ± 200 3800 ± 200 3900 ± 200 3600 ± 200 4000 ± 200 4300 ± 200 4100 ± 200 5000 ± 600 4200 ± 200 4400 ± 200 5400 ± 300 4700 ± 300 6000 ± 500 5500 ± 500 5500 ± 500 6200 ± 500 2 2 2 2 2 2 2 2 4 4 5 4 5 3 3 3 5 1 1 1 1 1 7800 ± 500 8 8100 ± 800 8 9700 ± 2000 7 8900 ± 2000 4 8930 ±5790 9 1200 3 10300 ± 2000 10 8500 ± 1500 11 10000 ± 2000 12 9100 ± 2000 2 12000 ± 1500 1 10050 ± 770 3600 ± 200 7 4250 ± 250 7 4 9880 ± 170 2 10450 ± 480 7600 ± 1400 11 8600 ± 1600 6 8900 ± 2000 8 5 9100 ± 2000 3 12750 ± 320 5 8300 ± 1500 1 10300 ± 1200 10 13630 ± 440 10 12200 ± 1500 -- 38 -- Table 6. Radial Oxygen Abundance Gradients No. of Central Galaxy HII Regions 12+log(O/H) [dex/r25] NGC 0628 NGC 0925 NGC 1068 NGC 1232 NGC 1637 NGC 2403 NGC 2805 NGC 2903 NGC 3184 NGC 4395 NGC 5457 26 53 13 16 15 40 17 36 32 14 53 9.46 ± 0.11 8.79 ± 0.06 9.26 ± 0.02 9.49 ± 0.12 9.18 ± 0.07 8.70 ± 0.07 9.11 ± 0.12 9.22 ± 0.06 9.50 ± 0.04 8.48 ± 0.13 9.29 ± 0.05 -- 0.99 ± 0.14 -- 0.45 ± 0.08 -- 0.30 ± 0.07 -- 1.31 ± 0.20 -- 0.34 ± 0.15 -- 0.77 ± 0.14 -- 1.05 ± 0.17 -- 0.56 ± 0.09 -- 0.78 ± 0.07 -- 0.32 ± 0.19 -- 1.52 ± 0.09 Gradients [dex/rd] -- 0.23 ± 0.03 -- 0.12 ± 0.02 -- 0.16 ± 0.04 -- 0.36 ± 0.05 -- 0.06 ± 0.02 -- 0.14 ± 0.02 -- 0.42 ± 0.07 -- 0.13 ± 0.02 -- 0.20 ± 0.02 -- 0.13 ± 0.07 -- 0.27 ± 0.02 [dex/kpc] -- 0.067 ± 0.009 -- 0.032 ± 0.006 -- 0.020 ± 0.004 -- 0.056 ± 0.009 -- 0.068 ± 0.028 -- 0.074 ± 0.013 -- 0.049 ± 0.008 -- 0.048 ± 0.008 -- 0.083 ± 0.007 -- 0.037 ± 0.022 -- 0.049 ± 0.003 -- 39 -- Fig. 1. -- R -- band and continuum subtracted Hα images of the galaxies in the sample. The positions of the 2′ slits are illustrated on the Hα images. The images are oriented with north to the top and east to the left. -- 40 -- Fig. 1 continued -- 41 -- Fig. 1 continued -- 42 -- Fig. 1 continued -- 43 -- Fig. 2. -- Representative spectra of H II regions in NGC 1232. The inner H II regions (a) and (b) are clearly high abundance, with strong [N II] and [S II] and weak [O III] lines. The outermost H II region (c) is low abundance, with weak [N II] and strong [O III] lines. -- 44 -- Fig. 3. -- Reddening as a function of radius for each galaxy. In each panel the radius has been normalized by the isophotal radius of that galaxy. -- 45 -- Fig. 4. -- Emission -- line diagnostic diagram for [N II]. In this and subsequent figures, H II regions from NGC 628 are denoted by upright triangles, NGC 925 by squares, NGC 1068 by crosses, NGC 1232 by downward triangles, NGC 1637 by diamonds, NGC 2403 by upward indented triangles, NGC 2805 by pentagons, IC 2458 by six pointed stars, NGC 2820 by five pointed stars, NGC 2903 by indented crosses, NGC 3184 by hexagons, NGC 4395 by four pointed open stars, and NGC 5457 by five pointed open stars. (a) The H II region sequence formed by [N II] and [O III] normalized by Balmer lines. A theoretical curve from Baldwin et al. (1981) is superposed. (b) The H II region sequence formed by R23 vs. [N II]/Hα. In both panels, NGC 4395 -- 003 -- 003 is marked by an arrow. -- 46 -- Fig. 5. -- Emission -- line diagnostic diagrams for [S II]. The symbols are the same as in Figure 4. (a) The H II region sequence formed by [S II] and [O III] normalized by the Balmer lines. (b) The H II region sequence formed by R23 vs. [S II]/Hα. In both panels, NGC 4395 -- 003 -- 003 is marked by an arrow. -- 47 -- Fig. 6. -- The density sensitive line ratio [S II] λ6717/6731 as a function of R23. The symbols are the same as in Figure 4. The dashed line indicates the maximum value of the [S II] ratio in the low density limit. The majority of the H II regions fall within the low density limit. -- 48 -- 2 1 0 0.1 0.01 -1 0.001 0.0001 -2 -1.5 -1 -0.5 0 0.5 1 Fig. 7. -- The model grid of the R23 relation from McGaugh (1991). The locations of the H II regions are marked; the symbols are the same as in Figure 4. The ambiguity between the high and low abundance sides of the surface can be resolved based on the strength of the [N II] lines. -- 49 -- 1 0.5 0 -0.5 -1 -1.5 -1 -0.5 0 0.5 1 Fig. 8. -- The [N II]/[O II] diagnostic diagram. The symbols are the same as in Figure 4. The one outlying point is NGC 4395 -- 003 -- 003. Low values of [N II]/[O II] and high values of [O II]+[O III]/Hβ are found for the outermost H II regions in NGC 925, NGC 2403, NGC 2805, NGC 4395, and NGC 5457, indicating that these are low abundance H II regions. 9.5 9 8.5 8 7.5 7 -- 50 -- -2 -1 0 Fig. 9. -- A semi -- empirical technique for determining oxygen abundances. The line shows the least squares fit to the low (12 + log (O/H) < 8.2) and high (8.8 < 12 + log (O/H) < 9.1) abundance H II regions and HII regions from a dwarf galaxy sample (van Zee et al. 1997). Also plotted are data points for H II regions in NGC 4214 (Kobulnicky & Skillman 1996), NGC 1569 (Kobulnicky & Skillman 1997), and I Zw 18 (Skillman & Kennicutt 1993). -- 51 -- Fig. 10. -- Comparison between R23 calibrations. The large scatter for 12 + log (O/H) < 8.5 is due to the turnover in the R23 relation which has not been accounted for in the ZKH (Zaritsky et al. 1994) calibration. The six high abundance points which are offset from both the EP84 (Edmunds & Pagel 1984) and ZKH calibrations are from H II regions where the [N II]/Hα calibration was used instead of the pure R23 relation. Two of these points are from the edge -- on spiral NGC 2820, where the combination of uncertain extinction corrections and the superpositions of multiple H II regions make the empirical calibration highly uncertain. While there are systematic differences between the calibrations, at high abundances the ZKH calibration is very similar to the calibration used here. In contrast, the calibration of EP84 would result in much steeper abundance gradients. -- 52 -- Fig. 11. -- The relative enrichment of the α elements. In all three panels, the mean value is denoted by a solid line. The solar value (Anders & Grevesse 1989) is represented by an open circle, where the symbol size is indicative of the errors. (a) Log (Ne/O) as a function of oxygen abundance. (b) Log (S/O) as a function of oxygen abundance. (c) Log (Ar/O) as a function of oxygen abundance. -- 53 -- Fig. 12. -- The observed oxygen abundance gradients in all 11 spiral galaxies. The filled symbols represent H II regions from the present study. The open circles represent data from the literature: NGC 628 -- McCall et al. (1985); NGC 925 -- Zaritsky et al. (1994); NGC 1068 -- Evans & Dopita (1987), Oey & Kennicutt (1993); NGC 2403 -- McCall et al. (1985), Fierro et al. (1986), Garnett et al. (1997); NGC 2903 -- McCall et al. (1985), Zaritsky et al. (1994); NGC 3184 -- Zaritsky et al. (1994); NGC 4395 -- McCall et al. (1985); NGC 5457 -- Kennicutt & Garnett (1996). The solid lines illustrate the derived oxygen abundance gradients.
astro-ph/0012527
1
0012
2000-12-29T05:18:10
On the source of QPO of the black hole candidate GRS1915+105: some new observations and their interpretation
[ "astro-ph" ]
A few classes of the light curve of the black hole candidate GRS 1915+105 have been analyzed in detail. We discover that unlike the previous findings, QPOs occasionally occur even in the so-called `On' or softer states. Such findings may require a revision of the accretion/wind scenario of the black hole candidates. We conjecture that considerable winds which is produced in `Off' states, cool down due to Comptonization, and falls back to the disk and creating an excess accretion rate to produce the so-called `On' state. After the drainage of the excess matter, the disk goes back to the `Off' state. Our findings strengthen the shock oscillation model for QPOs.
astro-ph
astro-ph
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (2000) Printed 27 October 2018 On the source of QPO of the black hole candidate GRS1915+105: some new observations and their interpretation. Anuj Nandi1, Sivakumar G. Manickam1, A.R. Rao2 and Sandip K. Chakrabarti1,3 1 S.N. Bose National Centre for Basic Sciences, JD-Block, Sector III, Salt Lake, Calcutta 700098, India; 2 Tata Institute of Fundamental Research, Homi Bhabha Road, Colaba, Mumbai 400005 3 Centre for Space Physics, 114/v/1A Raja S.C. Mullick Rd., Calcutta 700047, India [email protected], [email protected], [email protected], [email protected] Accepted . Received ; in original form ABSTRACT A few classes of the light curve of the black hole candidate GRS 1915+105 have been analyzed in detail. We discover that unlike the previous findings, QPOs occa- sionally occur even in the so-called 'On' or softer states. Such findings may require a revision of the accretion/wind scenario of the black hole candidates. We conjecture that considerable winds which is produced in 'Off' states, cool down due to Comptonization, and falls back to the disk and creating an excess accretion rate to produce the so-called 'On' state. After the drainage of the excess matter, the disk goes back to the 'Off' state. Our findings strengthen the shock oscillation model for QPOs. Key words: Black Hole Physics -- Accretion Disks -- Outflows -- Stars:individual (GRS1915+105) MNRAS (to appear) 1 INTRODUCTION The black hole candidate GRS1915+105 continues to excite astrophysicists by having one of the most, if not the most mysterious light curves. In a matter of days, the light curve changes its character, and within each day, photon counts show variations of a factor of two to five or more (Morgan, Remillard & Greiner, 1997; Belloni et al, 1997; Paul et al. 1998; Manickam & Chakrabarti 1999). The power density spectra shows clear evidence of quasi-periodic oscillations (QPOs) with frequency ranging from ∼ 0.001Hz to 67Hz (see, e.g., Morgan, Remillard & Greiner, 1997; Chakrabarti & Manickam 2000, hereafter CM00, and references therein). The origin of the quasi-periodic oscillations cannot be a complete mystery since there are very clear evidence that sub-Keplerian flows, which must occur in a black hole accre- tion, exhibit very wild time-dependent behaviour including large amplitude shock oscillations. This is true, especially when the infall time matches with the cooling time (Molteni, Sponholz & Chakrabarti, 1996; Ryu, Chakrabarti & Molteni 1997; Paul et al, 1998; Remillard et al, 1999ab; Muno, Mor- gan & Remillard, 1999). This became more evident when it was observed that the QPOs are absent in low energy photons but is very strong at high energies, supporting the view that the photons participating in QPOs originate at hot post-shock flow (CM00, Rao et al. 2000). It has also been observed that QPOs are seen mostly when the photon counts are low ('Off' states) and QPO is very weak or absent in the On states when the photon counts are about two to five times larger. In the present Letter, by On and Off states, we would mean the high and low count states of the light curves. This nomenclature is not related to spectral states. Because of the presence of the On and the Off states in the light curves, explanation of QPOs using radial os- cillation of the centrifugal pressure supported shocks alone (which produce the right amplitude and frequency for the QPOs) cannot be the complete story. The very fact that the Off states terminate and the On states emerge in almost regular basis (but not exactly regular) gives rise to another time scale which must, at the same time, be dependent on disk/jet parameters. This is because Belloni et al. (2000) found that at least twelve types of light curves are seen, and within each type, the duration and behaviour of the On and the Off (if both exist) states were not at all fixed. For in- stance, ρ class exhibits extremely regular light curves (Taam, Chen & Swank, 1997; Vilhu & Nevalainen, 1998) with broad Off- or low-count states and very narrow, spiky, high-count or On-states (see, Fig. 2b below). Light curves in the ν class 2 Anuj Nandi, Sivakumar G. Manickam, A.R. Rao and Sandip K. Chakrabarti is similar to those of ρ, but are highly irregular (see, Fig. 2a below). In λ class, both Off and On states are of longer time duration (Belloni et al. 1997) while in κ class these durations are relatively shorter (see, Fig. 5 below). Nandi, Manickam and Chakrabarti (2000) using a completely differ- ent procedure divided the light curves into four fundamental types (Hard, Soft, Semi-soft and Intermediate). The Inter- mediate class shows On/Off transitions. Chakrabarti (1999) and Chakrabarti & Manickam (2000) showed that outflow rates from the centrifugal barrier must play a major role in deciding the duration, as the wind matter at least up to the sonic sphere can be Comptonized and cooled down. Most of this cold matter (below the sonic surface of the cooler wind) can fall back on the immediate vicinity of the disk increasing its accretion rate temporarily while the rest must separate out at a supersonic speed. As this excess matter drains out, the Off state together with the QPO start appearing again. In this Letter, we analyze light curves in detail and find that QPOs may be observed at very unlikely times in the light curves. For instance, we find that very often, there is a sharp peak ('first hiccup') at the onset of the On state and there is a sharp peak ('last hiccup') just prior to going to the Off state. The first peak, though in the On state, very often shows QPOs while the last peak, though the radiation is much harder, does not show a QPO. We also note that the radiation progressively hardens in the On state. In several cases, using data from both the RXTE and Indian X-ray Astronomical Experiment (IXAE), ρ class curves (mostly mini-ρ type) are seen to be peeled off from a κ class. In κ and λ, the On state duration is long and just before going to the Off state, the light curve becomes noisy and oscillating in nature, and the features indicate as though the light curve is made up of 'sums' of ρ types. Only the lower half of the oscillations (local Off states of mini-ρ states) exhibit QPO! We believe that these observations definitely point to the drainage of extra matter in the disk which was accumulated from the wind. One of the problems in analyzing data of one of the most complex objects, such as GRS1915+105 is that one has to ask the right questions and as many of them as pos- sible. Once a paradigm is kept in the back of the mind, ask- ing questions become easier. We therefore concentrate on models which required sub-Keplerian and Keplerian flows simultaneously and where winds are also produced self- consistently. In the next Section, we present some of the 'subtle' observational results which have not been reported by workers before (despite the fact that GRS 1915+105 is probably the most studied object in recent years). In §3, we interpret this observations in terms of the advective disk paradigm. In §4, we draw our conclusions. 2 OBSERVATIONS 2.1 Properties of two peaks in the On state GRS 1915+105 exhibits both broad and narrow On states (Paul et al. 1998; Yadav et al, 1999; Belloni et al. 2000; CM00). When the On state is narrow or spiky, i.e., the du- ration with high photon counts is very small, and very often, just after transition from On to Off state, a second peak is observed. This has also been cursorily reported by Paul et al. (1998). Figure 1 shows details of a narrow section of the Fig. 1: Subdivisions of a quasi-repetitive structure of the light curve of Oct. 15th, 1996 (ID:10408-01-41-00). Relatively hard ra- diations are emitted in the region marked Off with relatively lower count rates. There are typically two major peaks, P1 followed by P2. all-channel light curve corresponding to the RXTE obser- vation of 15th of Oct. 1996 (The observation ID number is 10408-01-41-00.). We denote the peak appearing first as the primary peak (P1 for short) and the peak following P1 as the secondary peak (P2 for short) respectively. The words Off, 'P1' and 'P2' are marked. The entire light curve is al- most a repetition of this section. Figure 2a presents the light curves of this data clearly showing two peaks P1 and P2 in the On state. The time lag between the two peaks is roughly constant on a given day (in this case about 10 seconds) and we do not find any correlation between this lag and the du- ration of the Off state. The light curve is shown in four panels: the panels are drawn for channel energies (from top to bottom) 0 − 5.07keV, 5.43 − 6.88keV, 7.24 − 9.43keV, and 9.79 − 13.09keV respectively (corresponding channels are 0 − 13, 14 − 18, 19 − 25, 26 − 35 respectively). Note that while in low energies, photon counts in P1 is much larger compared to P2, in higher energies they are roughly equal, suggesting that the spectrum of P2 is harder. Taam, Chen and Swank (1997) also noted the existence of these peaks and that P2 is harder compared to P1. A similar light curve is shown in Fig. 2b, for the RXTE observation of 22nd of June, 1997 (see also Vilhu & Nevalainen (1998) and Yadav et al (1999) for displaying this light curve.). The correspond- ing observation ID is 20402-01-34-01. In this case, the time lag is about 4 seconds but other features are similar to those of Fig. 2a. The channel energies are marked on each panel (corresponding channels numbers are 0−13, 14−19, 20−25, 26 − 35 respectively). Figures 3(a-c) show the power density spectra (PDS) of the three regions marked in Fig. 1. While selecting photons from 'P1' and 'P2' regions we took special care that they are not contaminated by the photons from the Off states. Also, to improve statistics, we added data from many peaks over the entire duration of the observation on that day. Note that in the Off state there is a distinct QPO of frequency νqpo = 7.4Hz. Photons in 'P1' also exhibit QPO though it is weaker (νqpo ∼ 5.7Hz). QPO is completely absent from 'P2'. On the source of QPO of the black hole candidate GRS1915+105: some new observations and their interpretation. Fig. 2a: Four panels showing light curves of a part of the observa- tion on Oct. 15th, 1996 (ID: 10408-01-41-00) at different channels (energies are marked). Note that ratio of photon counts in P1 and P2 tends to become unity at higher energies. Fig. 3(a-c): Power Density Spectrum (PDS) of three regions marked in Fig. 1 of the observation dated Oct. 15th, 1996. Several data segments have been added to improve statistics. Note the absence of QPO in P2. Fig. 2b : Four panels showing light curves of a part of the ob- servation on June 22nd, 1997 (ID: 20402-01-34-01) at different channels (energies are marked). We examined the RXTE data of 22nd June, 1997 also, the PDS of which is shown in Fig. 4(a-c). The result is generally the same. The QPO frequency in the Off-state is given by νqpo = 6.3Hz and in P1 peak νqpo = 6.8Hz. We therefore believe that the observed features may be generic. Fig. 4(a-c): Power Density Spectrum (PDS) of three regions marked in Fig. 1 of the observation dated June 22nd, 1997. Several data segments have been added to improve statistics. Note the absence of QPO in P2. 2.2 Properties of On++ state It is generally observed that whenever the duration of On or the high count state is large, light curves become very noisy and count rates start oscillating wildly as the Off state is approached. The later half may be termed as 'On++' state and Manickam & Chakrabarti (1999) showed that this re- gion exhibits a weak QPO. In fact, similar to Fig. 2(a-b) above, where the photons in P2 are harder compared to P1, one finds that On++ state is harder compared to the first half of the On state. This is demonstrated in Fig. 5, 4 Anuj Nandi, Sivakumar G. Manickam, A.R. Rao and Sandip K. Chakrabarti frequency and when the shock is fully developed, the Off state begins with a QPO at 3.09Hz. An easily missed phenomenon in all these light curves is that most of the On states begin with a 'hiccup' or a small peak, which may be likened with 'P1' of ρ or ν class and also end with another 'hiccup' which may be likened with 'P2'. 3 POSSIBLE INTERPRETATIONS OF THESE OBSERVATIONS 3.1 The Paradigm While there is as yet no fully self-consistent model which includes disks, winds and radiative transfer simultaneously, one can collect bits and pieces of the solutions and con- struct a viable model for the system. Chakrabarti (1996) and more recently Chakrabarti (2000a) have discussed such solutions in detail. Fig. 7 shows a cartoon diagram of an accretion/wind system of GRS 1915+105. Generally, it is assumed the the accretion is advective in nature and has a centrifugal barrier dominated region which may or may not have become a fully developed shock (at r = rs) throughout the period of observation. Chakrabarti (1999) showed that the centrifugal barrier dominated boundary layer (CENBOL for short) is not only responsible for the Comptonized ra- diation, but also responsible for the wind/jet formation. A very simple analysis which envisages an isothermal wind at least up to the sonic surface (at r = rc) shows that winds should not be emitted in soft states, and very hard states should have very little winds. If the shock is of intermediate strength, outflow rate is very much high. If the outflow rate is high enough, it can fill the sonic sphere (of size rc ∼ 2 − 3rs; sub-sonic region up to the sonic surface) rapidly till the optical depth due to Compton scattering become larger than unity. Comptonization cools down this region rapidly. CM00 suggests that the duration of the Off state is the time in which this region achieves this threshold of cooling. Since the specific energy and angular momentum of the flow decide shock location and its strength (Chakrabarti 1989), a small variation of the overall viscosity would change the shock location, and therefore cause the variation of the duration and the QPO frequency from one Off state to another. The correlation between duration and frequency based on this consideration has been discussed in CM00. An important bye-product of all these is that as the disk loses matter (and pressure) from the post-shock region in the form of outflows, the shock, as well as the inner edge of the Keplerian disk moves inwards, gradually increasing the QPO frequency. This has been demonstrated by the dynamic power density spectra (e.g., Muno et al., 2000; Trudolyubov et al. 2000). A part of the wind which does not reach escape veloc- ity must return back to the disk and the accretion rate of the disk is temporarily modulated. This has been demon- strated by two dimensional simulations of advective disks (Molteni, Lanzafame & Chakrabarti, 1994; Molteni, Ryu and Chakrabarti, 1996 and Ryu, Chakrabarti & Molteni, 1997). This feedback mechanism becomes more complex in presence of radiative cooling of the outflow. After the wind is cooled down, the sound speed is reduced, and location of the sonic surface comes closer to the black hole, very abruptly. Fig. 5: A part of the light curve from the so-called κ-state is shown. The energy ranges are marked in each panel. Upper four panels are from RXTE and the lower two panels are from IXAE data. Note that in On++ states light curves are more noisy and oscillatory, more so for photons at higher energy. A small arrow indicates a ρ type burst. Presence of mini-ρ bursts within a κ type On states indicates that ρ may be the fundamental type of all burst. where a part of the light curve from the so-called κ-state is shown. The energy ranges are marked in each panel. Upper four channels are from RXTE and the lower two channels are from the Indian X-ray Astronomy Experiment (IXAE) data. The observation ID for RXTE is 20186-03-01-02. In both the cases one second bin-size is chosen. First note that RXTE and IXAE show very similar be- haviours throughout the period of overlap of observations. Second, towards the end of each of the On states (which we term as On++ state for brevity), the light curves are noisy and generally oscillatory in nature. Third, in both the ex- periments, the relative oscillations are increased with the increasing photon energy in On++ state. Fourth, though generally the light curve may be called that of a κ-class (Belloni et al. 2000), several pieces of ρ class is evident. In fact, mini-ρ type light curves are also evident in the On++ states giving clear evidence that the light curve in the ρ-class is more primitive. In the next Section we discuss possible in- terpretations of this important observation. In Fig. 6, we show a part of the light curve on the same day, and draw two boxes (in dotted curves) at the On++ state. The photons from the upper box show no sign of QPO, while the photons from the lower box shows clear evidence of QPO. Since lower box contains photons which are from mini-Off states of the mini-ρ class mentioned above, it is not surprising that this show QPOs. Indeed, while the Off state on this day shows QPO of frequency 3.09Hz, these mini-Off states show QPOs of frequency 6.25Hz. This indicates that towards the end, the shock again starts developing much closer to the black hole giving rise to a higher oscillation On the source of QPO of the black hole candidate GRS1915+105: some new observations and their interpretation. Fig. 6: A part of the light curve from the so-called κ-state is shown. In On++ states, two dotted boxes are drawn and below the PDS of each box is plotted. Note the presence of QPO in lower box, which are nothing but mini-Off states corresponding to mini-ρ bursts. Part of the originally subsonic outflowing matter still re- mains below the new sonic surface and falls back since it loses drive to escape. The rest becomes supersonic (because of reduced sound speed) and separates in the form of blobs. Thus, blobs are produced at the end of the off-states. This conjecture has been tested by multiwavelength observations (Dhawan et al. 2000). During the On states, the shock and the CENBOL are non-existent and hence the QPOs are generally absent. To- wards the end of the On state, in the so-called On++ state, the shocks start forming close to the black hole because of heat generated by excess accretion. The shock then rapidly recedes to a distance consistent with the steady state so- lution as soon as the excess matter is drained out of the disk. This causes the onset of the Off state. This process of receding are regularly seen in numerical simulation of sub- Keplerian flows (Chakrabarti & Molteni, 1993; MLC94). 3.2 Origin of hiccups or P1 and P2 peaks It is clear that since these peaks are separated by a few sec- onds, and generally one has a QPO while the other does not, their origins cannot be the same. From the paradigm described above, one may imagine that P1 forms when catas- trophic cooling of the CENBOL-sonic sphere system takes place. Since P1 is a continuation of the Off state, QPO is thus expected unless the shock is hidden under the cooler wind. P2 is due to the steepening of last bits of excess matter on the disk which is delayed by the viscous time- scale. The viscous time which a ring of matter takes after it leaves the Keplerian disk from a transition radius (rtr, where the flow deviates from a Keplerian disk (Chakrabarti & Titarchuk, 1995), and enters the post-shock region is given 100 )1/2s where α by tvisc ∼ 1 is the Shakura-Sunyaev viscosity parameter, h(r) ∼ csr3/2 = 10( 0.01 α )( 0.1 cs α ( h r )−2 r vKep )2( rtr Fig. 7: Cartoon diagram showing possible physical processes close to a black hole during an average time of the (a) Off and (b) On states. Centrifugal barrier or CENBOL produces winds. Subsonic region is cooled and falls back on the disk while supersonic region separates out as blobs. Excess accretion inner part causes higher counts in On states. is the dimensionless instantaneous height of the disk (in vertical equilibrium) at a radius r (measured in units of Rg = 2GM/c2, the Schwarzschild radius), vKep is the ro- tational velocity of the Keplerian orbit, and cs is the speed of sound in units of velocity of light. As this ring of matter propagates through this region, it is illuminated by soft pho- tons coming out of the Keplerian disk, but since it is outside the shock, its radiations do not participate in the oscillation. On the other hand, since rising side of P1 is in Off state, it shows QPO. Due to Compton cooling the spectrum of P1 is softer. However P2 occurs when the excess matter is almost entirely drained out from the disk. Hence its spectrum is harder. It is to be noted that whether or not shocks would form depend on viscosity. It has been shown (Chakrabarti, 1990; Chakrabarti & Molteni, 1995; Chakrabarti 1996) that there is a critical viscosity parameter αc below which shocks can form and this parameter is about αc ∼ 0.015. Beyond this chosks disappear and Keplerian flows directly enter into the black hole. Out choice of α = 0.01 to explain the time scale is thus consistent with the presence of shocks. In general, as excess matter drains out of the CENBOL, 6 Anuj Nandi, Sivakumar G. Manickam, A.R. Rao and Sandip K. Chakrabarti This work is partly supported by Indian Space Research Organization (ISRO) through a project entitled Quasi Peri- odic Oscillations in Black Holes. AN's work is supported by a DST project entitled Analytical and Numerical Studies of Astrophysical Flows Around Compact Objects. We grate- fully acknowledge using RXTE data from public archive of GSFC/NASA. REFERENCES Belloni, T, M´endez, King, A.R., Van der Klis, M. & Van Paradijs, J. 1997, APJ 488, L109 Belloni, T., Klein, Wolt, M., M´endez, M., van Der Klis, M. and van Paradijs, J. 2000, A & A, 355, 271 Chakrabarti, S.K. 1989, ApJ 347, 365 Chakrabarti, S.K. 1990, M.N.R.A.S. 243, 610 Chakrabarti, S.K. & Molteni, D., 1995, M.N.R.A.S. 272, 80 Chakrabarti, S.K. 1996, Phys. Rep., 266, 229 Chakrabarti, S.K. 1999a, Ind. J. Phys. 73B(6), 931 Chakrabarti, S.K. 1999b, A&A 351, 185, (C99) Chakrabarti, S.K. 2000a, Il Nouvo Cimento (In press). Chakrabarti, S.K. et al. 2000b, Proceedings of 9th Marcel Gross- man meeting (Ed.) R. Ruffini Chakrabarti, S.K. & Molteni, D. 1993, ApJ 417 671 Chakrabarti, S.K. & Manickam, S.G. 2000, ApJL 531, L41 (CM00) Chakrabarti, S.K. & Titarchuk, L.G. 1995, ApJ 455, 623 Dhawan, V., Mirabel, I. F. & Rodriguez, L. F. 2000, ApJ, sub- mitted. Manickam, S.G. & Chakrabarti, S.K. 1999 in Proceedings of Young Astrophysicists of Today's India, Ind. J. Phys. 73(6), 967 Molteni, D., Lanzafame, G. & Chakrabarti, S.K. 1994, ApJ 425, 161 Molteni, D., Sponholz, H. & Chakrabarti, S.K. 1996, ApJ 457, 805 Molteni, D., Ryu, D. & Chakrabarti, S.K. 1996, ApJ 470, 460 Morgan, E. H., Remillard, R. A. & Greiner, J. 1997, ApJ 482, 993 Muno, M.P., Morgan, E.H. & Remillard, R.A. 1999 ApJ 527, 321 Muno M.P., Morgan E.H., Remillard, R. A. 1999, 527, 321. Nandi, A., Manickam, S. & Chakrabarti, S.K., 2000, Ind. J. Phys., 74B(5), 3312 Paul, B., et al. 1998, ApJ 492, L63 Rao, A.R, Naik, S.,Vadawale, S.V. And Chakrabarti, S.K. 2000, A&A, 360, L25 Remillard, R.A. et al., 1999a ApJL 517, L127 Remillard, R. A., Morgan, E.H., McClintock, J.E., Bailyn, C. D., Orosz, J.A. 1999b, ApJ 522, 397 Ryu, D., Chakrabarti & Molteni, D. 1997, ApJ 474, 378 Taam, R., Chen, X.-M. & Swank, J. 1997, ApJ, 485, 83 Trudolyubov S., Churazov E., Gilfanov, M. 1999, A & A 25, 718 Vilhu, O. & Nevalainen, J. 1998, ApJ, 508, L85 Yadav, et al., 1999, ApJ, 517, 935 its optical depth decreases and the spectrum gets harder in the On++ state. This is observed in both RXTE and IXAE data. 3.3 Origin of ρ type light curve One of the reasons we plotted Fig. 5 using the particular re- gion of κ class observations is that it may hold the key of un- derstanding the general light curves. The Figure clearly in- dicates that ρ type regions are peeled off gradually in On++ states of κ class light curve. Arrows in Fig. 5 indicate that the forms of rise and fall are qualitatively similar, but quan- titatively one is a miniature version of a fully developed ρ type of bursts. Details of the modeling is being discussed elsewhere (Chakrabarti et al. 2000b). Briefly, when the entire CENBOL behaves like a single blob, it's light curve is of ρ type: count rate going up in Comptonization time scale and rapidly drops in infall time- scale. However, in presence of strong winds, and subsequent return of matter on the disk, turbulence are generated and one may imagine that each of the turbulent cells after being steepened into small shocks, producing mini-ρ light curves, depending on the shapes and sizes of the turbulent cells. Each mini-shock produces a mini-ρ curve after filing smaller sonic surface. 4 CONCLUDING REMARKS In this Paper, we have pointed out some new and interesting behaviour of the light-curve of the black hole candidate GRS 1915+105. We showed that very often the On states of ρ and ν types of light curves produce more than one peak which behave differently. For instance, the first peak may show QPOs whereas the second peak shows no QPO. The second peak also has a harder spectrum. In the κ and λ classes when the On state duration is longer, the On state shows very noisy and oscillatory behaviour towards the end, which we term as On++ state. What is more, photons from upper- half of these light curves do not show QPOs while those from the lower-half do. Though these observations are related to very small re- gions of the overall light curves, they help us understand the behaviour of matter close to a black hole. If one assumes that the duration of the Off states are determined by the time in which the optical depth of the sonic sphere becomes unity, as CM00 suggested, then many puzzles are resolved. In this picture, part of the cooler matter of the outflowing wind falls back on the CENBOL and the pre-shock disk and its drainage time gives rise to the On states. It is possible that the first catastrophic cooling gives rise to the first hiccup in the On state and the last significant density perturba- tion due to the fallen matter may cause the final hiccup. This can explain why P1 often shows QPO, but P2 does not. Duration of On states in between may vary, depend- ing on drainage time, giving rise to various classes of light curves. Also, towards the end of the drainage period, i.e., in On++ states, the excess matter is depleted and the signs of QPO starts appearing. What is most interesting, mini-Off states show QPO, while mini-On states do not, indicating that states with broader On are possibly made of ρ type bursts.
astro-ph/0207385
2
0207
2002-07-19T05:01:37
Thermal Emission as a Test for Hidden Nuclei in Nearby Radio Galaxies
[ "astro-ph" ]
The clear sign of a hidden quasar inside a radio galaxy is the appearance of quasar spectral features in its polarized (scattered) light. However that observational test requires suitably placed scattering material to act as a mirror, allowing us to see the nuclear light. A rather robust and more general test for a hidden quasar is to look for the predicted high mid-IR luminosity from the nuclear obscuring matter. The nuclear waste heat is detected and well isolated in the nearest narrow line radio galaxy, Cen A. This confirms other indications that Cen A does contain a modest quasar-like nucleus. However we show here that M87 does not: at high spatial resolution, the mid-IR nucleus is seen to be very weak, and consistent with simple synchrotron emission from the base of the radio jet. This fairly robustly establishes that there are "real" narrow line radio galaxies, without the putative accretion power, and with essentially all the luminosity in kinetic form. Next we show the intriguing mid-IR morphology of Cygnus A, reported previously by us and later discussed in detail by Radomski et al. (2002). All of this mid-IR emission is consistent with reprocessing by a hidden quasar, known to exist from spectropolarimetry by Ogle et al. (1997) and other evidence.
astro-ph
astro-ph
Thermal Emission as a Test for Hidden Nuclei in Nearby Radio Galaxies D. Whysong and R. Antonucci Physics Department, University of California, Santa Barbara, CA 93106 ABSTRACT It is widely believed that the optical/UV continuum of quasars (the "Big Blue Bump") represents optically thick thermal emission from accretion onto a black hole. Narrow line radio galaxies don't show such a component directly, and were historically thought for that reason to be rotation-powered, with large kinetic luminosity in the radio jets but very little accretion or optical radiation. When the Unified Model came along, identifying at least some narrow line radio galaxies as hidden quasars, the compelling observational motivation for this radio galaxy scenario lost some of its force. However, it is far from clear that all narrow line radio galaxies contain hidden quasar nuclei. The clear sign of a hidden quasar inside a radio galaxy is the appearance of quasar spectral features in its polarized (scattered) light. However that ob- servational test requires suitably placed scattering material to act as a mirror, allowing us to see the nuclear light. A rather robust and more general test for a hidden quasar is to look for the predicted high mid-IR luminosity from the nuclear obscuring matter. The nuclear waste heat is detected and well isolated in the nearest narrow line radio galaxy, Cen A. This confirms other indications that Cen A does contain a modest quasar-like nucleus. However we show here that M87 does not: at high spatial resolution, the mid-IR nucleus is seen to be very weak, and consistent with simple synchrotron emission from the base of the radio jet. This fairly robustly establishes that there are "real" narrow line radio galaxies, without the putative accretion power, and with essentially all the luminosity in kinetic form. Next we show the intriguing mid-IR morphology of Cygnus A, reported pre- viously by us and later discussed in detail by Radomski et al. (2002). All of this mid-IR emission is consistent with reprocessing by a hidden quasar, known to exist from spectropolarimetry by Ogle et al. (1997) and other evidence. Subject headings: galaxies: active -- galaxies: individual (M87, Cen A, 3C 405, 3C 273) -- infrared: galaxies -- 2 -- 1. Introduction Now that the existence of supermassive black holes in AGN seems fairly secure, perhaps the next most fundamental questions are the sources of energy and the nature of the accretion flow in the various classes of objects. Historically (see Begelman, Blandford and Rees 1984 for an early review) it was thought that the optical/UV continuum (or "Big Blue Bump") in quasars (hereafter: and broad line radio galaxies) represents thermal radiation from some sort of cool optically thick accretion flow. Radio galaxies didn't show this component, so were posited to be "nonthermal AGN" with hot radiatively inefficient accretion, at a very low rate; in these cases the the jet power would derive from the hole rotation rather than release of gravitational potential energy from accretion. These arguments took a surprising turn when it was realized that many radio galaxies do have the quasar-like nuclei∗ that are invisible from our line of sight. One Fanaroff- Riley II (edge-brightened, very luminous) radio galaxy, 3C234, was shown in 1982 to have quasar features (broad permitted emission lines and Big Blue Bump) in polarized light (Antonucci 1982, 1984; Antonucci 1993 for a review). Thus 3C234 does have the "thermal" optical/UV emission, which is only visible via scattering. Many other examples have been shown subsequently (e.g. Hines and Wills 1993; Young et al. 1996). Some invocations of the Unified Model postulated that this was generally true of the FR II (powerful, edge- brightened) class (e.g. Barthel 1989). 3C234 has powerful high-ionization narrow lines, consistent with its being a hidden quasar. However it is still contentious how those FR IIs with weak and/or low ionization narrow emission lines fit in (Singal 1993; Laing 1994; Gopal-Krishna et al. 1996; Antonucci 2001). The situation is even less clear for the FR I galaxies, almost all of which have undetectable or low-ionization emission lines. For radio galaxies with no observable high-ionization narrow emission line region present, there is no a priori evidence for the presence of a quasar. These could still have a quasar nucleus, but any narrow line region would need to be mostly obscured as well. In fact there is some evidence that the high ionization narrow emission lines are partially obscured in many 3C radio galaxies (Hes et al. 1993). In NGC 4945 and many other ordinary looking galaxies, the only present evidence for a hidden AGN is in the hard X-ray (e.g. Madejski et al. 2002). To summarize, many FR II radio galaxies fall into the apparently weak, low-ionization emission line category. Most of the FR I radio galaxies do as well. These spectra have been described as "optically dull." Quite recently, Chiaberge et al. (1999, 2000) have shown that among these optically dull sources of both FR types, a majority show optical point sources in -- 3 -- HST images, and of course more might do so with better imaging data. Note that their result seems surprising at first: one might have guessed that those with stronger high-ionization spectra would have the point sources, but the opposite is true. Chiaberge et al. argue that those with detectable optical point sources cannot in general have thick obscuring tori on. The reason is, if the point sources are truly nuclear, then we can see to the very center in most of these objects. These radio galaxies are effectively selected by an isotropic emission property - the radio lobe flux - so the source orientations should have an isotropic distribution. That is, the radio axes should be randomly oriented with respect to our line of sight. Since we can see optical point sources in most optically dull objects, most lines of sight to their nuclei must be unobscured in general. Recall that in the Chiaberge et al. picture, the HST optical point sources are the bases of the synchrotron jets which emit in the radio. As they point out, one caveat must be given with their line of argument. It is possible that obscuring tori exist below the HST resolution, and hide the nucleus and the very innermost region of the conical jet. For example, their optical point sources may really be jet emission on ∼ 1 pc scales, and optical Big Blue Bump sources are much smaller. A torus might be large enough to obscure the latter while still allowing the pc scale jet emission to be seen over the top. In the Chiaberge et al. scenario the optical point sources represent synchrotron emission from the bases of the jets. In support of this idea, those authors show that the optical fluxes correlate roughly with the core radio fluxes at 6 cm. It should be easy to check this: the relationship should tighten up substantially using millimeter fluxes instead of those at 6 cm. It is very important to remember that some narrow line radio galaxies of both FR types do have strong high ionization emission lines, and in some cases definitive spectropolarimetric evidence for a hidden "thermal" nucleus. In general such objects have no detectable optical point source. They behave instead like Seyfert 2 nuclei, with just spatially resolved scattered light and the extended narrow line region directly visible. Our goal is to determine robustly which if any radio galaxies lack a hidden "thermal" (optical/UV continuum) nucleus. This is crucial for AGN theory since it would prove the existence of an accretion mode different from that in quasars. In particular, current wisdom would posit a very low accretion rate and a very low radiative efficiency, that is, some variant of the advection dominated accretion flow for those radio galaxies. Then by default the enormous kinetic luminosity of the radio jets would be attributed to black hole rotational energy. How can we tell whether or not a hidden nucleus is present? One method is via the hard X-rays. Many hidden AGN that are not Compton-thick have been discovered with X-ray -- 4 -- observations. Some references to penetrating X-rays in optically dull sources are gathered in Antonucci 2001; NGC4945 is a spectacular example (e.g. Madejski et al. 2000). Our approach here is to look instead for reradiation of the absorbed light from any hidden quasar-like nucleus by the dusty obscuring matter (torus). Modulo factors of order unity, the various models of the obscuring tori predict that the "waste heat" from the obscuring matter will emerge in the mid-IR, and that this emission is roughly isotropic in all but the highest inclination cases. (Note that this assumption is the weakest point of our project, but that the various torus models predict it is true to within a factor of a few: Pier and Krolik 1992, 1993; Granato et al. 1997; Efstathiou and Rowan-Robinson 1995; a related model in Konigl and Kartje 1994 should also be consulted; Cen A is discussed specifically in Alexander et al 1999). We are making this test for hidden quasars using the Keck I telescope's mid-IR instrument (the "LWS"). This is the only way to isolate the nucleus well, and to achieve the required sensitivity. Here we show the power of the technique, and demonstrate the existence of at least one "nonthermal" AGN†. 2. Observations 2.1. 3C 405 (Cygnus A) in Figure 1 we present a diffraction limited mid-IR image of the nuclear source in 3C 405 (Cyg A), obtained with the Long Wavelength Spectrometer (LWS) instrument at the Keck I telescope‡. All data were taken with the 11.7µm filter, which has an ∼ 1 µm bandpass from 11.2 to 12.2 µm. The image was shown previously in Whysong and Antonucci (2001). The nucleus of 3C 405 was imaged at 11.7µm with Keck I/LWS on 1999 September 30. The chop/nod throw was set to 10" in order to allow imaging of larger scale extended structure; this places the chop beam off the chip, which has a 10" field. We do not report on structures larger than the 10" chop distance. The images were dithered in a 5 position box pattern, 2" to a side, with 53.1 seconds on-source for the positive image per dither position. The entire 5-position exposure was repeated three times, for a total on-source time of 796.6 seconds. Data were processed by subtracting all background chop/nod frames, shifting each dithered image to the correct position, and coadding all dither images. Morphology is ex- tended, with structure to the east and southeast of the nucleus (Fig. 1). The standard star was alpha Ari§, with a FWHM of 0.27". Our 11.7µm image and a partial spectral energy distribution are shown in Figs. 1 and 5. For comparison, the IRAS (large aperture) data for Cygnus A are listed in Table 3. -- 5 -- 2.2. Cen A The high resolution mid-IR data on Cen A also come from the Keck I telescope. Data were obtained on our behalf by Randy Campbell on 28 June 2002. Three filters were used: the 11.7 µm filter with 1 µm bandpass, a wider "SiC" filter centered at 11.7 µm but with a 2.4 µm bandwidth, and the 17.75 µm filter with a ∼ 1 µm bandwidth. The standard star was Sigma Sco, a multiple star which was partially resolved so that it was necessary to increase the synthetic aperture to 2.24 arcsec in order to include all components. No photometric data are available for Sigma Sco at wavelengths longer than M, so we extrapolated to longer wavelengths using the Rayleigh-Jeans approximation. The Cen A images show only an unresolved source in all filters. Photometric calibration results are F(11.7µm)= 1.6 Jy, F(SiC)= 1.8 Jy with ≈ 0.3 arcsec FWHM resolution, and F(17.75 µm)= 2.3 Jy with ≈ 0.5 arcsec FWHM resolution. Since Cen A is so close, there are published data for relatively small physical apertures. A small physical aperture is key, because for both Cen A and M87, the large-aperture (e.g. IRAS, ISO) fluxes are much larger than that from the nucleus. But the much higher resolution Keck data isolate a nuclear point source with size ≈ 5 pc at 11.7µm and ≈ 7.8 pc at 17.75µm given a distance of 3.1 Mpc. We will assume that this is mostly dust emission heated by the nucleus. This is consistent with the 10-20 micron nuclear spectrum taken by ISOCAM CVF in a ∼ 4" aperture (Mirabel et al. 1999, reproduced here as Fig. 3). In fact, our 11.7 µm flux is in good agreement with that of the spectrum, suggesting that Fig. 3 is in fact the nuclear spectrum. For comparison, the published, larger aperture mid-IR fluxes are much higher. The Cen A central region flux has been measured in the mid-IR by Grasdelen and Joyce (1976). Their 3.5" aperture has the same flux as the 5.2" aperture, so there is a compact source surrounded by a region of low or zero flux. The 3.5" aperture corresponds to 50 pc, and the enclosed flux is given as ∼ 2.6 Jy at 11 microns, and 4.3 Jy at 12.6 microns. The IRAS 12.6µm flux measurement is even higher at 13.3 Jy. -- 6 -- 2.3. M87 Here a new high-resolution Keck I image is also crucial. We obtained this data for M87 on 2000 January 18. The observation was made in chop-nod mode using a small 3.5" amplitude so as to keep both the object and chop beams on the CCD chip. The 11.7 µm filter was used, with a 1µm bandpass. Integration time was 96 seconds per dither each for positive and negative (background). Unfortunately, due to a loss of guiding, the positive nucleus image was only fully imaged in one dither frame. However, we are still left with two independent images, both strong pointlike detections. Our measurement was taken from the better placed of the two. Beta And and Mu UMa were used as standards for photometric calibration, yielding a flux scale of 0.0874 and 0.0931 mJy/(ADU/s) and FWHM of 0.33" and 0.31" respectively. This calibration for the unresolved nuclear component in M87 results in a flux of 13 +/- 2 mJy. The uncertainty is dominated by systematic errors in the background subtraction; we adopt a value of 13 mJy. A synthetic aperture of 0.96 arcsec was used, but the source is unresolved so the flux is insensitive to the aperture. Again we note that the large aperture fluxes are much higher. In particular the IRAS fluxes (Moshir et al. 1990) are 231+/−37 mJy at 12µm, <241 mJy at 25µm, and 393+/−51 mJy at 60µm. 3. Discussion 3.1. Cen A Fig 1 shows the SED for Cen A, combining all the ∼ 1" measurements. For AGN which show the quasar-like optical nucleus in scattered light only, the mid-IR emission is typically two orders of magnitude greater than the (scattered) optical light. For Cen A the situation is similar, but with a slight modification (Bailey et al. 1996; Hough et al. 1987; Antonucci and Barvainis 1990; Alexander et al. 1999). The ratio is inflated by absorption of the pc scale scattered optical light by the kpc dust lane famous from photographs. At 2u, a highly polarized point source indicates detection of a ∼ pc scale reflection region. The near-IR reflected light from pc scales penetrates the kpc scale dust lane which has only moderate optical depth. This situation was deduced for 3C323.1 and Cen A by Antonucci and Barvainis (1990). The pc scale dust screen can be consistently identified with the cold absorber seen in the X-ray spectrum, which has column density ∼ 3 × 1023 cm−2. -- 7 -- Additional optical and near-IR scattered light on kpc scales has been mapped and disucssed by Capetti et al 2000 and Marconi et al 2000. The case of Cen A shows that even radio galaxies with weak or low-ionization lines (the "optically dull" ones) can have hidden Type 1 nuclei. Others can be found in e.g. Ekers and Simkin 1983 and Sambruna et al. 2000. To show the level of the reflected light in the nuclear region, we have plotted the value at 2 microns rather than that in the optical (which is highly absorbed). The K-band nuclear flux with starlight subtracted (Marconi et al. 2000) is F(K) ∼ 3 mJy, and νLν ∼ 6 × 1039 ergs/sec for a distance of 3 Mpc. The SED in Figure 2 shows that the mid-IR luminosity of the nucleus is much larger than the optical/near-IR value, as expeted. This is consistent with all other reflected-light objects. The value of the mid-IR flux quoted above is ∼3.5 Jy at 11.7 microns, interpolated from measurements at two nearby wavelengths. The corresponding νLν is 1.2 × 1042 erg/sec. As- suming a normal quasar SED (e.g. Sanders et al 1989; Barvainis et al. 1990), that translates to a bolometric luminosity of 2 × 1043 ergs/sec for the hidden AGN. The nature of the mid-IR emission is important. The SED has been fit to a synchrotron self-Compton model (Chiaberge et al. 2001) which would lead to a classification for Cen A as a "misaligned BL Lac". However, the fit was to the mid-IR continuum slope instead of the different ISO bands because the latter were heavily affected by absorption and emission features (M. Chiaberge 2001, private communication). Small aperture mid-IR spectra show PAH and Si features (Figure 3), and the SED in ≤ 4" apertures is consistent with predom- inantly dust rather than synchrotron (Alexander et al. 1999; note also that the ISOCAM CVF flux from the Mirabel et al. 1999 spectrum is consistent with our 11.7 µm measure- ment, suggesting that the spectrum is representative of our smaller aperture). These features indicate that most of the mid-IR flux is thermal emission as expected for the torus model. 3.2. M87 This radio galaxy is on the FR I-II borderline, both in morphology and in radio power (Owen et al. 2000). It is one of the majority of FR I radio galaxies with an optical/UV point source (Chiaberge et al. 1999, 2000). The optical point source is tentatively ascribed to synchrotron radiation associated with the radio core (see Ford & Tsvetanov 1999, as well as the Chiaberge et al. papers). However it's not known whether this light is highly polarized, or whether broad emission lines are strong in either total or polarized optical/UV flux. -- 8 -- Our small ∼ 0.3 arcsec FWHM beam isolates the innermost ∼ 25pc in M87. The enclosed 11.7µm flux in an 0.6" synthetic aperture is 15 mJy; this aperture matches that used for Cen A in physical size. The flux measured in this way is much lower than the large aperture measurements in the literature. Published large-aperture data leave plenty of room for waste heat from a hidden AGN, but our data do not (Fig. 1). The mid-IR luminosity νLν is only of order that in the optical rather than much greater as for the hidden AGN sources. In fact, much or all of the mid-IR flux could simply be synchrotron radiation associated with the innermost part of the jet, so the measured flux is an upper limit to the dust luminosity. Note that this outcome for M87 is just what Chiaberge et al. implicitly predicted. Unless the nuclear dust is too obscured to emit in the mid-IR, this rules out a powerful hidden nucleus. The observed 11.7µm flux corresponds to νLν = 1.0 × 1041 erg/sec, for a distance of 15 Mpc. Suppose the mid-IR core is in fact all dust emission. For the SED for the PG quasar composite of Sanders et al. 1989, a bolometric luminosity of ∼ 1.6 × 1042 ergs/sec is expected. For comparison, a lower limit to the jet kinetic luminosity in M87 is ∼ 5 × 1044 erg/sec (Owen et al. 2000), so the jet is by far the dominant channel for energy release.k If correct, this suggests that M87 is the true "misaligned BL Lac." Published ADAF models (Reynolds et al. 1996) predict very low IR-optical-UV lumi- nosities compared with those in the radio and X-ray. Our 11.7µm point is about equal to the optical value, but certainly our 11.7µm point may be partially or completely jet emission∗∗. 3.3. Cygnus A This is a very powerful FR II radio galaxy at a redshift of 0.056. It has strong high- ionization narrow lines, suggestive of a hidden AGN. A broad Mg II 2800 emission line is detectable in total flux (Antonucci et al. 1994). That line may or may not be highly polarized, and thus scattered from a hidden nucleus. Several detailed papers report spectroscopic and spectropolarimatric data (Goodrich and Miller 1989; Tadhunter et al. 1994; Shaw and Tadhunter 1994; Vestergaard and Barthel 1993; Stockton et al. 1994; see also Tadhunter et al. 2000 and Thornton et al. 1999 - and there are several others), culminating in Ogle et al. 1997, which shows an extremely broad H-alpha line in polarized flux. It is virtually invisible in total flux because its great width makes it hard to distinguish from continuum emission. A nuclear point source in the near-IR was noted by Djorgovski et al. (1991), but they don't seem to have considered hot dust emission for this excess over the extrapolation from the optical light, as we believe. A powerful hidden nucleus should manifest a mid-IR dust luminosity much larger than -- 9 -- the observed optical luminosity. However for this object and M87 (and virtually all others!) the only IR data available were taken with very large beam sizes. We (and Radomski et al. 2002)†† isolate the core much better with the ∼ 0.3 arcsec (∼ 1.1 kpc) resolution provided by the Keck I telescope, and find a nuclear flux of ∼ 60mJy. An uncertainty here derives from the extended emission, but flux as a function of aperture size does flatten out for apertures larger than the seeing disk, so the 0.96" measurement should be approximately correct (see Table 2). However, we can't be sure from this observation alone that the emission is on pc scales. Since the emission is powerful and at the relatively short wavelength of 11.7µm, it is very likely that this comes from nuclear dust rather than a starburst. It is entirely possible that the extended emission is thermal dust even at radii up to 1 kpc. The temperature of nuclear-heated dust can be estimated according to Barvainis 1988. Adopting a Hubble constant of 75 km/s Mpc, the νLν luminosity at 11.7µm is 9.2 × 1043 erg/sec. If the intrinsic SED is similar to those of PG quasars (Sanders et al. 1989), the 11.7µm value implies a bolometric luminosity of 16.5 νLν (11.7µm) ∼ 1.5 × 1045 erg/sec. An optical luminosity of ≈ 1045 erg/s produces a dust temperature of ≈ 120 K at a 500pc radius. A similar calculation of the dust temperature was done by (Radomski et al. 2002), yielding a slightly higher results (150 K) due to a higher estimate of the optical luminosity. Among the uncertainties are the nuclear UV luminosity and the possiblility of single photon heating. The IRAS (large aperture, see Table 3) dust spectrum is quite cool, suggesting a large starburst contribution. Extended emission is seen in our image at 11.7µm. The core can't be exactly separated from the extensions (see Table 2), but we can estimate around 60 mJy for the nuclear dust. Fig. 1 shows the Cyg A 11.7µm image, and Fig. 5 shows a partial spectral energy distribution. The nuclear luminosity νLν at 11.7µm is 10 times higher than that at 0.5µm. The latter wavelength needs two roughly canceling corrections: subtraction of optical light from the host galaxy, and dereddening (Ogle et al. 1997). The starbust contribution to the nuclear 11.7µm emission is expected to be small, but this should be checked with a spectral slope measurement. The conclusion is simple and expected from prior evidence: Cyg A has a moderately powerful hidden nucleus. As noted above, the estimated bolometric luminosity is 1.5 ×1045 erg/s. For comparison the jet power is estimated several different ways (Carilli and Barthel 1996; Sikora 2001; Punsly 2001). The values are rough, but generally lie in the >∼ 1045 ergs/sec range. This is consistent with the finding that jet power and optical/UV luminosity are often comparable in double radio quasars (e.g., Falcke 1995). -- 10 -- Thus this is a moderate luminosity broad line radio galaxy with a very high luminosity radio source. It has in fact been inferred already that Cygnus A is an over-achiever in the radio (Carilli and Barthel 1996; Barthel and Arnaud 1996). This is well explained qualitatively by the fact that it is the only known FR II radio source in a rich X-ray-emitting cluster. 3.4. Blazars A goal of this paper is to detect thermal IR bumps, often in the presence of a synchrotron In general the SED sampling is limited for our targets. Even in cases with continuum. convincing dust emission, the dust component is not very well isolated in the infrared SED. Thus we thought it would be helpful to show a well observed object which clearly isolates a nonthermal infrared bump from a broadband nonthermal component. Blazar radio sources are typically dominated over most of their SED by variable syn- chrotron emission (and/or inverse Compton bump in the X-ray - γ-ray region). This means that in powerful cases, relatively few continuum components of emission are important for the SED. For example, starlight and dust warmed to tens of K by stars are relatively weak. Thus the infrared consists almost entirely of the broadband nonthermal and the (relatively) narrow band thermal components. The SED, shown in Fig. 4, is that of 3C 273 (data from Robson et al. 1993). It can be referred to while assessing the other SEDs in this paper. 4. Relation to other Radio Galaxies and Conclusion Radio lobe emission is fairly isotropic, so it's easy to make lists of double radio sources that are nearly unbiased with respect to orientation. The visibility of optical point sources in most of the optically dull (weak or low ionization emission line) galaxies show that there are no >∼ parsec scale tori present which are able to obscure the unresolved optical sources in those cases (Chiaberge et al. 1999, 2000). Whatever the nature of the M87 optical point source, we know it's small from the variability on year timescales. The same would apply to the other optically dull nuclei if they vary as M87 does. This would strengthen the Chiaberge et al. argument that we have unobscured sightlines nearly all the way to the central engines. Since the M87 optical/UV flux is quite variable (e.g., Tsvetanov et al. 1998) jet syn- chrotron emission is a possibility. By correlating the radio synchrotron core fluxes and the optical point source fluxes in FR I radio galaxies generally, Chiaberge et al. (1999) infer that the latter are in fact likely to be beamed synchrotron sources. Crucial tests of the nature of -- 11 -- the optical point sources can be made with spectroscopy and polarimetry. We hope to do this with adaptive optics, excluding most of the starlight that dominates in arcsec apertures. However, a large minority of low ionization FR I and FR II radio galaxies show no point source, and are similar in this way to AGN with hidden nuclei. In fact the closest FR I, Cen A, does have a big molecular torus (see Fig. 2 of Rydbeck et al. 1993!), and substantial evidence for a hidden nucleus as well (see references cited earlier). As a working hypothesis we might suppose that the same is true for all those without detectable optical pointlike nuclei. (Of course sensitivity of the optical/UV observations also enters in.) Thus the FR I family is heterogeneous: some contain hidden optical/UV nuclei and some do not. It's been difficult to find FR I objects with strong evidence for hidden AGN, but in fact at least a few are known to be quasar-like from direct spectroscopy (3C120 is well known; see also Lara et al. 1999 and Sarazin et al. 1999). The infrared peaks in some blazars such as 3C 66A are clearly distinct from the radio synchrotron, providing further evidence for FR I objects with hidden thermal emission. Therefore the "nonthermal" model does not apply to all optically dull or low ionization radio galaxies, or to all FR I galaxies. The FR class has no apparent direct relation to the mode of energy production, consistent with much recent evidence that FR Is behave very much like FR IIs at VLBI scales. We will understand this better after the completion of our mid-IR program. Finally we reiterate an assumption in our program: we assume that any mid-IR dust emission is isotropic to within a factor of a few. -- 12 -- 5. Acknowledgements We wish to thank M. Chiaberge, P. Ogle, B. Wills, and R. Barvainis for good advice. A thousand huzzas to Randy Campbell at Keck for obtaining the Cen A data. Alexander, D. M., Efstathiou, A., Hough, J. H., Aitken, D. K., Lutz, D., Roche, P. F., & REFERENCES Sturm, E. 1999, MNRAS, 310, 78 Antonucci, R 1982 Nature 299, 605 Antonucci, R 1993 Ann Rev Astron Astrophys 31, 473 Antonucci, R 1984 ApJ 278, 499 Antonucci, R 2001 preprint (astro-ph/0103048) Antonucci, R, Hurt, T, and Kinney, A 1994 Nature 371, 313 ∗We will use the terms "thermal AGN" and "hidden AGN" to refer to the optical/UV continuum and broad emission lines which characterize the spectra of quiescent quasars, broad line radio galaxies, and Seyfert galaxies in this spectral region. †Preliminary reports of this work were published in Anotnucci 2001 and Whysong and Antonucci 2001 (astro-ph 0106381); shortly thereafter, Gemini mid-IR images were published and analyzed by Perlman et al. 2001. Their measurements and conclusions were similar to ours. Since their images were very deep, they were also able to detect extended jet emission. ‡Instrument reference is available at: http://www2.keck.hawaii.edu:3636/realpublic/inst/lws/lws.html. §A table of photometric standards is available at: http://www2.keck.hawaii.edu:3636/realpublic/inst/lws/IRTF Standards.html. kWe can also make the empirical test of the ratio of the mid-IR flux to that of the radio lobes. We measured 15mJy at 11.7µm, and NED shows ∼ 220Jy at 1.4GHz. The ratio is 6.8 × 10−5, much lower than the ratio for the objects with hidden AGN, such as Cen A. ∗∗It is unclear to us why the radio points, which fit the ADAF model, are taken as measurements while the (non-fitting) optical fluxes were not. Also, the Reynolds et al. figure apparently uses 3C273 as a "thermal" quasar-like template, but that object definitely has a large jet contribution in the radio and infrared (Robson et al. 1993). ††Our Keck I image appeared publicly before that of Radomski et al. 2002, but we did relatively little analysis. -- 13 -- Barth, A J, Hien, H D, Brotherton, M S, Filippenko, A V, Ho, L C, van Breugel, W, Antonucci, R, and Goodrich, R 1999 A J 118, 1609 Barthel, P 1989 ApJ 336, 606 Barthel, P D, and Arnaud, K A 1996 MNRAS 283, L45 Barvainis, R E 1987 ApJ 320, 537 Barvainis, R E 1990 ApJ 353, 419 Begelman, M, Blandford, R, and Rees, M 1984 Rev Mod Phys 56, 255 Blandford, R and Znajek, R 1977 MNRAS 179, 433 Carilli, C L, and Barthel, P D 1996 A&AR 7, 1 Capetti, A et al. 2000 ApJ 544, 269 Chiaberge, M, Capetti, A, and Celotti, A 1999 Astron Astrophys 349, 77 Chiaberge, M, Capetti, A, and Celotti, A 2000 Astron Astrophys 355, 873 Chiaberge, M, Capetti, A, and Celotti, A 2001, MNRAS, 324, 33 Clarke, D, Burns, J, and Norman, M 1992 ApJ, 395, 444 Cohen, M H, Ogle, P M, Tran, H D, Goodrich, R W, and Miller, J S, 1999 AJ 188, 1963. Djorgovski, S, Weir, N, Matthews, K, and Graham, J R 1991 ApJL 372, 67 Miller, J S 1999 Astron J 118, 1963 Efstathiou, A. & Rowan-Robinson, M. 1995, MNRAS, 273, 649 Ekers, R D, and Simkin, S M 1983 ApJ 265, 85 Falcke, H, Malkan, M A, Biermann, P L 1995 A&A 298, 375 Ford, H, and Tsvetanov, Z 1999, in "The Radio Galaxy M87: proceedings of a workshop held at Ringberg Castle" (NY:Springer), Roser and Meisenheimer, eds., p.∼278 Goodrich, R W, and Miller, J S 1989 ApJ 346, L21. Gopal-Krishna, Kulkarni, V K, and Wiita, P J 1996 ApJ 463, L1 Granato, G. L., Danese, L., & Franceschini, A. 1997, ApJ, 486, 147 -- 14 -- Grasdalen, G L, and Joyce, R R 1976 ApJ 208, 317 Hes, R., Barthel, P. D., and Fosbury, R. A. E., 1993 Nature 362, 326 Hines, D C, and Wills, B J 1993 ApJ 415, 82 Impey, C D and Neugebauer, G 1988, Astronom. J, 95, 307 Konigl, A. & Kartje, J. F. 1994, ApJ, 434, 446 Kishimoto, M, Antonucci, R, Cimatti, A, Hurt, T, Dey, A, and van Breugel, W 2001, ApJ, in press; also astro-ph 0010001 Laing, R 1994 Physics of Active Galaxies, ASP Conf Series #54, eds. Bicknell et al. Lara, L, Marquez, I, Cotton, W D, Feretti, L, Giovannini, G, Marcade, J M, and Venturi, ,T 1999 New Astr Reviews 43, 643 Marconi, A, Schreier, E J, Koekemoer, A, Capetti, A, Axon, D, Macchetto, D, and Caon, N 2000 ApJ 528, 276 Moshir, M. et al. Infrared Astronomical Satellite Catalogs, 1990, The Faint Source Catalog, version 2.0 Ogle, P M, Cohen, M H, Miller, J S, Tran, H D, Fosbury, R A E, and Goodrich, R W 1997 ApJ 482, L37 Owen, F W, Eilek, J A, and Kassim, N E 2000 ApJ 543, 611 Perlman, E S, Sparks, W B, Radomski, J, Packham, C, Biretta, J, and Fisher, R S, 2001 ApJ 561, 51L Punsly, B 2001, Black Hole Gravitohydromagnetics (Berlin: Springer) Radomski, J T, Pina, R K, Packham, C, Telesco, C M, Tadhunter, C 2002 ApJ 566, 675. Reynolds, C S et al. 1996 MNRAS 283, L111 Robson, E I et al. 1993 MNRAS 262, 249 Robson, E I, Leeuw, L L, Stevens, J A, Holland, W S 1998 MNRAS 301, 935 Rydbeck, G, Wiklind, T, Wild, W, Eckart, A, Genzel, R, and Rothermel, H 1993 A&A 270, L13 -- 15 -- Sambruna, R M, Chartas, G, Eracleus, M, Mushotzky, R F, and Nousek, J A 2000 ApJ 532, L91 Sanders, D B, Phinney, E S, Neugebauer, G, Soifer, B T, and Matthews, K 1989 ApJ 347, 29 Sarazin, C L, Koekemoer, A M, Baum, S A, O'dea, C P, Owen, F N, and Wise, M W 1999 ApJ 510, 90 Singal, A 1993 MNRAS 262, L27 Shaw,, M and Tadhunter, C 1994 MNRAS 267, 589 Sikora, M 2001 ASP Conference Series: Blazar Demographics and Physics (in press); also astro-ph 01011381 Stockton, A Ridgway, S E, and Lilly, S J 1994 AJ 108, 414 Thornton, R J, Stockton, A, and Ridgway, S 1999 AJ 118, 146 Tadhunter, C N, Metz, S, and Robinson, A 1994 MNRAS 268, 989 Tadhunter, C, et al. 2000 MNRAS 313, L52 Tsvetanov, Z et al. 1998 ApJ 493, L83 Vestergaard, M, and Barthel, P D Ap&SS 205, 135 Young, S, Hough, J H, Efstathiou, A, Wills, B J, Axon, D J, Bailey, J A, and Ward, M J 1996 MNRAS 281, 1206 Whysong, D and Antonucci, R, preprint (astro-ph/0106381) Wilkes, B J, Schmidt, G D, Smith, P S, Mathur, S, and McLoed, K K 1995 ApJ 455, L13 This preprint was prepared with the AAS LATEX macros v5.0. -- 16 -- Fig. 1. -- Keck I/LWS 11.7µm image of Cygnus A. The arrow indicates north and is 1 arcsec long. -- 17 -- Fig. 2. -- Spectral energy distribution for Cen A (triangles, rectangles, and thin lines, adopted from Chiaberge et al. 2001), and M87 (open circles), with ADAF models for three different accretion rates (Reynolds et al. 1996). Additional M87 points are from the IRAS faint source catalog (Moshir et al. 1990) (solid circles) and Keck I/LWS (cross). The Cen A points have been shifted up two decades for clarity. -- 18 -- Fig. 3. -- ISOCAM CVF spectrum of Centaurus A (from Mirabel et al. 1999). -- 19 -- Fig. 4. -- Spectral energy distribution for 3C 273. Data were obtained from Robson et al. 1993. -- 20 -- Fig. 5. -- Partial core SED for Cygnus A. Data are from IRAS (squares, Impey and Neuge- bauer 1988), Keck/LWS (triangle), and Palomar 200 inch (circles, Djorgovski et al. 1991). Errors for the Keck/LWS point represent uncertainty in nuclear emission due to the extended structure. Other data are plotted with 10% error bars. While measurement errors were much smaller than this, we include these larger uncertainties due to the different apertures. -- 21 -- Table 1: Flux Ratios Mid-IR results are from our Keck I survey. 1.4 GHz (mJy) 3800 4000 8000 5400 5100 10600 11.7µm (mJy) 10 9.7 7.0 186 27.5 22 Name 3C 47 3C 216 3C 219 3C 234 3C 382 3C 452 Ratio 2.6E-3 2.4E-3 8.8E-4 3.4E-2 5.4E-3 2.1E-3 Table 2: LWS photometry results for Cygnus A: aperture diameter (arcsec) 0.64 0.96 1.28 1.60 1.92 2.56 3.20 flux (mJy) 44 71 93 111 122 139 152 Table 3: IRAS photometry for Cygnus A (Impey and Neugebauer 1988): 12 µm 25 µm 60 µm S = 144 +/- 5 mJy S = 870 +/- 5 mJy S = 2908 +/- 13 mJy
astro-ph/0309096
1
0309
2003-09-03T11:25:10
The Ionized Gas and Nuclear Environment in NGC 3783. IV. Variability and Modeling of the 900 ks CHANDRA Spectrum
[ "astro-ph" ]
We present a detailed spectral analysis of the data obtained from NGC 3783 during the period 2000-2001 using Chandra. This analysis leads us to the following results. 1) NGC 3783 fluctuated in luminosity by a factor ~1.5 during individual observations (~170 ks duration). These fluctuations were not associated with significant spectral variations. 2) On a longer time scale (20-120 days), we found the source to exhibit two very different spectral shapes. The main difference between these can be well-described by the appearance and disappearance of a spectral component that dominates the underlying continuum at the longest wavelengths. The spectral variations are not related to the brightening or the fading of the continuum at short wavelengths in any simple way. 3) The appearance of the soft continuum component is consistent with being the only spectral variation, and there is no need to invoke changes in the opacity of the absorbers. 4) Photoionization modeling indicates that a combination of three ionized absorbers, each split into two kinematic components, can explain the strengths of almost all the absorption lines and bound-free edges. All three components are thermally stable and seem to have the same gas pressure. 5) The only real discrepancy between our model and the observations concerns the range of wavelengths absorbed by the iron M-shell UTA feature. This most likely arises as the result of our underestimation of the poorly-known dielectronic recombination rates appropriate for these ions. 6) The lower limit on the distance of the absorbing gas in NGC 3783 is between 0.2 and 3.2 pc. The assumption of pressure equilibrium imposes an upper limit of about 25 pc on the distance of the least-ionized component from the central source. (abridged)
astro-ph
astro-ph
THE ASTROPHYSICAL JOURNAL, ASTRO-PH/0309096 Preprint typeset using LATEX style emulateapj v. 7/15/03 THE IONIZED GAS AND NUCLEAR ENVIRONMENT IN NGC 3783. IV. VARIABILITY AND MODELING OF THE 900 KS CHANDRA SPECTRUM HAGAI NETZER,1 SHAI KASPI,1 EHUD BEHAR,2 W. N. BRANDT,3 DORON CHELOUCHE,1 IAN M. GEORGE,4,5 D. MICHAEL CRENSHAW,6 JACK R. GABEL,7 FREDERICK W. HAMANN,8 STEVEN B. KRAEMER,7 GERARD A. KRISS,9,10 KIRPAL NANDRA,11 BRADLEY M. PETERSON,12 JOSEPH C. SHIELDS,13 AND T. J. TURNER,4,5 Received 2003 July 7; accepted 2003 September 1 ABSTRACT We present a detailed spectral analysis of the data obtained from NGC 3783 during the period 2000 -- 2001 using Chandra. The data were split in various ways to look for time- and luminosity-dependent spectral variations. This analysis, along with the measured equivalent widths of a large number of X-ray lines and photoionization calculations, lead us to the following results and conclusions. 1) NGC 3783 fluctuated in luminosity by a factor ∼ 1.5 during individual observations (most of which were of 170 ks duration). These fluctuations were not associated with significant spectral variations. 2) On a longer time scale (20 -- 120 days), we found the source to exhibit two very different spectral shapes. The main difference between these can be well-described by the appearance (in the "high state") and disappearance (in the "low state") of a spectral component that dominates the underlying continuum at the longest wavelengths. Contrary to the case in other objects, the spectral variations are not related to the brightening or the fading of the continuum at short wavelengths in any simple way. NGC 3783 seems to be the first AGN to show this unusual behavior. 3) The appearance of the soft continuum component is consistent with being the only spectral variation, and there is no need to invoke changes in the opacity of the absorbers lying along the line of sight. Indeed, we find all the absorption lines which can be reliably measured have the same equivalent widths (within the observational uncertainties) during high- and low-states. 4) Photoionization modeling indicates that a combination of three ionized absorbers, each split into two kinematic components, can explain the strengths of almost all the absorption lines and bound- free edges. These three components span a large range of ionization, and have total column of about 4×1022 cm−2. Moreover, all three components are thermally stable and seem to have the same gas pressure. Thus all three may co-exist in the same volume of space. This is the first detection of such a multi-component, equilibrium gas in an AGN. 5) The only real discrepancy between our model and the observations concerns the range of wavelengths absorbed by the iron M-shell UTA feature. This most likely arises as the result of our underestimation of the poorly-known dielectronic recombination rates appropriate for these ions. We also note a small discrepancy in the calculated column density of O VI and discuss its possible origin. 6) The lower limit on the distance of the absorbing gas in NGC 3783 is between 0.2 and 3.2 pc, depending on the component of ionized gas considered. The assumption of pressure equilibrium imposes an upper limit of about 25 pc on the distance of the least-ionized component from the central source. Subject headings: galaxies: active -- galaxies: individual (NGC 3783) -- galaxies: nuclei -- galaxies: Seyfert -- techniques: spectroscopic -- X-rays: galaxies 1 School of Physics and Astronomy, Raymond and Beverly Sackler Fac- ulty of Exact Sciences, Tel-Aviv University, Tel-Aviv 69978, Israel. 2 Physics Department, Technion, Haifa 32000, Israel 3 Department of Astronomy and Astrophysics, 525 Davey Laboratory, The Pennsylvania State University, University Park, PA 16802. 4 Laboratory for High Energy Astrophysics, NASA/Goddard Space Flight Center, Code 662, Greenbelt, MD 20771. 5 Joint Center for Astrophysics, Physics Department, University of Mary- land, Baltimore County, 1000 Hilltop Circle, Baltimore, MD 21250. 6 Department of Physics and Astronomy, Georgia State University, At- MD 21218 11 Astrophysics Group, Imperial College London, iBlackett Laboratory, Prince Consort Rd., London SW7 2AZ, UK 12 Department of Astronomy, Ohio State University, 140 West 18th Av- enue, Columbus, OH 43210-1106. 13 Department of Physics and Astronomy, Clippinger Research Labs 251B, Ohio University, Athens, OH 45701-2979. lanta, GA 30303. MD 20771. 7 Catholic University of America, NASA/GSFC, Code 681, Greenbelt, 8 Department of Astronomy, University of Florida, 211 Bryant Space Sci- ence Center, Gainesville, FL 32611-2055. 9 Center for Astrophysical Sciences, Department of Physics and Astron- omy, The Johns Hopkins University, Baltimore, MD 21218-2686. 10 Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, 1. INTRODUCTION The barred-spiral galaxy NGC 3783 (V ≃ 13.5 mag., z = 0.0097) hosts a well-studied, type-I active galactic nucleus (AGN) with prominent broad emission lines and strong X- ray absorption features. The object has been observed ex- tensively with almost all X-ray instruments, most recently by Chandra (Kaspi et al. 2002, hereafter Paper I) and XMM- Newton (Blustin et al 2002). The 2 -- 10 keV flux of NGC 3783 varies in the range ∼ (4 -- 9) × 10−11 ergs cm−2 s−1, and its mean 2 -- 10 keV luminosity is ∼ 1.5 × 1043 erg s−1 (for H0 = 70 km s−1 Mpc−1 and q0 = 0.5). Paper I gives an extensive list of references and a comprehensive summary of historical observations, including ground-based and UV (HST) observa- tions. It also discusses the unique Chandra data set obtained in 2000 -- 2001. These observations consist of a relatively short observation performed in 2000 January, and five longer ob- servations performed in 2001 February -- June, separated by various intervals from 2 to 120 days (see §2). Paper I con- tains numerous illustrations of the mean spectrum, absorption line profiles and detailed measurements of many absorption and emission lines. Two other papers (Gabel et al. 2003a,b) 2 NETZER ET AL. discuss the HST and FUSE data obtained as part of this multi- waveband campaign. There have been several previous attempts to model the characteristics of the X-ray absorbing gas along the line of sight to NGC 3738. Here we comment only on the more de- tailed works. A very detailed work Kaspi et al (2001) at- tempted to fit the spectrum obtained with Chandra in 2000. The Kaspi et al. most successful model consists of two ab- sorbing shells, both outflowing from the central source with a velocity of ∼ 600 km s−1. The gas in both shells was as- sumed to be turbulent with internal turbulent motion of ∼ 250 km s−1. The two shells had similar column densities, but dif- ferent ionization parameters. The model is consistent with the intensity and equivalent width (EW) of many (but not all) emission and absorption lines observed in the spectrum. A major limitation of the Kaspi et al. (2001) work was the lim- ited signal-to-noise (S/N) of the spectrum, which results in large uncertainties on the model parameters. In addition, the limited duration of the only observation then available did not allow a meaningful analysis of any time-dependence. A new paper, by Krongold et al. (2003, hereafter K03), discussing the full Chandra data set, was accepted for publication after the submission of our paper. Some results of that work are relevant to our study and are discussed in the various sections below. This paper discusses the complete spectroscopic Chandra data of NGC 3783. We present our modeling of this material, with emphasis on the properties of the absorber(s) along the line of sight. In §2 we describe the two X-ray spectral states discovered using the new observations. In §3 we explain the various ways we measured and modeled the spectrum. §4 con- tains a discussion of the new findings, and §5 summarizes our new results. 2. THE TWO-STATE X-RAY SPECTRUM OF NGC 3783 A full description of the Chandra data considered here is given in Paper I. In brief, there were 5 observations (each of ∼170 ks duration) obtained over the period 2001 February -- June, and a shorter observation (of ∼56 ks duration) obtained in 2000 January (and also discussed in Kaspi et al. 2001). All six observations were performed with the High-Energy Trans- mission Grating Spectrometer (HETGS) in place, and consist of a total exposure of 888.7 ks. All measurements in Paper I refer to a time-averaged spectrum, produced by combining the first-order spectra from both grating arms, using 0.01 A wide bins. Third-order data from the medium-energy grat- ing (MEG) were also used to compare the profiles of several short-wavelength lines with the first-order profiles obtained for several lines at longer wavelength. It was found that most of the resolved absorption lines (e.g. O VII, Ne IX and Si XIV lines) consist of at least two kinematic components, outflow- ing at −500 ± 100 km s−1 and −1000 ± 200 km s−1 (respec- tively). The overall absorption profiles in NGC 3783 covers the velocity range of 0 to -1600 km s−1 (e.g. see Figs. 5 & 10 in Paper I). A major goal of the present work is to investigate subsets of the Chandra data. We subdivided the data in various ways to search for spectral variations as a function of source lumi- nosity and/or time. This was done for both short (i.e. within the individual 170 ks observations) as well as for long (be- tween observations i.e. 20 -- 90 days) time scales. The short time-scale behavior of NGC 3783 is illustrated in Fig.1 which shows the short wavelength (2 -- 10 A) flux as a function of time for all observations binned in intervals of 3170 sec in the left panels. The right panels shows the "softness" ratio, defined as the flux in the 15 -- 25 A band divided by that in the 2 -- 10 A band, to illustrate now the different variability characteristics exhibited by the source above and below ∼1 keV. From Fig. 1 it can be seen that the 2 -- 10 A flux varied by about 50% over the 170 ks duration of each of the ob- servations. However, inspection of the corresponding soft- ness light curves shows that the spectral variations are much smaller and almost insignificant. Thus, the data show very little cahnges of the spectral energy distribution (SED) over time scales .4 days ("obsids" 2090 and 2091 are consecutive with total duration of about 4 days). To further test this find- ing, we have also extracted various spectra when the source was in different intensity states, but did not find any signifi- cant spectral variations during this period. We also note that Behar et al. (2003) did not find any spectral variations dur- ing a three-day observation of NGC 3783 in 2001 December using XMM-Newton. Contrary to the above, there is a very significant change of SED on longer time scales. In particular, there is a major change in the softness ratio between obsids 2092 and 2093 (separated by 20 days) and obsids 2093 and 2094 (separated by 90 days). As shown in Fig. 1, obsid 2093 shows a higher mean count rate but exhibits a much softer spectrum. Direct comparison of the time-averaged spectra obtained during ob- sid 2093 with that obtained during obsid 2094 reveals the low energy part (l > 15 A) decreased in flux by a factor of ∼ 4, but that the shorter wavelength continuum exhibited a much smaller decrease (factor ∼ 1.5) The opposite change occured between obsids 2092 and 2093. Most importantly, the soft- ness ratio variations are not simple luminosity-related effects. For example, the 2 -- 10 A count rate in the middle of obsid 2093 (Fig. 1) is a little lower than the 2 -- 10 A flux toward the end of obsid 2094, yet the softness ratios are significantly different. A similar effect is seen in obsid 2093, during which the 2 -- 10 A count rate varied by about 50% yet the softness ratio remaining approximately constant. In light of these results, we have divided the entire data set into groups with high and low softness ratios. We find four observations (2090, 2091, 2092 and 2094) with low softness ratio and two (0373 and 2093) with high softness ratio. Here- after we refer to these as the "low-" and "high-state" observa- tions, respectively. We find no significant differences between the mean spectra of the four individual low-state observations (except for a small intensity variations). Similarly we find no significant differences between the mean spectral shapes of the two high-state observations. Fig. 2 is a more detailed example of this phenomenon. It shows the softness ratio as a function of the 2 -- 10 A flux for all the data using 3170s bins. A separation into two groups is apparent, and a standard KS statistical test confirms its sig- nificance. We have also investigated possible linear correla- tions of the softness ratio with the 2 -- 10 A flux within each of the states. For the low-state observations, we find a sig- nificant (> 99%) linear correlation between softness ratio and flux. However, as can be seen from Fig.2, clearly this line does not connect the low- and high-state data. No significant correlation of softness ratio with flux was found for the high- state data. We conclude that the X-ray spectrum of NGC 3783 fluctuates between two states of different softness ratio. The combined spectra of the two are shown in Fig. 3 and much of the rest of this paper is devoted to the analysis of this unusual behavior of NGC 3783. VARIABILITY AND MODELING OF NGC 3783 3 obsid=2094 2001 Jun 26 2087 2087.5 2088 2088.5 0.24 0.2 0.16 0.12 0.08 0.04 2001 Jun 26 2087 2087.5 2088 2088.5 obsid=2093 2001 Mar 31 2000 2000.5 2001 2001.5 obsid=2092 2001 Mar 10 1979 1979.5 1980 1980.5 obsid=2091 2001 Feb 27 2001 Mar 31 2000 2000.5 2001 2001.5 2001 Mar 10 1979 1979.5 1980 1980.5 2001 Feb 27 8 7 6 5 4 3 8 7 6 5 4 3 1999.5 8 7 6 5 4 3 1978.5 8 7 6 5 4 3 ] 1 - Å 1 - s 2 - m c s g r e 1 1 - 0 1 [ r e d r o t s r i f G E M Å 0 1 - 2 x u l F 0.24 0.2 0.16 0.12 0.08 0.04 1999.5 0.24 0.2 0.16 0.12 0.08 0.04 1978.5 r e d r o t s r i f G E M ) x u l f Å 0 1 - 2 ( / ) x u l f Å 5 2 - 5 1 ( [ o i t a r s s e n t f o S 0.24 0.2 0.16 0.12 0.08 0.04 0.24 0.2 0.16 0.12 0.08 0.04 1968 1968.5 1969 1969.5 obsid=2090 2001 Feb 24 1965.5 1966 1966.5 1967 obsid=0373 2000 Jan 21 8 7 6 5 4 3 8 7 6 5 4 3 1564.5 1565 1565.5 1566 1566.5 JD - 2450000 1968 1968.5 1969 1969.5 2001 Feb 24 1965.5 1966 1966.5 1967 0.24 0.2 0.16 0.12 0.08 0.04 1564.5 2000 Jan 21 1565 1565.5 1566 1566.5 JD - 2450000 FIG. 1. -- MEG 2 -- 10 A flux light curves (left) and softness ratios (ratio of the 2 -- 10 and the 15 -- 25 A count rates, right) for all Chandra observations of NGC 3783. The dashed lines across observations no. 2093 and 2094 represent constant flux (4.8) and softness ratio (0.15). They illustrate the fact that similar 2 -- 10 A fluxes (middle part of obs. 2093 and last part of obs. 2094) can be associated with very different softness ratios. 4 ] ) x u l f Å 0 1 - 2 ( / ) x u l f Å 5 2 - 5 1 ( [ o i t a r s s e n t f o S 0.24 0.21 0.18 0.15 0.12 0.09 0.06 Obsn. 2090 Obsn. 2091 Obsn. 2092 Obsn. 2094 Obsn. 0373 Obsn. 2093 3 3.5 4.5 4 2-10 Å Flux [10-11 ergs cm-2 s-1 Å-1] 5.5 5 6 NETZER ET AL. TABLE 1. EQUIVALENT WIDTHSa Ion & Line low state high state Kaspi et al. (2002) Si XIV l 6.182 Si XIII l 6.648 Si XII l 6.718 b Si XI l 6.778 Si X l 6.859 Si IX l 6.931 Si VIII l 6.999 S XVI l 4.729 S XV l 5.039 20.4 ± 1.2 16.0 ± 1.3 2.9 ± 0.5 4.8 ± 0.8 10.6 ± 1.1 7.2 ± 1.0 4.3 ± 1.0 11.5 ± 1.6 8.7 ± 1.7 22.6 ± 1.7 14.6 ± 1.3 3.8 ± 0.7 5.9 ± 1.2 9.4 ± 1.4 7.7 ± 1.6 4.4 ± 1.4 11.6 ± 2.5 9.1 ± 2.5 20.5 ± 0.8 14.9 ± 0.7 3.0 ± 0.4 4.8 ± 0.6 10.9 ± 0.8 7.1 ± 0.7 4.3 ± 0.8 10.7 ± 1.2 9.2 ± 1.2 aEquivalent widths were measured as described in Paper I and are given in m A. bThis line is blended with Mg XII l 6.738 with 1:1 ratio. Tabulated values are only for the Si XII l 6.718 line. 6.5 7 7.5 A / 2 m c / c e s / n o o h p 4 − 0 1 t FIG. 2. -- Softness ratio vs. 2 -- 10 A flux for the data in Fig. 1. The low and high-state observations are shown as filled and open symbols, respec- tively, and the various observations are denoted with different colors. The solid straight line is the linear regression fit to the low-state observations. No significant correlation of softness ratio vs. flux was found for the high state observations. 25 20 15 10 5 0 5 10 15 rest wavelength (A) 20 FIG. 3. -- High-state (red, observations 0373 and 2093 combined) and low- state (blue, all other observations combined) spectra of NGC 3783 binned to 0.04 A. The solid lines are the two chosen intrinsic continua (see §3.2.5). The lower continuum is a single power-law with G = 1.6. The high continuum is made out of two power-laws with slopes G = 1.6 and G = 3.2 and relative normalization of 1:0.4 at 1 keV. These power-laws are discussed in §3.2.5. 3. SPECTRAL ANALYSIS AND MODELING 3.1. Spectral differences between the high and the low states We have investigated the unusual variations observed in the spectrum of NGC 3783 in an attempt to understand their na- ture and their origin. In particular, we have attempted to an- swer the question of whether they are due to the response of the absorbing gas to the variations observed in the continuum. There are two ways to answer the above question -- by a direct and detailed spectral comparison, and by modeling the two spectra trying to establish the origin of the differences. Fig. 4 is an example of the first approach. The diagrams shows a comparison of the low- and high-state spectra after applying a simple scaling factor to the former such that the (local) continua have the same intensity. As the scaling factor is wavelength dependent, we divided the two spectra into sev- eral bands and applied different factors in each case. The two examples shown in Fig. 4 (and all others we have examined) illustrate that, except for a luminosity scaling, the two absorp- tion line spectra are indistinguishable within the uncertainties (∼10% for the low-state spectrum and about twice that for the high-state spectrum). The somewhat weaker looking lines of low ionization species in the high-state spectrum (e.g. around 7 A) are well within the noise. Table 1 lists the EWs measured for several key absorption lines (see §3.2.3) in both the high- and low-state spectra. Again, the small differences between the two spectra are well within the measurement errors. We proceed under the assumption that there are no variations in the line EWs between the two states. We comment on the implications of this in §4.1. 3.2. Spectral modeling The second and complementary approach for investigat- ing the spectral changes is by detailed modeling of the ab- sorbing and emitting gas. The underlying idea is that the gas is photoionized by the central X-ray source, and that the observed spectra represent its physical state during the two states. The principles and the ingredients of such modeling were outlined in Netzer (1996) and previous applications to the case of NGC 3783 were discussed by Kaspi et al. (2001). A very recent analysis is provided by K03 who describe a detailed model composed of two absorbers that are different from those suggested by Kaspi et al. (2001). Below we com- ment on the similarity and differences between our work and the new K03 paper. Here we summarize our basic method and explain its application to the high and the low-state spectra of NGC 3783. 3.2.1. General method and model ingredients The X-ray gas is assumed to be photoionized by a central source, and in photoionization and thermal equilibrium (see §4 for discussion of the last point). Modeling was performed using ION2003, the 2003 version of the code ION (Netzer 1996). The code includes all relevant atomic processes, and computes the ionization and thermal structure of the gas along with the intensities and EWs for more than 2000 X-ray lines. The code is able to consider various geometries, from a sin- gle cloud to a multi-component, expanding atmosphere. The VARIABILITY AND MODELING OF NGC 3783 5 A / 2 m c / c e s / / n o t o h p d e z i l a m r o n 12 10 8 6 4 2 0 high vs. 1.5xlow 6.2 6.4 6.6 6.8 7 7.2 rest wavelength [A] A / 2 m c / c e s / n o t o h p d e z i l a m r o n 6 4 2 0 high vs. 3xlow 16 16.5 17 rest wavelength [A] 17.5 FIG. 4. -- Left: High-state (red) vs. low-state (blue) 6 -- 7 A spectra of NGC 3783 showing the great similarity in absorption line EWs (cf. table 1). The low-state spectrum was multiplied by 1.5 to match the flux of the high-state continuum. For clarity, error bars are only plotted for the low state spectrum. Right: Same for the 15.6 -- 18 A range except for the different scaling of the low-state continuum. TABLE 2. ASSUMED COMPOSITION Element Relative abundance H He C N O Ne Mg Al Si S Ar Ca Fe 1.0 0.1 3.7 × 10−4 1.1 × 10−4 5.0 × 10−4 1.0 × 10−4 3.7 × 10−5 3.0 × 10−6 3.5 × 10−5 1.6 × 10−5 3.3 × 10−6 2.3 × 10−6 4.0 × 10−5 basic parameters of the model are the gas density (assumed to be in the range 102−6 cm−3 - see justification in §4), the hydrogen column density (N, in units of cm−2), the covering factor, the gas composition, turbulent motion (§3.2.2) and the oxygen ionization parameter (UOX ) defined over the energy range 0.54 -- 10 keV. As explained by Netzer (1996), and dis- cussed further in George et al. (1998), this choice of ioniza- tion parameter (compared with, for example, UX defined over the 0.1 -- 10 keV range) gives the most meaningful definition of the ionization field of X-ray photons for AGN. Regarding the covering factor, we distinguish between the emission covering factor (W /4p ) applicable to the emission line gas, and the ab- sorption (line of sight) covering factor which can be different. The "solar composition" used throughout this work is given in Table 2 (note the reduced oxygen abundance compared with older estimates). The incident continuum is taken to be the broken power- law defined in Kaspi et al. (2001, Table 4). The only changes we have experimented with are related to the slope of the 0.1 -- 50 keV continuum. As discussed below, the UV part of the spectrum can be different from the one assumed here with important implications for the UV absorption lines. This will be investigated in a forthcoming paper. The effects of the UV continuum on the strongest features seen in the X-ray band are of far less importance. Thus we consider this SED to be adequate for the present analysis. The models calculated here are entirely self-consistent and are not simple attempts to fit the observed spectra by measur- ing line EWs and deducing column densities for the differ- ent ions. We have searched for the combination of physical components that can be produced in nature in an environment where low density gas is exposed to a typical AGN contin- uum. These components were then combined in a realistic manner, taking into account screening, attenuation of the ra- diation field, etc. 3.2.2. Multi-component models We have examined the hypothesis that the spectral changes observed arise purely as a result of variations in the opacity of the absorbing gas which are caused solely by changes in the intensity of the ionizing continuum. Thus, for this experi- ment, we assume the shape of the SEDs is the same for both the low- and high-states and only differ in total luminosity. We start by calculating a variety of models in order to mimic the low-state spectrum. Each model is made of several emis- sion and absorption components. The absorbers are specified by UOX , N and the absorption (i.e. line of sight) covering fac- tor. The latter is assumed to be the same for all absorbers (but see also §4). The absorbers are assumed to be aligned such that the observed spectrum is the result of the intrinsic contin- uum passing through them all. The emission components are specified by their UOX and N (that are not necessarily the same as those of the absorbers), their covering factor and whether or not they are occulted by the absorbing gas. The dynamics and kinematics of the absorbing gas are im- portant factors in comparing the data with the model. Fol- lowing numerous UV observations, and our analysis in Pa- 6 NETZER ET AL. TABLE 3. COLUMN DENSITIESa per I, we assumed internal motion in the gas which is much larger than the thermal motion. This "turbulent velocity" is assumed to have a Doppler profile and the velocity quoted is the Doppler b parameter. Paper I showed that all absorp- tion lines with good S/N and sufficient resolution (the lines of O VII, Ne X, Mg XII and Si XIV) can be characterized by two kinematic components. The central velocities of these are between -400 and -600 km s−1 and between -1000 and -1300 km s−1, relative to the systemic velocity. A good represen- tation of the observations can be obtained with two Doppler profiles each with b ≃ 250 km s−1. The relative EW and covering factor of the two velocity components are critical to our modeling of the source. The observations show that the EW ratio of the velocity compo- nents is about 1:0.7 (lower:higher) in all lines with sufficient S/N (see below). Since this ratio is seen in several saturated lines, it can be interpreted as the result of different absorp- tion covering factors for each component. In this case, the lower-velocity component has a covering factor of 0.8 -- 1.0 and the larger-velocity component a covering factor of 0.6 -- 0.8. However, the composite profiles also include unsaturated lines that seem to have similar shapes. In this case, the dif- ferent EWs are due to different column densities. We cannot distinguish between the two possibilities since we do not have high-quality, high-resolution profiles for many weak lines. Given these uncertainties, we investigated two cases where the relative column densities for the two velocity components are in the ratio of 1:0.7. In one case the covering factor of all absorbers is unity and in the other case it is of order 0.8. In summary, each ionization component of the models pre- sented in this paper is made of two kinematic components. For the outflow velocities we chose -500 and -1000 km s−1, and assumed a turbulent velocity of 250 km s−1 for all com- ponents. This implies that the total derived column density of a certain ionization component is the sum of the column densities in the two kinematic components, even for saturated lines (since there is very little velocity overlap between the two). For brevity, in the rest of this paper we just quote the total column densities. Given these assumptions, we used the EWs measured in Paper I, combined with a few new mea- surements, to obtain the various column densities and optical depths. The more important lines that were used to constrain our models, and their adopted column densities, are listed in Table 3. 3.2.3. The silicon and sulphur line method A major clue for the conditions in the absorbing gas is ob- tained from EW measurements of various lines in the 5 -- 7.1 A band. This wavelength range contains lines from Si VII to Si XIV as well as the strongest lines of S XV and S XVI. Most of these lines are not blended. The range of ionization and ex- citation is very large and represents the ionization of almost all line-producing ions in the Chandra spectrum. The column densities deduced from these EWs, assuming the two compo- nent profiles, are given in Tables 1 & 3. The atomic data for the lines are known either from standard calculations (f-values for the H-like and He-like transitions) or from the recent work of Behar and Netzer (2002) discussing the inner-shell lines of silicon, sulphur and other elements. We have developed a simple algorithm to compare the mea- sured optical depths of all the observed silicon and sulphur lines with the results of the photoionization calculations. We first compute a large grid of models with calculated optical depths for a large range of ionization parameters and column Ion & Line Si XIV l 6.182 Si XIV l 5.217 Si XIV l 4.947 Si XIV adopted value Si XIII l 6.648 Si XIII l 5.681 Si XIII l 5.405 Si XIII l 5.286 Si XIII adopted value Si XII l 6.718 Si XI l 6.778 Si X l 6.859 Si IX l 6.931 Si VIII l 6.999 Si VII l 7.063 S XVI l 4.729 S XVI l 3.991 S XVI adopted value S XV l 5.039 S XV l 4.299 S XV adopted value Mg VIII l 9.506 Mg IX l 9.378 O VIII l 14.821 O VIII l 14.634 O VIII adopted value O VII l 17.396 O VII l 17.200 O VII adopted value O VI l 21.01 O VI l 19.341 O VI adopted value O V l 22.334 O V l 19.924 O V adopted value measuredb > 17.50c 17.94 ± 0.10 17.83 ± 0.20 17.90 ± 0.20 > 16.87c > 17.23c 17.57 ± 0.20 17.71 ± 0.25 17.65 ± 0.20 16.10 ± 0.06 16.29 ± 0.06 16.83 ± 0.05 16.66 ± 0.06 16.66 ± 0.08 16.26 ± 0.20 > 17.29c 17.58 ± 0.13 17.58 ± 0.13 16.87 ± 0.09 17.08 ± 0.25 17.00 ± 0.20 16.58 ± 0.05 16.39 ± 0.06 18.67 ± 0.25 18.59 ± 0.14 18.63 ± 0.25 18.03 ± 0.20 18.04 ± 0.20 18.04 ± 0.20 > 16.42c > 17.00c > 17.00 > 16.34c model log(Uox) −2.4 −1.2 −0.6 · · · · · · · · · 12.50 · · · · · · · · · · · · 14.68 15.57 16.43 16.90 17.04 16.81 16.13 · · · · · · 10.13 · · · · · · 12.70 16.77 17.06 · · · · · · 17.65 · · · · · · 18.39 · · · · · · 17.96 · · · · · · 17.54 · · · · · · · · · 17.00 · · · · · · · · · · · · 17.31 16.55 16.15 15.88 15.27 14.26 12.71 · · · · · · 16.23 · · · · · · 16.90 14.46 15.13 · · · · · · 18.05 · · · · · · 16.84 · · · · · · 14.34 · · · · · · 12.59 · · · · · · · · · 17.48 · · · · · · · · · · · · 16.94 15.41 14.11 13.10 11.68 9.93 7.80 · · · · · · 17.16 · · · · · · 17.00 11.13 12.61 · · · · · · 17.55 · · · · · · 15.49 · · · · · · 12.23 · · · · · · 9.3 total · · · · · · · · · 17.60 · · · · · · · · · · · · 17.46 16.62 16.61 16.94 17.05 16.81 16.13 · · · · · · 17.21 · · · · · · 17.25 16.77 17.07 · · · · · · 18.28 · · · · · · 18.40 · · · · · · 17.96 · · · · · · 17.54 17.07 ± 0.40 17.07 ± 0.40 aLog of column density in units of cm−2. bColumn densities and uncertainties derived from the EWs in Paper I. The EW was divided in the ratio 1:0.7, then the column density corresponding to the 1/1.7 part was computed assuming a Doppler width of 250 km s−1. The derived column density was multiplied by 1.7 to result with the total column density listed here. Uncertainties were calculated from the EWs uncertanties. See text for details. cSaturated line, lower limit only. densities. We then pick up to four models at a time and com- pare the combined optical depth for the chosen set of lines with the values deduced from the observations for the larger column density (lower outflow velocity) kinematic compo- nent. The best combination of models is obtained by mini- mizing the differences between the observed and the calcu- lated optical depths in all 10 lines (the 9 silicon and sulphur lines listed in Table 1 plus a line of Si VII). The result is a list of up to four models, with various column density and UOX , whose combination gives the best match to the observed col- umn densities. Using this procedure we find that at least three ionization components (i.e. six kinematic components) are re- quired to fit the data to within the observational accuracy. The reason is the very large difference in ionization between the lowest (e.g. Si VII and Si VIII) and the highest (e.g. S XVI) ionization lines. This requires at least one component to fit the EWs of the lowest ionization lines, one to fit the intermediate ionization lines and one for the highest ionization lines. Experimenting with various modifications of the method suggests that there are several combinations of three or four ionization components that give similar quality fits to the ob- VARIABILITY AND MODELING OF NGC 3783 7 Log(Uox)=−0.6 Log(N)=22.3 Log(Uox)=−1.2 Log(N)=22 Log(Uox)=−2.4 Log(N)=21.9 Log(UOX)=−0.6 Log(N)=22.3 Log(UOX)=−1.2 Log(N)=22.0 Log(UOX)=−2.4 Log(N)=21.9 A / 2 m c / n o t o h p d e z i l a m r o n 1 .5 0 1 .5 x u l f d e z i l a m r o n S Si 5 10 15 20 wavelength 25 30 35 0 4.5 5 5.5 6 wavelength [A] 6.5 7 FIG. 5. -- The three generic components that were used to model the low- state spectrum of NGC 3783. All components are shown on the same nor- malized scale where 1 is the incident continuum level. FIG. 6. -- A blown up version of Fig. 5 showing the 4.5 -- 7 A region used to constrain the model parameters (see text). Note that all three models are required to explain the large range of ionization (from Si VII to S XVI.) this column gives rises to an uncertainty on the slope. served optical depths of the silicon and sulphur lines. All combinations share the same general properties -- they all require at least three very different sets of physical condi- tions (i.e. ionization parameters) and three similar column densities. A generic model with the required properties is the following three-component model (all column densities refer to the total column of the two kinematic components): a low ionization component with log(UOX ) = −2.4 ± 0.1 and log(N) = 21.9 ± 0.1, a medium ionization component with log(UOX ) = −1.2 ± 0.2 and log(N) = 22.0 ± 0.15, and a high ionization component with log(UOX ) = −0.6 ± 0.2 and log(N) = 22.3 ± 0.2. The uncertainties on UOX and N in the case of the low ionization component arise primarily from the uncertainty on the slope of the ionizing continuum. Relax- ing this constraint allows a somewhat smaller column density (see §3.2.4). The larger uncertainties on the parameters of the higher ionization components are due to the fact that gas with such properties produce similar EWs for medium (e.g. Si XI) and high (e.g. Si XIV) ionization lines over a larger range of ionization parameters and column densities. The three the- oretical ionization components are plotted over a large wave- length range in Fig. 5, and a more detailed view of the 4.5 -- 7 A range is shown in Fig. 6. 3.2.4. Intrinsic continuum and global model properties The next step is to test the combination of components that fit the optical depths of the 5 -- 7.1 A silicon and sulphur ions best over other wavelength bands. This requires the shape of the underlying ionizing continuum to be defined more pre- cisely. The slope of the short wavelength continuum can be measured directly if bound-free absorption is negligible. However, Fig. 5 clearly shows that two of the generic ab- sorbers are characterized by large opacity at wavelengths as short as 4 -- 5 A. Thus a unique determination of the 2 -- 5 A con- tinuum slope (the part available for the grating observations) is not trivial. More specifically, the exact O VII column den- sity in the lowest-ionization component strongly influences the short wavelength continuum shape, and the uncertainty on Given the above constraints on the model properties, and the observed 2 -- 5 A continuum, we have experimented with various power-law continua with the requirement that both the short wavelength continuum and the measured EWs of the silicon and sulphur lines, are within the observational un- certainties. All fits were performed on the low-state spec- trum. We found that a single power-law (where the number of photons per unit energy, E, is proportional to E −G ) with G = 1.65 ± 0.15 is a good overall representation of the low- state spectrum. We note that the limited data quality cannot exclude the possibility of some steepening of the intrinsic con- tinuum at long (l > 20 A) wavelengths. We note in passing the different continuum slope adopted by K03 in their recent modeling of the source. These authors did not consider the two spectral states found here and fitted the full Chandra data set. As a result, the SED they adopted is based on the combined (high- and low- state) spectrum, and hence is different from ours. This is also the reason why their powerlaw slope is flatter than the one adopted by Kaspi et al. (2001), who fitted the high state only. K03 also include in their SED a low-energy blackbody component. Another uncertainty on the intrinsic SED arises from the fact that some of the high energy photons are likely to be produced far away from the source, due to Compton scatter- ing and reflection of the central continuum radiation. This can be very noticeable in faint sources, like NGC 3516 dur- ing 2000 (see Netzer et al. 2002). Scattering by gas in a low state of ionization is strongly wavelength dependent, and for NGC 3783 can amount to a change of ∼ 0.1 in the de- rived value of G . As a result, we might have overestimated the column density of the lowest ionization component (the only component with significant influence on the slope of the short wavelength continuum) which would results in overesti- mating the column densities of low ionization species such as O V and O VI. §4.1 contains some discussion about the low- ering of the column density in our low UOX component and a more detailed examination of this effect is deferred to a sepa- 8 NETZER ET AL. rate paper (George et al. 2003, in preparation). Finally, a broad, relativistic iron Ka line can also be impor- tant in affecting the short wavelength continuum slope. The presence or absence of such a feature is an open question. Kaspi et al. (2001) see no evidence for this feature in the 2000 Chandra observation. Given the uncertainties discussed above, and the large amount of absorption at almost all wave- lengths, we cannot resolve this issue even with our much im- proved data. Given the chosen continuum and the three ionization com- ponents, we have calculated a combined multi-component model that covers the entire wavelength range. As stated above, the model includes two kinematic components for ev- ery ionization component. The theoretical spectrum is com- pared with the observed low-state spectrum in Fig. 7. The re- sults are very good. In fact, we could not find another combi- nation of models (three or more absorbers) that is significantly superior. Not only does the model provide a good overall fit to many lines of all elements, but it also predicts the strengths of the deep bound-free edges correctly. Thus, the constraints on the 5 -- 7.1 A silicon and sulphur lines are enough to com- pletely specify a model for the entire spectrum. 3.2.5. Deficiencies in the Modeling As demonstrated above, the low-state absorption spectrum of NGC 3783 can be satisfactorily explained by our combina- tion of line-of-sight absorbers. The agreement with the ob- servations is very good for all lines with l < 15 A. However, there is a noticeable discrepancy at around 16 A. At this wave- length range there are two main sources of opacity -- the O VII bound-free continuum (edge at 16.7 A), and the iron M-shell UTA feature (Behar, Sako & Kahn 2001). Our photoioniza- tion calculations suggest a disagreement between the two in the sense that the strongest iron lines observed are due to ions indicative of a lower level of ionization than that indicated by the strongest oxygen lines observed. Specifically, the peak UTA absorption corresponds to Fe VIII -- Fe X. For such a level of ionization, our models predict most of the oxygen being in the form of O III -- O VI. However, the observations indicate that most of the oxygen bound-free absorption is due to O VII. The possible origin of this discrepancy is discussed in §4. Regarding the spectrum beyond 19 A, the comparison be- tween the model and the observations is limited by the poorer S/N. While the model shown in Fig. 7 is consistent with the data, the predicted depths of several absorption lines in this range (in particular those due to O V and O VI) seem to be stronger than actually observed. This may be a real short- coming of the model (see comments in §3.2.4 and in §4). It may also be the result of the assumed absorption cover- ing factor. As already mentioned, line-profile analysis sug- gest a mean (over the line profile) covering factor of 0.7 -- 1.0, and all models considered so far assumed the extreme case of an absorption covering factor of unity. The "leakage" of the incident continuum makes little difference at short wave- lengths except for a need to somewhat decrease the assumed . However, leakage can influence the comparison at long wavelengths much more. This is illustrated in Fig. 8 which shows a model with G = 1.5 and a line-of-sight covering factor of 0.85. Indeed, the agreement at long wavelengths is much better. The differences between the two assumed slopes and covering factors are well within the model uncertainties. We have attempted to model the high-state spectrum by as- suming the same ionization components and by changing the luminosity and the ionization parameter by the factor inferred from the observed increased luminosity in the short wave- length continuum (∼ 1.5). The result is a very poor fit to the long wavelength continuum luminosity and to the inten- sities of many lines. The reason is that such a small lumi- nosity change results in an opacity change which is too small to account for the very large difference (factor ∼ 4) between the low and the high state spectra at long wavelengths. This means that the high and the low state spectra are inconsistent with the assumption of a simple response of the gas to contin- uum luminosity variations. There is a simple and satisfactory solution for the high-state spectrum that involves the appearance of an additional con- tinuum component. This second, long wavelength component (sometimes referred to as a "soft excess") appears only during high state. We have therefore modeled the high-state contin- uum by two powerlaws: the low-state power law (G = 1.6) and a steeper component with G = 3.2 which dominates at long wavelengths. We find a satisfactory solution for the high-state spectrum when the G = 3.2 component emits 40% of the flux of the G = 1.6 component at 12.398 A (1 keV). (We note that this is not a unique combination, and fits of similar quality can be obtained for other combinations of slope and normal- ization.) The two continua are shown in Fig. 3, along with the high and the low state spectra. Given the two component con- tinuum, we find a very good agreement for the high-state line and continuum spectrum by assuming no opacity variations. The quality of the fit is similar to the low-state fits shown in Figs. 7 & 8. To further illustrate this point in Fig. 9 we show the high- state spectrum divided by the low-state spectrum (binned to reduce the noise) along with the ratio of the two assumed con- tinua. The diagram shows that any remaining spectral features are entirely consistent with the noise. The S/N in this diagram is not high enough to completely rule out some opacity-like variations at long wavelengths. In particular, there is some ex- cess emission near the O VII and O VIII recombination edges that may hint to extra emission in the high-state spectrum (note that we do not expect the emission features to disap- pear by this division). However, such opacity variations must contribute very little to the spectral variations observed at long wavelengths. Finally, the emission-line spectrum of NGC 3783 can be ex- plained by assuming X-ray emitters with the same properties found for the ionized absorbers. The emitted line photons are probably observed through the absorber, as indicated by the combined emission-absorption profiles of several resonance lines. The model is problematic at around 21.8 A, where it fails to reproduce the O VII intercombination emission line. The emission covering factor required is about 0.1 (see §4.1) consistent with the value of approximately 0.05 -- 0.15 obtained by Behar et al. (2003). 4. DISCUSSION This section discusses the main results of the new analysis of all the Chandra observations of NGC 3783 performed to date. The reader is referred also to an appendix where we give a detailed comparison with the recent work of Krongold et al. (2003). The main new findings of our work are the large increase in flux between the low- and high-states, which is not accom- panied by any significant changes in the properties of the ab- sorption. The central source in NGC 3783 fluctuates between two such flux states on a 20 -- 120 days time scale, and the best G VARIABILITY AND MODELING OF NGC 3783 9 14 12 10 8 6 4 2 8 6 4 2 0 8 6 4 2 0 8 6 4 2 0 14 12 10 8 6 4 2 0 4 4.5 5 5.5 6 6.5 7 7.5 8 8.5 9 9.5 10 10.5 11 11.5 12 12.5 13 13.5 14 14.5 15 15.5 16 16.5 17 17.5 18 18.5 19 19.5 20 20.5 21 Rest Wavelength [Å] 21.5 22 ] 1 - Å 1 - s 2 - m c s n o t o h p 4 - 0 1 [ x u l F FIG. 7. -- A comparison of the low state spectrum with the six component model (three ionization components each split into two kinematic components). Note that most of the apparent discrepancy at around 5 -- 6 A is due to a known calibration problem (H. Marshall, private communication). description of the observed variations is the appearance and disappearance of a "soft excess", low energy continuum com- ponent. These lead to important conclusions regarding the nature of the continuum source. Other important conclusions are the multi-component nature of the absorber and the large range of ionization. Below we discuss some interpretations. 4.1. Physical properties of the absorbing gas We found that three ionization components with different properties and a large range of ionization, are required to fit 10 ] 1 - Å 1 - s 2 - m c s n o t o h p 4 - 0 1 [ x u l F NETZER ET AL. 14 12 10 8 6 4 2 0 19 19.5 20 20.5 21 Rest Wavelength [Å] 21.5 22 FIG. 8. -- Model vs. observation for the long wavelength spectrum assuming G = 1.5 and a line-of-sight covering factor of 0.85 for all absorbers. w o l / h g h i 6 4 2 0 5 10 15 rest wavelength [A] 20 25 FIG. 9. -- A ratio spectrum of NGC 3783: Points with error bars are the result of dividing the high-state spectrum by the low-state spectrum. The binning is 0.04 A for l < 11 A and 0.1 A at longer wavelengths. The smooth line is the same division applied to the two continua shown in Fig. 3. the spectrum of NGC 3783. Each of those components is split into two kinematic components, as implied by the ob- served line profiles reported in Paper 1. The three "generic ab- sorbers" of our model are listed in Table 4 and shown in Figs. 5 & 6. As explained above, the number of real components and the exact values of UOX and N are not very meaningful since several somewhat different combinations of ionization components give equally good fits. Rather, the values chosen here are representative, "average" ionization parameters, and the total column densities required. Our modeling of the low- TABLE 4. PARAMETERS OF THE THREE GENERIC MODELS Model log(UOX ) log(N) Low ionization Intermidiate ionization High ionization -2.4 -1.2 -0.6 21.9 22.0 22.3 state spectrum assumes the underlying X-ray continuum is a single power-law. However, the S/N of the long wavelength spectrum is not high enough to completely rule out the pres- ence of an additional softer continuum component, similar to the one seen during the high states but of much lower lumi- nosity. We also note that there is little (if any) response of the absorbing gas to the short (∼ 2 days) time scale contin- uum variations. Thus our measurements are models represent some time-averaged spectral properties. The main difference between the present model and the one given by Kaspi et al. (2001) is the presence of the low- est ionization component. Only the first Chandra observa- tion (obtained in 2000 January) was available to Kaspi et al., and hence was insufficient to reveal the two spectral states. Furthermore, this relatively short observation resulted in the S/N at l > 18 A which was insufficient to reveal the true shape of the continuum. Hence Kaspi et al. naturally (but incorrectly) assumed the underlying continuum was a single power-law for this high-state spectrum. As a result, their preferred model contained a component with a strong O VII bound-free edge, but very little absorption at longer wave- lengths (i.e. at l > 18 A the underlying continuum was "re- covered"). Our new observations do not show such a recovery. In fact, the absorption in the long wavelength region is dom- inated by features of K-shell carbon and nitrogen and L-shell nitrogen and oxygen. The presence of a lower ionization component is also a key VARIABILITY AND MODELING OF NGC 3783 11 point in the K03 work. In their study they assume two ab- sorbers with log(UOX ) = −2.77 (i.e. a factor of about two lower than our lowest ionization component) and log(UOX ) = −1.23 (very similar to our imtermediate ionization compo- nent). Their fit does not require a third, very ionized com- ponent. Our reason for suggesting this higest-ionization (log(UOX ) = −0.6) component is based on the fit to the lines of Si XIV, S XV and S XVI. When this component is not in- cluded, the discrepancy in column densities of these ions is a factor 2 -- 3. Several of the lines of these ions are in fact missing from the K03 analysis. Moreover, some of the high- est -ionization lines are badly saturated, (e.g. Si XIVl 6.18 A). Thus a comparison of the observed and the calculated EWs of those lines, as performed by K03, is not sufficient for evaluat- ing the ionic column densities. Perhaps the largest observed discrepancy between model and observation is due to absorption by the iron M-shell UTA. This feature has been observed in several other AGN (Steen- brugge et al. 2003 and references therein) with similar shape and central wavelength. Its smooth shape is due to the large number of absorption lines of many iron ions and the wave- length of largest absorption is ionization dependent. Our mod- eling of this spectral region includes all the lines in the Behar et al. (2001) calculations (i.e. not only the shorter published list of lines). The central observed wavelength of the fea- ture in NGC 3783 is at around 16.4 A suggesting that most of the contribution is due to Fe VIII -- Fe X. For the assumed SED, this corresponds to gas whose dominant oxygen ions are O III -- O VI. In our modeling, most long wavelength ab- sorption is due to the component with log(UOX ) = −2.4 (Fig. 5). The dominant ions in this components are O VI -- O VII and Fe X -- Fe XII. This corresponds to peak UTA opacity at around 16.1 A. As shown in Fig. 7, the model fits the observed oxy- gen lines and continua reasonably well but badly misses the position of the UTA absorption. We have experimented with various other models to try to eliminate this discrepancy. For example, an overabundance of iron relative to oxygen may decrease the level of ioniza- tion of iron due to the increased iron opacity. Experiment- ing with iron abundance which is three times larger resulted with changes that are not large enough to explain the observed discrepancy. We have also experimented with models that in- clude several low ionization components, instead of the single low UOX component shown here. This, again, gave very lit- tle improvement. It seems that the apparent conflict between the oxygen and iron ionization cannot be resolved by these models and we suggest two alternative explanations. 1. The absorbing gas in NGC 3783 may not be in ioniza- tion equilibrium due to its low density and the rapid flux variations on time scales that are shorter than the recombination and ionization time scales. This is prob- ably not very important for the short wavelength flux since its variability time scale is short and the ampli- tude not very large. The gas responding to this con- tinuum is probably at some mean level of ionization. This however is not the case for the soft excess con- tinuum which varies on a much longer time scale with a much larger amplitude. This can contribute, signifi- cantly, to the ionization of the lower ionization species. Different ions react on different time scales and the gas may never reach an equilibrium. Since all our mod- els assume steady state gas, they may not be adequate to describe the absorbing gas properties. In particular, the iron and oxygen recombination times may be dif- ferent enough to result in gas where iron is less ionized or oxygen is more ionized compared to the equilibrium situation. The complicated issue of time dependent ion- ization is beyond the scope of this paper. 2. The ionization balance of iron depends critically on low temperature dielectronic recombination (DR) rates that are not well known for the iron M-shell ions. Experi- ence with L-shell iron ions, whose DR rates have been measured and computed recently (Savin et al. 2002 and references therein), shows that in almost all cases the previous low temperature DR rates were consistently smaller than the newly calculated values. Assuming a similar effect in the M-shell ions, we can envisage a sit- uation where, in the absence of realistic DR rate, the en- tire ionization balance of iron is shifted toward higher ionization. Thus, more realistic low temperature DR rates may bring models to better agreement with the ob- servations. We note that low temperature DR rates are also not available for the magnesium and silicon ions that we have modeled. For some of these ions (Mg VIII and Mg IX) these are probably not very important and for others (the relevant silicon ions) the situation is less clear (D. Savin, private communication but see also Gu 2003). Nevertheless, since we do obtain good agree- ment between the calculated EWs of these lines and the EWs of other lines observed in the spectrum, we did not consider changing those rates in the calculations. It is interesting to note that the K03 model assumes a lower level of ionization for all elements, including iron, and hence provides a better fit to the UTA. However, their lower ion- ization component fails to fit the Si X and Si XI lines around 6.8 A. Our model was designed to give a very good fit over the 5 -- 7 A wavelength range and hence produces a lower quality fit to the UTA feature. Thus, there seem to be no satisfac- tory solution based on a single low-ionization component that explains all these features. An interesting and somewhat problematic issue is the abun- dance of O VI, which can also be clearly observed in the far UV (Gabel et al. 2003a; Gabel et al., 2003b). Paper I found even the Ka line of O VI undetectable with the cur- rent S/N. However, revisiting this measurement more care- fully, and with more accurate atomic data, we can now detect several O VI lines. From two such lines, both of which are probably saturated, we have obtained a lower limit of 1017.0 cm−2 on the column density of O VI. This is much higher than the value (1016.0±0.3 cm−2) obtained from the XMM-Newton RGS spectrum (Behar et al. 2003). Moreover, the values pre- dicted from our model (1017.96 cm−2) is even larger. Such a large value is problematic for several reasons. It may be in contradiction with the UV observations of the source (see be- low), it is much larger than the values obtained from the RGS observations, and it results in a prediction of a strong absorp- tion line at around 21.87 A which is not seen in the spectrum (Fig. 7, the absorption line underneath the O VII intercombi- nation line). An explanation of the discrepancy between the values for the O VI column density measured by Chandra and XMM- Newton is the crowded region of the spectrum where O VI Ka resides. In particular, the lower spectral resolution of the XMM-Newton RGS makes it extremely difficult to resolve the weak O VI absorption line at 22.01 A in the presence of the 12 .8 .6 .4 .2 0 n o i t i a z n o i l a n o i t c a r F oxygen silicon magnesium VII VIII VIII IX V VII VII VI IX −3.2 −3 −2.8 −2.6 −2.4 −2.2 −2 −1.8 Log UOX FIG. 10. -- The fractional ionization of oxygen, magnesium and silicon ions in a low density gas of our chosen composition exposed to the NGC 3783 low-state continuum. The curves show the mean ionization of various ions averaged over a column density of 1021.5 cm−2. Note the great similarity in the ionization of O VI, Mg VIII and Si VIII (all plotted with dotted lines). bright O VII forbidden emission line (22.10 A). We also wish to note that with the lack of laboratory measurements for the O VI wavelengths, high-resolution astrophysical spectra such as the present one constitute the best determination of these wavelengths currently available. The Chandra data, along with the assumption that O VI is outflowing at the same ve- locity as O VII, places the O VI Ka line at a rest-wavelength of 22.01 ± 0.01 A. This value is in agreement with our model wavelength of 22.005 A, calculated with the HULLAC atomic code, as well as with the value measured in NGC 5548 (Steen- brugge et al., 2003), but is somewhat different from the value of 22.05 A calculated by Pradhan (2000) using the R-matrix method. For more information see Behar and Kahn (2002). A O VI column density less than 1017 cm−2 is also incon- sistent with theoretical predictions based on modelling the abundances of Mg VIII, Si VIII and O VI. This is illustrated in Fig. 10 where we show calculations for the fractional abun- dance of oxygen, silicon and magnesium ions as a function of UOX assuming an incident continuum similar to the one used here. The diagram shows the very similar fractional abun- dance of O VI, Mg VIII and Si VIII as a function of ioniza- tion parameter. Using the measured EW of the Si VIIIl 6.998 line, which is an unblended, easy-to-measure feature, we can use Fig. 10, and the assumed silicon-to-oxygen abundance ratio, to derive an estimate for the column density of O VI. This number is very close to the model prediction. The sit- uation is more complicated regarding magnesium since sev- eral of the relevant lines (around 9.4 A) are blended with neon lines. However, we could obtain an estimate of the EW of the Mg VIII l 9.506,9.378 lines that are good enough to constrain the Mg VIII column density. The O VI column density based on this measurement is, again, much larger than 1017 cm−2 and in good agreement with the model prediction. A comparison with the K03 work shows a disagreement by a factor of two in the total (hydrogen) column density of the lowest-ionization absorber. One reason for the difference is NETZER ET AL. that K03 analyse the combined (high- and low-state) spec- trum, which is then interpreted in terms of a different amount of bound-free absorption. Another reason is the somewhat larger oxygen abundance assumed by K03 (the oxygen col- umn densities of the low-ionization components in the two papers actually differ only by a factor 1.35). We also note that the inner-shell O VI lines that we have used to determine the limit on the column density of this ion are missing from the K03 model. This makes a noticeable difference in the spectral fitting at around 21 A. Finally, we have looked for the lowest O VI column density allowed by the observations if we remove the constraint im- posed by the short wavelength continuum slope (§3.2.4) and keep the ionization parameter unchanged. This means find- ing the lowest column density for the log(UOX)=-2.4 compo- nent which is still consistent with the measured EWs of the 5 -- 7.1 A silicon and sulphur lines. We found this limit to be log(N) & 21.7 for the low-state spectrum, and log(N) & 21.6 for the high-state spectrum. These lower limits on the column densities are consistent with the uncertainties given in Table 1. Under these assumptions, the column density of O VI in the low-state spectrum can be as low as 1017.76 cm−2. This column is large enough to produce saturated O VI lines at 21.01 and 21.87 A. It also requires a much harder continuum (G = 1.4). 4.2. Outflow velocity and covering factor Some of the parameters our model are based on measured profiles of several absorption lines that were shown in Paper I to include at least two components with different ouflow ve- locities. The measurements in Paper I also show that the rela- tive EWs of the two components seems to be ionization inde- pendent and that most of the lines used for the profile analysis were heavily saturated. This suggests that the apparent opti- cal depths of those lines are determined by the covering fac- tor rather than the line opacity, and hence similar to the case seen in many UV absorption systems in AGN (e.g. Barlow, Hamann & Sargent 1997;Arav et al. 1999). The covering fac- tor of the lower-velocity components of the O VII, Ne X and Si XIV lines is in the range 0.8 -- 1.0. The covering factor of the larger-velocity component of the same lines, assuming satu- rated profiles, is 0.6 -- 0.8. We also note that the large number of iron lines near the UTA center gives an independent esti- mate of the covering factor which is at least 0.85. Fig. 8 shows that models with a covering factor less than unity provide fits to the data of similar quality. Using the sil- icon and sulphur lines (as in §3.2.3), we find three ionization components are again required, each with parameters similar to those listed in Table 4. The only difference is the need to assume a somewhat harder ionizing continuum. The fit to the long wavelength lines is in fact somewhat better than that produced by the full covering models. However given the low S/N at those wavelengths, this is not a strong con- clusion. Thus, the data-model comparison cannot constrain covering factor beyond the actual observations. The most im- portant conclusion is that the covering factor is similar in low (e.g. O VII and the iron UTA) and in high (e.g. Si XIV, Ne X) ionization lines. 4.3. Density, location and thermal stability of the absorbing gas Our measurements and analysis are consistent with no vari- ations in the absorbers' properties on time scales of 1 -- 4 days. VARIABILITY AND MODELING OF NGC 3783 13 There are also indications for no spectral changes over much longer periods, perhaps a few months. For instance we find no evidence for any significant, narrow (< 1 A) spectral features in the high- to low-state spectral ratio shown in Fig.9. The reason we cannot constrain the variability of the absorbers on short time-scales are the fast, relatively-small amplitude fluctuations in the luminosity of the central source (maximum amplitude of about a factor 2). As a result of these variability characteristics of the illuminating continuum, the lack of sig- nificant spectral variations may be due to the ionization level of the gas reflecting the time averaged ionizing luminosity, rather than very long ionization or recombination times. This conclusion therefore depends primarily on the direct compar- ison between the low and the high state spectra, and thus on the S/N in the lower-quality spectrum used for the comparison (the high state spectrum). To further examine this point, we have calculated the theo- retical spectrum appropriate for the high state continuum un- der the assumption that the absorbing gas responds instantly to changes in the ionizing continuum. This would be the case if the gas along the line-of-sight gas were of very high den- sity. The initial conditions are those assumed for the low-state continuum, and the change in the SED as described earlier: a two component X-ray continuum, where the long-wavelength component is assumed to extrapolate down to 40 eV. The in- crease in flux between the low- and high states was assumed to be a factor 1.5 at 4 A (see Fig. 3). Fig. 11 shows two the- oretical spectra: the low-state spectrum (in blue) is the same model shown in Fig. 7, the high state spectrum (in red) is the result of these new calculations. The two are clearly differ- ent. In particular the Si VII, Si VIII and Si IX lines are are much weaker for the high-state model. This is due to the large increase in flux at longest wavelengths, with the largest consequences for the lowest ionization component (UOX has increased by a factor close to 2 and UX by a factor of 6.6). The variations predicted by this theoretical exercise are much larger than those observed (Table 1 and Fig. 4). This illus- trates the gas did not simply react to the differences in the illuminating continuum between the low- and high-states. An even stronger conclusion is obtained from the fact that at long wavelengths the differences observed between the two states are much larger than can be explained by simple variations in opacity. Given the above findings, we can place limits on the elec- tron density (ne) and the distance from the central source (D) for each absorption component. We used the temperatures derived from the models, average recombination times for the dominant oxygen and silicon ions, and assuming there is no response to continuum variations on a time scale of 10 days. We find ne < 5 × 104 cm−3 and D > 3.2 pc for the log(UOX ) = −2.4 component, ne < 105 cm−3 and D > 0.63 pc for the log(UOX ) = −1.2 component , and ne < 2.5 × 105 cm−3 and D > 0.18 pc for the log(UOX ) = −0.6 compo- nent. The corresponding masses are 8.1 × 103Cf , 390Cf and 64Cf solar masses, respectively, where Cf is the absorption (4p ) covering factor. The mass outflow rate is dominated by the lowest ionization component and is approximately 75Cf e voutflow/(500 km s−1) solar masses per year, where e is the radial filling factor of the flow. The very large num- ber suggests a short duration outburst rather than continuous ejection. We also note that the covering factor can be small, of order 0.1. The above distances are in agreement with the find- ings of Behar et al. (2003), who have suggested the outflow is x u l f d e z i l a m r o n 15 10 5 0 low state high state 6.2 6.4 6.6 6.8 7 7.2 wavelength [A] FIG. 11. -- Low state and high state theoretical spectra assuming the ob- served variable NGC 3783 continuum and line of sight gas that responds in- stantly to continuum variations. The low state model (in blue) is the one described earlier and shown in Fig. 7. The high state spectrum (in red) as- sumed the high state SED discussed in the text. Such a continuum results in much weaker Si VIII and Si IX lines. located at a distance of a few pc and is extended beyond 10 pc based on a comparison between the soft X-ray emission and absorption characteristics of the XMM-Newton data. A significant new aspect of our model is that the product ne × T is similar (within a factor .2) for all three absorp- tion components. This raises the interesting possibility that all the components are in pressure equilibrium (assuming gas pressure dominates the total pressure, i.e. radiation pressure and turbulent pressure are not important), and they all occupy the same volume of space. To test this idea, we have calcu- lated thermal stability curves (log(T ) vs. log(UOX /T ) for the low-state continuum and our assumed composition (Table 2) under various assumptions. We have kept the UV SED un- changed and varied G (0.1 -- 50 keV) over a wide range. We have also investigated the possibility of a line-of-sight atten- uation by the absorbing gas. This can be relevant for the sit- uation under study since, even in the simplest geometry, two of the absorbers do not have a clear view of the central radi- ation source. Screening by the lowest-ionization component (log(UOX ) = −2.4) is the most important since this gas mod- ifies the transmitted spectrum much more than the other com- ponents (see Fig. 5). Three such stability curves are shown in Fig. 12, along with the location of the three ionization com- ponents. The exact location depends on the details of the as- sumed SED, the gas metallicity and opacity. The interesting feature of the stability diagram is that all ra- diation fields considered here result in extended, almost verti- cal parts between T ≃ 3 ×104 K and T ≃ 2 ×106 K. The curve for G = 1.6 has a more extended vertical part since the higher mean energy of the radiation field results in a Compton tem- perature. Screening by the lowest-ionization component in NGC 3783 is indeed important, but does not change the main result that, given the uncertainties on the SED and compo- sition, all three ionization components lie on the stable part of the curve. We also note that, while here we only consid- ered gas with a single composition, the details of the stability 14 7 6 5 4 ) T ( g o l G =1.6 G =1.8 −8 −7.5 −7 −6.5 log(UOX/T) −6 −5.5 −5 FIG. 12. -- Thermal stability curves for a low density gas exposed to the low-state continuum of NGC 3783. The two curves marked by G are for a bare continuum with a 0.1 -- 50 keV slope as marked. The dotted line is the stability curve for a gas exposed to the G = 1.8 continuum seen through the log(UOX ) = −2.4 absorber. Thick sections mark the locations of the three absorbers considered in this work with the corresponding uncertainties on the ionization parameters. All ionization components considered in this work are situated on thermally stable parts and they all have roughly the same gas pressure. curve are abundance dependent. Needless to say, very differ- ent curves will be obtained for significantly different SEDs. We note that K03 reach similar conclusions regarding the pressure equilibrium of their two ionization components. However, in our model all components are situated on stable branches of the thermal stability curve (Fig. 12) while accord- ing to K03, their high-ionization component is situated on the unstable branch (their Fig. 16). We suspect that the difference is in part due to the different SED and metalicity assumed by K03, and in part due to the method they used to calculate the curve. The existence of an additional stable-region along the curve of thermal-equilibrium (intermediate between the cold and the hot branches) where "warm" ( f ew × 105 K) can survive is not a new idea. It has been discussed by Marshall et al. (1993) in their study of the multi-phase medium in NGC 1068, and later by many others. Reynolds & Fabian (1995) discussed the stability criteria and noted the narrow range of ionization parameter allowed for warm absorbers. Hess et al (1997) dis- cussed the various parameters governing the stability curve (metallicity and more). Komossa and collaborators (e.g. Ko- mossa & Fink, 1997; Komossa & Meerschweichen 2001) cal- culated many such curves, for different SEDs and metallici- ties, and discussed the location of warm absorbers as well as the possible link between the NLR and the X-ray gas. Kro- lik & Kriss (2001) considered multi-phase, warm-absorber winds and noted the various locations of stable, over-heated and over-cooled gas. Kinkhabwala et al. (2002) re-visited the thermal stability issue in NGC 1068 and showed that a wide range of UOX is needed to account for the observed X- ray emission lines. Chelouche & Netzer (2003) investigated the changes in the stability criteria for clouds with large inter- nal line radiation pressure. NETZER ET AL. Most of these earlier works focused either on the allowed parts of the curve from general stability conditions, or on the location of a certain absorber observed in a certain source. Our work shows that three very different aborbers in a source can all occupy extended, stable parts of the curve where the gas pressure is roughly the same. This means there may well be more than three absorbers spread over the vertical branch of the curves in Fig. 12. The three ionization parameters con- sidered here may therefore represent some volume-averaged properties of the entire cloud ensemble. The column densi- ties are therefore the total column densities of a large num- ber of such clouds. Moreover, the most-ionized component may provide the confining medium for the two other ioniza- tion components. Needless to say, real confined outflowing clouds are characterised by a more complicated density and pressure structure than with the simplified, constant density clouds considered here (e.g. see Chelouche & Netzer 2001). The pressure equilibrium between all three absorbers in NGC 3783 suggest they occupy the same general location. This allows additional constraints to be placed on their density and location. Specifically, we require that the distance of this region is at least 3.2 pc (the minimum distance of the lowest ionization component). However, this zone cannot be much further away because of total size and volume considerations. Under the assumed conditions, the highest-ionization param- eter component has the lowest density, and hence the largest dimension. However, its radial extent cannot exceed about 500 pc (the approximate radius of the narrow-line region), and thus its mean density cannot fall below about 10 cm−3. Pressure equilibrium then dictates the density of the lowest- ionization component, and a maximum distance of about 25 pc. This limitation is only imposed on discrete clouds -- i.e. on the mean properties of the gas. Real absorbers may not be "clouds", they can cover a large range of distances, and can have properties which depend on location. The detailed in- vestigation of such models is beyond the scope of the present paper. 4.4. The emission lines There is no indication for flux variation in any of the ob- served emission lines. However the measurement uncertainty on these line is rather large since the strongest lines are ob- served at the long wavelengths where the S/N is poor. We obtain satisfactory fits for most lines by assuming that the emission -- line gas has the same column density and ionization parameter as the absorbing gas. The covering factors required for the emitting gas are 0.2 -- 0.3 for the log(UOX ) = −2.4 com- ponent, 0.1 -- 0.2 for the log(UOX ) = −1.2 component, and ∼ 0.1 for the log(UOX ) = −0.6 component. The large range is due to the uncertainties in the continuum placement and in estimating the fraction of the O VII and O VIII emission lines that are being absorbed by the line-of-sight gas (the model shown in Figs. 7 and 8 assumes that all the emitted photons are seen through the absorbers). A clear shortcoming of the model is the fitting of the in- tercombination O VII line. The intensity predicted for this line is too small by a large factor (2-4). A large part of the discrepancy is due to the O VI l 21.87 A absorption line sit- uated close to this emission (Fig. 7). This may be related to the O VI problem discussed earlier. Alternatively it may be a consequence of our assumption that all emission-line photons are seen through the absorbers. Indeed, a model that assumes no absorption of the emitted photons by the line-of-sight gas (not shown here) would give a much better agreement to the VARIABILITY AND MODELING OF NGC 3783 15 O VII absorption complex. This would also required a smaller emission covering factor. Obviously we cannot determine the exact geometry using current observations, and the real case may well be intermediate between those two extremes. We also note on the difference between the absorption cov- ering factor (0.8 -- 1.0) and the emission covering factor (0.1 -- 0.3). Given the uncertain geometry and the fact that the first is a line-of-sight covering factor and the secong a globel (4p ) property, the two are perhaps consistent with each other. We did not investigate all possibilities and cannot comment in de- tail on special cases like biconical flows or a torus geometry (e.g. Krolik & Kriss 2001). Finally, comparing our method with that of K03 we note that those authors do not calculate the emission line flux and profile in a self-consistent way. Instaed, they add Gaussian shaped emission profiles to their calculated absorption pro- files (e.g. their Fig. 6). This can result in significant differ- ences regarding the line flux and EW. For example, our default approach assumes that all emission lines are seen through the absorber. In such a case, the entire "blue" wing of many emis- sion lines is absorbed. Clearly this affects the EW of the emis- sion lines, but has almost no effect on the EW of the absorp- tion features. In contrast, the approach adopted in K03 always results in a decrease of the EW of the absorption lines. 4.5. UV absorbers in NGC 3783 NGC 3783 contains time-variable absorbers in the UV that have been described in various earlier publications (Gabel et al. 2003a and references therein). Our analysis clearly indi- cates that at least one X-ray component contains gas which is of sufficiently low ionization so as to produce strong absorp- tion lines of C IV, N V and O VI in the UV. The predicted EWs of these UV lines depend on two fac- tors that have been given little attention in this work: the ex- act shape of the UV continuum, and covering factor appro- priate for the UV absorbers. A UV continuum which is softer than that assumed here can result in higher levels of ionization for carbon, nitrogen and oxygen, with little effect on most X- ray lines. This is very important for the C IVl 1549 A and N Vl 1240 A lines, but less so for the UV lines of O VI that are directly linked to the O VI column density discussed here. As for the covering factors, the size of the UV continuum source is likely to be much larger than the size of the central X-ray source. Thus the covering factor appropriate in the UV can be different to that appropriate in the X-ray. This will re- sult in saturated UV absorption lines whose EWs are smaller than the ones indicated by the column densities derived here. However, our upper limit on the density of the low-ionization component, combined with its column density, suggest a very large line-of-sight dimension (> 1017 cm). Assuming lateral dimension of the same size or large, we conclude that the physical dimension of this component is much larger than the expected dimension of any likely UV source (e.g. the surface of a thin accretion disk). A combined analysis of the UV and X-ray results, taking these such points into consideration, is in progress UV absorbers are also UV emitters, and the low-ionization component can also contribute to the observed UV emission lines. Our photoionization calculations show that this con- tribution is very small. For example, an emission covering factor of 0.1, similar to the one deduced from the X-ray emis- sion lines, will produce O VIl 1035 line with emission EW of about 1 A. 4.6. The long wavelength component Perhaps the most interesting result of this study is the ap- pearance and disappearance of the soft excess component. This broad-band continuum source was seen in two of the six observations. On average, these are the two observations with the highest luminosity. An increase in softness ratio with increasing source luminosity is well known and well docu- mented in a number of AGN (e.g. Magdziarz et al., 1998; Chiang et al., 2000; Markowitz & Edelson 2001). However we are not aware of a softness ratio increase that is not corre- lated with a short wavelength flux increase. NGC 3783 seems to be the first AGN to show this phenomenon, but we suspect that careful spectroscopic monitoring will reveal the same or a similar behavior in other sources. This has important conse- quences to the continuum production mechanism, as well as for the modeling of the warm X-ray gas around the center. We do not know the origin of the soft continuum source. In particular, we do not have information on its flux at wave- lengths l > 30 A (the source was not in a high-state in any of the published XMM-Newton observations). It may be re- lated to a broad-band phenomenon (e.g. due a flaring accre- tion disk), or a component covering a narrower band (e.g. a single temperature black body). Simple global energy consid- eration show that this phenomenon is unrelated to the appear- ance of broad emission features like those claimed to be seen in at least two sources and interpreted as due to relativistic disk lines (Branduardi et al 2001; Mason et al. 2002; but see also Lee et al. 2002). Finally, we must comment on the possibility of a more com- plex behavior. Our analysis is based on the fitting of a single powerlaw continuum at low-state and the addition of a soft X-ray component during the high-state. However, an equiv- alent analysis could be carried out in the reverse order -- i.e. starting from a pure powerlaw for the high-state and subtract- ing a continuum component to explain the low-state spectrum. However, as argued earlier, the observed spectral changes at long wavelengths are too large to be explained by pure opacity variations and it is not at all clear what other mechanism could explain an AGN continuum with flux deficit at long wave- lengths. 5. CONCLUSIONS Our detailed measurements and analysis of the 900 ks data set of NGC 3783 lead to the following results: 1. The source fluctuates in luminosity, by a factor ∼ 1.5, during individual 170 ks observations. The fluctuations are not associated with significant spectral variations. 2. On time scales of 20 -- 120 days, the source exhibits two very different spectral shapes denoted here as the high- and low-states. The two are associated with different softness ratios that seem unrelated to the total X-ray lu- minosity. The observed changes in the underlying con- tinuum can be described as due to the appearance (in the high-state) and disappearance (in the low-state) of a soft excess component. The origin of this continuum component is not clear. To the best of our knowledge, NGC 3783 is the first AGN to show such a behavior. 3. The appearance of the soft continuum component can explain all spectral variations observed within the mea- surement uncertainties. There is no need to invoke opacity changes between the high- and low-states. This 16 NETZER ET AL. conclusion depends mostly on the S/N in the high state spectrum. 4. A combination of three ionization components, each split into two kinematic components, provides a good description of the intensity of almost all the absorption lines and bound-free edges observed. The components span a large range of ionization, and have a total col- umn of about 4×1022 cm−2. The only real discrep- ancy between the observed and theoretical spectra are for the iron M-shell UTA feature at 16 -- 16.5 A. This is most likely due to inadequate dielectronic recombina- tion rates currently available to use in our calculations. The largest other uncertainty is in the column density of O VI. 5. The three generic absorbers discussed in this work have very similar values of ne × T . We speculate that the absorbers may be in pressure equilibrium with each other, occupying the same volume in the nucleus. This is the first confirmation of the location of several X- ray absorbers on the vertical part of the log(T ) vs. log(UOX /T ) stability curve of AGN. 6. We have obtained thee lower limits on the gas distance from the center, corresponding to our three generic ab- sorbers. The limits are 3.2 pc, 0.6 pc and about 0.2 for the low ionization, intermediate ionization and high ionization absorbers, respectively. The pressure equi- librium assumption implies distances in the range 3 -- 25 pc. This work is supported by the Israel Science Foundation grant 545/00, H.N. thanks S. Kahn and the astrophysics group at Columbia university for hospitality and support during a summer visit in 2002. E.B. was supported by the Yigal-Alon Fellowship and by the GIF Foundation under grant #2028- 1093.7/2001. We gratefully acknowledge the financial sup- port of CXC grant GO1-2103 (S. K., W. N. B., I. M. G.), NASA LTSA grant NAG 5-13035 (W. N. B.) and the Alfred P. Sloan Foundation (W. N. B.), REFERENCES Arav, N., Becker, R. H., Laurent-Muehleisen, S. A., Gregg, M. D., White, R. L., Brotherton, M. S., & de Kool, M. 1999, ApJ, 524, 566 Arav, N., Korista, K. T., & de Kool, M. 2002, ApJ, 566, 699 Barlow, T. A. & Sargent, W. L. W. 1997, AJ, 113, 136 Barlow, T. A., Hamann, F., & Sargent, W. L. W. 1997, ASP Conf. Ser. 128: Mass Ejection from Active Galactic Nuclei, 13 Behar, E., & Netzer, H. 2002, ApJ, 570, 165 Behar, E., Sako, M., & Kahn, S. M. 2001, ApJ, 563, 497 Behar, E. & Kahn, S. M. 2002, NASA Laboratory Astrophysics Workshop, Ed. Farid Salama (astro-ph/0210280) Behar et al., 2003 (astro-ph/0307467) Blustin, A. J., Branduardi-Raymont, G., Behar, E., Kaastra, J. S., Kahn, S. M., Page, M. J., Sako, M., & Steenbrugge, K. C. 2002, A&A, 392, 453 Branduardi-Raymont, G., Sako, M., Kahn, S. M., Brinkman, A. C., Kaastra, J. S., & Page, M. J. 2001, A&A, 365, L140 Chelouche, D. & Netzer, H. 2001, MNRAS, 326, 916 Chelouche, D. & Netzer, H. 2003, MNRAS, (in press) Chiang, J., Reynolds, C. S., Blaes, O. M., Nowak, M. A., Murray, N., Madejski, G., Marshall, H. L., & Magdziarz, P. 2000, ApJ, 528, 292 Gabel, J. R. et al. 2003a, ApJ, 583, 178 Gabel, J. R. et al. 2003b, ApJ, (in press) George, I. M., Turner, T. J., Mushotzky, R., Nandra, K., & Netzer, H. 1998, ApJ, 503, 174 Gu, M. F. 2003, ApJ, 589, 1085 Hess, C. J., Kahn, S. M., & Paerels, F. B. S. 1997, ApJ, 478, 94 Kaspi, S., Brandt, W. N., Netzer, H., Sambruna, R., Chartas, G., Garmire, G. P., & Nousek, J. A. 2000a, ApJ, 535, L17 Kaspi, S. et al. 2001, ApJ, 554, 216 Kaspi, S., et al. 2002, ApJ, 574, 643 (Paper I) Kinkhabwala, A. et al. 2002, ApJ, 575, 732 Komossa, S. & Fink, H. 1997, A&A, 322, 719 Komossa, S. & Meerschweinchen, J. 2000, A&A, 354, 411 Kraemer, S. B., Crenshaw, D. M., & Gabel, J. R. 2001, ApJ, 557, 30 Krolik, J. H., & Kriss, G. A. 2001, ApJ, 561, 684 Krongold, Y., Nicastro, F., Brickouse, N.S., Elvis, M., Liedahl, D.A., & Mathur, S., 2003, ApJ preprint doi:10.1086/378639 (K03) Lee, J. C., Ogle, P. M., Canizares, C. R., Marshall, H. L., Schulz, N. S., Morales, R., Fabian, A. C., & Iwasawa, K. 2001, ApJ, 554, L13 Magdziarz, P., Blaes, O. M., Zdziarski, A. A., Johnson, W. N., & Smith, D. A. 1998, MNRAS, 301, 179 Markowitz, A. & Edelson, R. 2001, ApJ, 547, 684 Marshall, F. E. et al. 1993, ApJ, 405, 168 Mason, K. O. et al. 2003, ApJ, 582, 95 Mathur, S., Elvis, M., & Wilkes, B. 1995, ApJ, 452, 230 Netzer, H. 1996, ApJ, 473, 781 Netzer, H., Chelouche, D., George, I. M., Turner, T. J., Crenshaw, D. M., Kraemer, S. B., & Nandra, K. 2002, ApJ, 571, 256 Pradhan, A. K. 2000, ApJ, 545, L165 Reynolds, C. S. & Fabian, A. C. 1995, MNRAS, 273, 1167 Savin, D. W. et al. 2002, ApJ, 576, 1098 Steenbrugge, K.C. et al. 2003, A&A, 402, 477
astro-ph/9604095
2
9604
1996-10-07T16:47:07
Halo White Dwarfs and the Hot Intergalactic Medium
[ "astro-ph" ]
We present a schematic model for the formation of baryonic galactic halos and hot gas in the Local Group and the intergalactic medium. We follow the dynamics, chemical evolution, heat flow and gas flows of a hierarchy of scales, including: protogalactic clouds, galactic halos, and the Local Group itself. Within this hierarchy, the Galaxy is built via mergers of protogalactic fragments. We find that early bursts of star formation lead to a large population of remnants (mostly white dwarfs), which would reside presently in the halo and contribute to the dark component observed in the microlensing experiments. The hot, metal-rich gas from early starbursts and merging evaporates from the clouds and is eventually incorporated into the intergalactic medium. The model thus suggests that most microlensing objects could be white dwarfs ($m \sim 0.5 \msol$), which comprise a significant fraction of the halo mass. Furthermore, the Local Group could have a component of metal-rich hot gas similar to, although less than, that observed in larger clusters. We discuss the known constraints on such a scenario and show that all local observations can be satisfied with present data in this model. The best-fit model has a halo that is 40% baryonic, with an upper limit of 77%.
astro-ph
astro-ph
HALO WHITE DWARFS AND THE HOT INTERGALACTIC MEDIUM Brian D. Fields and Grant J. Mathews Department of Physics, University of Notre Dame Notre Dame, IN 46635 [email protected] [email protected] and David N. Schramm University of Chicago, Chicago, IL 60637, and NASA/Fermilab Astrophysics Center, FNAL, Box 500, Batavia, IL, 60510 [email protected] ABSTRACT We present a schematic model for the formation of baryonic galactic halos and hot gas in the Local Group and the intergalactic medium. We follow the dynamics, chemical evolution, heat flow and gas flows of a hierarchy of scales, including: protogalactic clouds, galactic halos, and the Local Group itself. Within this hierarchy, the Galaxy is built via mergers of protogalactic fragments. Hot and cold gas components are distinguished, with star formation occurring in cold molecular cloud cores, while stellar winds, supernovae, and mergers convert cold gas into a hot intercloud medium. We find that early bursts of star formation lead to a large population of remnants (mostly white dwarfs), which would reside presently in the halo and contribute to the dark component observed in the microlensing experiments. The hot, metal-rich gas from early starbursts and merging evaporates from the clouds and is eventually incorporated into the intergalactic medium. The model thus suggests that most microlensing objects could be white dwarfs (m ∼ 0.5M⊙), which comprise a significant fraction of the halo mass. Furthermore, the Local Group could have a component of metal-rich hot gas similar to, although less than, that observed in larger clusters. We discuss the known constraints on such a scenario and show that all local observations can be satisfied with present data in this model. The most stringent constraint comes from the metallicity distribution in the halo. The best-fit model has a halo that is 40% baryonic, with an upper limit of 77%. Our model predicts that the hot intragroup gas has a total luminosity 1.5 × 1040 erg s−1, and a temperature of 0.26 keV, just at the margin of detectability. Improved X-ray data could provide a key constraint on any remnant component in the halo. Subject headings: dark matter -- galaxies: evolution -- galaxies: interactions -- nuclear reactions, nucleosynthesis: abundances cosmology -- dark matter, galactic evolution 1 1. Introduction Recently there has been renewed interest in the nature of the dark matter in galaxy halos, motivated by the results of microlensing experiments. Observa- tions toward the Magellanic clouds (e.g., Alcock et al. 1995; Aubourg et al. 1993) and toward the Galactic bulge (e.g., Udalski et al. 1993) have detected gravita- tional microlensing and inferred the presence of dark, massive compact halo objects (MACHOs). Further- more, a recent binary detection in the direction of the LMC, along with average event durations of about 2.5 months, may imply masses of order ∼ 0.5M⊙, sugges- tive of white dwarfs. At the same time, X-ray observations of clusters (Mushotzky 1993) and groups (e.g., Mulchaey, Davis, Mushotzky, & Burstein 1996, 1993; Pildis, Breg- man, & Evrard 1995; Ponman, Bourner, Ebeling, & Bohringer 1996) have discovered a large amount of hot, metal-rich gas. Where it is observed, this gas is a substantial fraction of the baryonic mass -- it is the dominant baryonic component of clusters, and is com- parable to the galactic component in groups. Indeed, it appears that most baryons in the universe are in the for of this hot X-ray emitting gas. The relatively high metallicity (Z ∼ 0.3Z⊙ for clus- ters, Z ∼ 0.1Z⊙ for groups) of this gas is impressive and demands that the material has undergone a sig- nificant amount of stellar processing during an ear- lier epoch of star formation. Such an epoch would also produce remnants, mostly white dwarfs. If this epoch is a general consequence of the formation of the bulge and halo of spirals like the Milky Way, as well as the ellipticals of rich clusters, it could account for the observed microlensing objects. Although white dwarfs are attractive MACHO candidates, there are important constraints on such objects and their formation (Ryu, Olive, & Silk 1990). These include background light from the early evolu- tion, the present luminosity of the halo, and the metal and helium content of the disk and halo stars. While these place important constraints on model parame- ters, they do not rule out a significant white dwarf halo population, as we shall show. In modeling these galaxy aggregates and their hot gas components, one must account for the dependence of these systems on the morphology of the constituent galaxies. On the one hand, the hot gas in clusters and groups appears correlated with the luminosity of el- liptical and S0 galaxies (Arnauld et al. 1992). This suggests that the hot gas arises from the violent merg- ing associated with these morphological types. On the other hand, it also seems well established (e.g., Rich 1990) that the morphology of the bulge and halo of spiral galaxies is quite similar to that of ellipticals and S0's. This suggests that the bulge and halo of spirals may have experienced a similar epoch of star formation and outflow during their formation. The difference in the morphologies may relate to the larger angular momentum or shallower gravitational poten- tial of spirals (Zurek et al. 1988), such that some gas survives halo formation, and settles afterwards into spiral arms. With this background in mind, we find that a likely, and perhaps inevitable, consequence of the forma- tion of the bulge and halo is the formation of a large remnant population in the Galactic halo, along with hot X-ray emitting gas in the Local Group and inter- galactic (i.e., extragroup) medium. While our model should be widely applicable, in this paper we concen- trate on the Local Group. Given the detection of dark microlensing objects, as well as the need for significant amounts of dark baryonic matter somewhere (Copi, Schramm, & Turner 1995; Fields, Kainulainen, Olive, & Thomas 1996) halo white dwarfs are a very conservative candi- date (e.g., Larson 1987; Ryu, Olive, & Silk 1990; Silk 1993). This is particularly so since red and brown dwarfs are apparently excluded as halo can- didates (Bahcall, Flynn, Gould, & Kirhakos 1994; Graff & Freese 1996). Here we make a specific though schematic model, and assess the plausibility of the white dwarf hypothesis.1 As we will see, the model can be made to work but not without some assump- tions (e.g., one must alter the halo initial mass func- tion). In any case, the model is eminently testable, and perhaps has already been tested by X-ray ob- servations in other groups. Indeed, should one find this model and the halo white dwarf hypothesis un- tenable, then it follows that the dark baryons and the MACHOs must take an even more exotic form. 2. Local Group Properties We take the Local Group to have a mass ∼ (3 − 5) × 1012M⊙ (Fich & Tremaine 1991). Its luminous component is dominated by two galaxies, one of which 1Others have suggested that the microlensing objects might be remnants from an early Population III; see Fujimoto, Sugiyama, Iben, & Hollowell (1995). 2 (the Galaxy) has a visible mass ∼ 7 × 1010M⊙. The total mass of the Galaxy is uncertain and depends on the radius, i.e., Mtot ∼ 5 × 1011M⊙(Rhalo/50 kpc). This implies that the dark halo has a mass within 50 kpc of ∼ 4 × 1011M⊙, some or all of which will be in the form of microlensing objects and other dark baryons. We consider models with up to 90% the total mass in nonbaryonic dark matter, at the group scale. The baryonic fraction at the halo scale, however, can be much less. In many galaxy groups, hot intergalactic gas is found (e.g., Mulchaey et al. 1993, 1996; Pildis, Breg- man, & Evrard 1995; Ponman, Bourner, Ebeling, & Bohringer 1996). This gas is a significant and some- times dominant component of the baryonic mass. The hot gas-to-galaxy mass ratio ranges from 0.3− 3. The gas also contains metals, with Z ∼ (0.1 − 0.2)Z⊙. While the ROSAT metallicity determinations are un- certain (Davis, Mulchaey, Mushotzky, & Burstein 1996), recent measurements with ASCA (Fukazawa et al. 1996) suggest that the gas metallicities in groups could have a larger spread than those of clusters. In any case, the metallicity is clearly not primordial, in- dicting that a significant fraction of material has been processed in stars before they are incorporated into the intragroup medium. If the gas is in hydrostatic equilibrium with the gravitational potential of the group, then the tem- perature distribution determines the total mass. Fur- thermore, the observed temperatures tend to tightly cluster around T = 1 keV for most groups observed thus far (Mulchaey et al. 1996). This implies that the total masses are similar, with most observed group masses lying in the range (1.5 − 2.5) × 1013M⊙. This is a tighter range than the span of luminous mass in galaxies and gas. However, it is not clear that all optically identified groups evidence hot gas. Several analyses of ROSAT observations have found that the presence of de- tectable gas is strongly correlated with the morpholo- gies of the constituent galaxies (e.g., Pildis, Bregman, & Evrard 1995). The trend is very similar to that found in galaxy clusters: the hot gas mass is cor- related with the presence of early-type (E and S0) Indeed, Mulchaey et al. (1996) emphasize galaxies. that there is at least one bright (LB >∼ 5 × 1010L⊙) elliptical galaxy in every group for which hot gas has been detected. On the other hand, Ponman, Bourner, Ebeling, & Bohringer (1996), also using ROSAT data, have re- 3 cently claimed the positive detection of hot gas in spiral-dominated Hickson compact groups. They find that the gas in these groups has a lower tempera- ture (∼ 0.3 keV), and thus, a lower surface brightness than that in groups with early-type galaxies. The dis- crepancies between these results and those of previous groups attests to the difficulty in trying to measure or put limits on such a relatively dim diffuse com- ponent. Ponman et al. (1996) note that differences between their results and others trace to details of the analysis, e.g., subtraction of galactic and back- ground emission. As Davis, Mulchaey, Mushotzky, & Burstein (1996) point out, these are not always straightforward issues. Hence, we regard the issue of hot gas in spiral-dominated galaxies as presently am- biguous. In the Local Group specifically, there has not been direct observation of hot intergalactic gas. However, Suto et al. (1996) have argued that such a component is allowed within current direct limits. Indeed, they suggest that this is a source of the excess low-energy component in the diffuse X-ray background. Suto et al. (1996) model a gas distribution that could lead to the excess radiation; their distribution implies a total mass in hot intergalactic gas of about ∼ 3.5×1011M⊙. Of course, the existence of hot gas in spiral-dominated groups is a central premise in this scenario. Pildis & McGaugh (1996) show that the observational limits on such gas (if it is similar to other spiral-rich groups) require any Local Group gas to have too low a mass to provide the soft X-ray background. 3. Hierarchical Collapse Model Motivated by hierarchical clustering scenarios for structure formation, we compute the evolution of a hierarchy of self-similar mass scales. The three scales are (1) protogalactic clouds, all of which reside in (2) galaxy halos, themselves moving within the (3) group. Within a spiral protogalaxy, the clouds merge to be- come ultimately the disk and bulge. However, we do not distinguish the disk and bulge. Their formation does not significantly affect the halo or group evolu- tion once the last merging has occurred and the star formation (and gas outflow) in the halo has dimin- ished. The dynamics of the three components are de- scribed (Mathews & Schramm 1993) as the radial evo- lution R(t) of a spherical overdensity, from its initial expansion (starting at t = 108 yr) through the de- parture from Hubble flow and collapse to a fixed final radius. The halo collapse is halted at 50 kpc. For each component in the structure hierarchy, we follow the evolution of matter in the form of gas, stars and remnants, and possibly also non-baryonic dark matter. We also compute the helium and metal evo- lution, and follow the temperature of the hot gas. Our model is schematic but contains the necessary features for testing and constraining our basic hypotheses. Its structure (summarized pictorially in figure 1) is as follows. 3.1. Level 1: Protogalactic Clouds In this simple model we assemble galaxies from a distribution of protogalactic fragments within ex- panding and contracting galactic halos. At the level of the protogalactic clouds, we consider the system to be composed of three components: cold star-forming molecular clouds; heated ejecta from the clouds; and stars and remnants. We assume that non-baryonic dark matter (if there is any) is unimportant at this scale. Let us first consider the cold star-forming gas and hot gas components. These components are most dramatically affected by star formation and merg- ing. Stars form from the cold gas and return a frac- tion of their material as hot ejecta. This hot stellar ejecta further mixes with and heats the local cold gas mass into the hot component. In addition, material is heated during mergers (e.g., White et al. 1993) as the relative kinetic energy of the merging clouds is converted into internal energy of the merged system. Also, cooling of the hot gas component returns mate- rial back to cold star forming regions. We write the coupled equations for the evolution of the cold and hot gas components for an average cloud experiencing all of these processes as: mcold g1 = −(1 + Rηmix)ψ(t) +λmergemcold g1 +λcoolmhot g1 , mhot g1 = R(1 + ηmix)ψ(t) (1) (2) +λmergemhot g1 −λcoolmhot + min1 − mhot g1 out1 . The first term on the right hand side of equations (1) and (2) describes the formation of stars and ejecta. R is the usual (e.g., Tinsley 1980) returned fraction of material from stars in the instantaneous recycling approximation. It depends on the initial mass func- tion. The hot ejecta is assumed to sweep up and mix with the cold cloud material; the resulting mixture contributes to the hot gas component. We therefore include an additional factor of ηmix describes the num- ber of equal masses of local interstellar material that is heated into the hot component along with the stel- lar ejecta. The rate of incorporation of cold gas into new stars is ψ(t). The second term on the right hand side in equa- tions (1) and (2) describes the effects of mergers. Here, λmerge is the merger rate per cloud. On average, during each merger, the cloud mass will double. The third term describes the cooling from hot to cold gas, with the cooling rate per particle given by λcool. The last terms in Eq. (2) accounts for infall and outflow of gas. The infall term min gives the rate at which hot halo gas is incorporated int to the hot gas component of the clouds. The outflow rate mhot out1 we attribute to the evaporation of hot gas. The physics behind each of these terms is discussed in §3.4. to be truly cold gas with T <∼ 100 K. The hot gas com- For our purposes, we take the cold star forming gas ponent is approximated by assuming that all gas at higher temperatures is isothermal and homogeneous within each of the levels of structure. If we combine equations (1) and (2) we can recover the familiar instantaneous recycling equation for one gas component; mcold g1 + mhot g1 = −(1 − R)ψ(t) + meff in − mout , (3) meff in = (mcold where the effective infall rate, due to mergers as well as the influx of halo gas, is g1 + mhot g1 )λmerge + min1. The outflow rate is just mout = mhot out1. As in MS93 we assume that protogalactic mergers disperse the stars and remnants into the halo. The evolution equation for the mass mr1 in stars and rem- nants remaining in a cloud becomes: mr1 = (1 − R)ψ(t) + (1 − κ)λmergemr1 . (4) Here κ is a measure of the efficiency for the mergers to disperse stars and remnants. Specifically, the average fraction f of stars and remnants born at time t0 which survive merging up to time t is f = e −κR t t0 dt′λmerge (5) 4 We will take κ = 1. Since stars return all of their ejecta into the hot component, the metallicity in the cold star forming gas is only indirectly enriched by cooling from the hot component. Thus, from Eq. (1) we write for the evolution of the total mass in metals in the cold star forming gas (Z cold 1 mcold g1 ):2 1 mcold g1 ) d(Z cold dt = Z hot 1 λcoolmhot g1 (cid:20)(1 + Rηmix)ψ(t) 1 −Z cold +mcold g1 λmerge(cid:21) . (6) This reduces to a simple evolution equation for the metallicity of the cold gas, Z cold 1 = (cid:0)Z hot 1 − Z cold 1 mhot g1 mcold g1 (cid:1) λcool . (7) Clearly, the equilibrium metallicity of the cold star- forming component is equal to the hot component metallicity. This equilibrium is, however, only achieved after a time given by the cooling time scale, λcool times the ratio of cold to hot gas masses. −1 By an analogous process to the derivation of Eq. (7), the evolution of metallicity in the hot gas com- ponent can be written Z hot 1 = (cid:2)yZ + (1 + ηmix)R(Z cold 1 − Z hot 1 +(Z hot g2 − Z hot g1 ) min1 mhot g1 ψ(t) mhot g1 )(cid:3) (8) mhot out mhot g1 −(cid:0)ǫout 1 − ǫ1(cid:1) + (ǫ2 − ǫ1) min1 mhot g1 R1 R1 −2 ǫ1 , where ǫSN in the average energy per unit mass in all stellar ejecta averaged over an appropriate initial mass function. This term is dominated by supernovae whose energy release is Eej ∼ 1051 erg per supernova. Averaging over the initial mass function φ(m) then gives ǫSN = R dm φ(m) Eej(m) R dm φ(m) mej (10) The quantity ǫout which is independent of the normalization convention of φ. For a typical initial mass function, this gives values of order ǫSN ∼ 1049 erg M−1 ⊙ . is the average energy per unit mass for the material exiting the cloud. For an ideal gas, the temperature is simply related to ǫ1 by, T1 = (2ǫ1µ/3kNA), where µ is the mean molecular weight of the gas and NA is Avagadro's number. 1 The last term in Eq. (9) accounts for the p dV work done as the cloud expands or contracts; the ra- dial dependence is given by the collapse of a spherical overdensity. The cloud radius R1 is the average tidal radius R1 = (cid:18) m1 m2(cid:19)1/3 R2 , (11) where m1 is the total average cloud mass m1 = mcold g1 + mhot g1 + mr1 , (12) where yZ is the mass fraction of newly synthesized material in the ejecta. As usual, yZ depends on both the initial mass function, and on the stellar nucleosyn- thesis yields as a function of mass (and metallicity). By a similar derivation, the temperature of hot gas in the halo is determined from an energy bal- ance equation. This leads to an evolution of internal energy per unit mass ǫ1 in the hot gas, ǫ1 = (cid:20)ǫSN − ǫ1(1 + ηmix)(cid:21)Rψ(t) mhot g1 (9) +λmergeǫmerge and m2 is the average halo mass defined below (Eq. 13). We assume that the relative momentum of the merger goes into heating the cloud gas. Thus, we have ǫmerge = Emerge/Mmerge = v2 rel/2. 3.2. Level 2: Galactic Halos For our purposes, galactic halos are treated as a ho- mogeneous assembly of protogalactic clouds and hot gas, possibly having a component of nonbaryonic dark matter. The total mass of the galactic halo is then m2 = ncm1 + mg2 + mr2 + mNB 2 , (13) 2We explicitly show only the metal evolution here and below, but from this the expressions for the helium evolution follows trivially. where nc is the number of protogalactic clouds, mg2 is the mass of hot gas which has exited the clouds to 5 reside in the halo, and mr2 is the mass of stars and remnants which have been dispersed from clouds into the halo. The mass of the nonbaryonic component, if present, is mNB 2 . The equation describing the evolution of the (hot) gas component in the halo is: mg2 = nc ( mout1 − min1) − mout2 + min2 , (14) where mout2 is the rate at which the hot gas is ejected from the halos as described below and min3 is the possible inflow of gas from the intragroup medium. The rate at which stars and remnants are injected into the halo from mergers can be inferred from Eq. (4), mr2 = κncmr1λmerge . (15) The equation governing the evolution of metallicity in the halos will be Z2 = (Z hot 1 − Z2) +(Z3 − Z2) nc mout1 mg2 min2 mg2 , (16) where nc is the number of cold protogalactic clouds in the halo. Similarly the internal energy per unit mass, ǫ2, of gas in the halo is determined from an energy balance equation: ǫ2 = (ǫout (17) nc mhot out1 mg2 1 − ǫ2) 2 − ǫ2(cid:19) mout2 −(cid:18)ǫout + (ǫ3 − ǫ2) mg2 −λcoolǫ2 − 2 ǫ2 . min2 mg2 R2 R2 The number of clouds decreases exponentially with the number of mergers: nc = −λmergenc . (18) 3.3. Level 3: Hot Intra-Group Medium The equations governing the evolution of the intra- group medium are analogous to those for the galactic halos, but at the level of the group nonbaryonic dark 6 matter may be a dominant contributor. Thus, for the total group mass we write, m3 = nhm2 + mg3 + mNB 3 , (19) where nh is the number of galactic halos in the group, mg3 is the mass of the hot X-ray intragroup medium, and mNB is the contribution from nonbaryonic dark 3 matter. The evolution equation for the intragroup medium is then, mg3 = nh( m2out − min2) − m3out , (20) and m3out is the rate at which the gas is lost from the group. The equation governing the evolution of metallicity in the intragroup medium is then Z3 = (Z2 − Z3) nh m2out mg3 , and the energy balance equation is ǫ3 = (ǫout nh mout2 mg3 mout3 mg3 2 − ǫ3) −(cid:0)ǫout 3 − ǫ3(cid:1) R3 −λcoolǫ3 − 2 R3 ǫ3 , (21) (22) from which temperature of the intragroup medium can be inferred. 3.4. Cooling, Star Formation, Mergers, and Mass Loss The evolution is given by equations (1 -- 22). What remains is to specify the various input quantities, i.e., the rates for cooling, star formation, merging, and mass loss. The metallicity-dependent cooling rate λcool is de- rived from the calculations of Bohringer & Hensler (1989). For the temperatures appropriate for the hot gas, the cooling is dominated by brehmsstrahlung emission from electrons. These losses are written in terms of the cooling function Λ, which gives the en- ergy loss rate per unit density. Given the number density at a given scale ni, one may then compute Ei = niΛ, and then the energy loss rate per particle: the cooling rate λcool ≡ E/E. We allow star formation to be induced both by mergers and intrinsic star formation processes within the clouds (MS93). In MS93 it was shown that in- trinsic star formation is more important during the subsequent evolution as material settles into a disk. To describe the merger-induced star formation for all of the merging substructures within a collapsing halo, MS93 proposed a schematic model of colliding viri- alized protogalactic clouds within an expanding and contracting halo. We adopt that formulation here. Specifically, we presume that the fragments virial- ize as they form. The collision rate per cloud within this virialized velocity distribution can be written, λmerge = (nc − 1)σv/V , (23) where n is the number of protogalactic clouds within a volume V , σ is an average collision cross section, and v is the virial velocity (v ∼ [0.4GMh/Rh]1/2), where Mh is the total gravitational mass of a galactic halo, and Rh is the radius of the galactic halos, approximated as a collapsing spherical overdensities. We define halos as those regions which evolve to become independent gravitationally bound ensembles of gas and stars at the present time. All halos themselves are viewed as the result of merging internal structure. Hence we only describe merging within the halos and not merging between halos. This is largely a matter of semantics. The number of protogalactic clouds decreases ex- ponentially with the integral of the merger rate (Eq. 18), nc(t) = nc(0) e−R t 0 λmergedt′ , (24) where the initial number of clouds, nc(0), is given by the ratio of the total initial halo baryonic mass to the initial cloud mass. Thus, nc(0) ∼ 106 for this schematic model. For the merger cross section σ we use σ = πR2 t , with a tidal radius Rt = 2R1 to include gravitational effects (Binney & Tremaine 1987). Near the end of the collapse of the halos, the com- bined effects of conservation of angular momentum and heating will dissipate the radial motion. In our schematic model, we approximate these effects by halting the halo collapse at a size of a typical present dark-matter halo, (R2 ∼ 50 kpc). The cloud radii are fixed by Eq. (11), and the group collapse is halted when the radius reaches the present size of the Local Group, R3 ∼ 700 kpc. The merger-induced stellar birth rate thus is taken as proportional to the mass of gas mass participating in mergers per unit time. The intrinsic, quiescent star formation rate is taken as ∝ ρn gas1, where coldmcold n = 1/2 in our case. The total star formation rate is the sum of the two terms: ψ(t) = (cid:16)αmrgλmerge + βQρ1/2 cold(cid:17) mcold g1 , (25) Note that we have parameterized the strength of merger-induced star formations in terms of the di- mensionless efficiency αmrg. Typical numbers for αmrg are ∼ 1% (MS93). The coefficient of the qui- escent piece is taken as βQ = βqΛ0, where Λ0 = 1.7 × 10−4/(M⊙ kpc−3)−1/2 Gyr−1. Typical values of the dimensionless scaling βq are 1 − 5 (MS93). Finally, we must specify the rate at which hot gas is ejected from, or falls into, the clouds, halos, and the group. We first calculate the outflow due to evap- orative mass loss. We assume that the hot gas is dis- tributed homogeneously within an object. The rate of loss of hot gas is then just given by the fraction of a Maxwellian thermal distribution of velocities in excess of the escape velocity at the surface of the structures at all three levels. Thus, we write, gi = 3ζloss hv(> ve)i mhot Ri mhot gi ≡ λoutmhot gi (26) Where, i denotes clouds, halos, or the group, and ζloss ≤ 1 is a dimensionless scale factor depending on the geometry of the cloud; ζloss = 1 for a sphere. The factor hv(> vesc)i is the average velocity of all particles above the escape velocity v2 esc = 2GMi/Ri: hv(> vesc)i = R ∞ dv v3 fMB(v) vesc R ∞ 0 dv v2 fMB(v) 1 4√π (1 + xi)e−xi vT = (27) where fMB is the Maxwellian distribution. The aver- age thermal velocity is v2 T = 2kTi/µmp, and µmp is the average mass of a gas particle. xi = GMi/RikTi is the ratio of gravitational binding energy to thermal energy, and is small typically. Similarly, the average energy loss per unit ejected mass of material is ǫout i = hǫv(> vesc)i hv(> vesc)i 1 + xi + x2 i /2 = 1 + xi 2kTi µmp (28) which for small xi gives ǫout i ≃ 2kTi/µmp = 4ǫi/3. 7 Using similar derivations, we have computed in- flow of halo (group) gas into the clouds (halos). We note that this inflow has two components. First, some material is accreted simply due to the motion of the halo or cloud through the surrounding medium -- the finite size of these substructures will lead to the cap- ture gas that falls within their geometric cross section σ = πR2 i . Second, the hot gas particles surround- ing the substructures will accrete due to their ran- dom thermal motion. We assume that these particles quickly equilibrate with the ambient gas. Together these two effects lead to an infall term of the form mhot ini = 4 3 ζloss (cid:18) Ri Ri+1(cid:19) mhot gi+1(cid:0)vT i+1 + vc(cid:1) Ri+1 , (29) with v2 locity. c ≡ Gmi+1/Ri+1 the cloud (halo) circular ve- 3.5. Initial Mass Function and Stellar Yields The initial mass function (IMF) is important for our scenario in several respects. As we have noted, it affects the parameters of our model, namely the returned fraction R, yields yY and yZ, and specific energy injection ǫSN. Moreover, the crucial impact of the IMF for our scenario is that it determines the ratio of low mass stars to high mass stars in the halo. Low mass stars (m <∼ 0.9M⊙ for halo metallicities) will still be burning today. While these stars are faint, they would be detectable, and heavily favored by an IMF similar to that of present disk stars. As a re- sult, such stars would today outnumber by far the halo remnants, and their net contribution to the halo luminosity would be large (Ryu, Olive, & Silk 1990). Thus, remnants cannot be a significant component of the halo if the halo IMF is similar to that inferred for disk stars. To allow for a significant population of halo white dwarfs, they must have been formed from an IMF which strongly favored the formation of intermediate- to high-mass stars over low-mass stars (Ryu, Olive, & Silk 1990). Given the different physical conditions during the formation of the halo (e.g., frequent mergers, higher temperatures, lower metallicities), it is at least plau- sible that the star forming process then was different than that of disk stars. Indeed, there are arguments for an early IMF that is skewed towards higher masses (see Silk 1993; Bond, Arnett, & Carr 1984; Adams & Fatuzzo 1996). Therefore, we have chosen the simple log-normal IMF parameterization, as suggested by both obser- 8 vational (Miller & Scalo 1979) and theoretical ar- guments (Adams & Fatuzzo 1996 and refs therein): ln φ(ln m) = ln φ0 − [ln2(m/mc)/2σ2]. We investigate the effect of variations of the centroid mass mc and the dimensionless width σ. We take the IMF bounds to be (0.1,100)M⊙, with a black hole cutoff at 18M⊙(Brown & Bethe 1995). In fact, our results for metal yields are insensitive to these limits. With these parameters, we obtain a returned fraction R = 0.375. The stellar yields are taken from Maeder (1992). We follow both helium and metallicity Z. These abundances are particularly powerful constraints when used together. The helium yields are mainly from in- termediate mass stars (m <∼ 8M⊙), while the metal yields come primarily from high mass stars. Thus, the IMF must strike a balance between the two mass ranges to avoid an inappropriate ratio of helium to metals. With our IMF, we find yY = 0.0139 and yZ = 0.0354. 4. Results In this hierarchical merging picture we take the clouds to be the initial building blocks of all structure. Thus, while we assume the clouds to be comprised initially of hot and cold gas, we assume baryons in the halos and the Group to be initially within clouds only. Thus, by definition the initial gas in the halos and group is zero. The initial metallicity is zero, while we take a primordial helium abundance of Y = 0.235. We have run the model for plausible ranges of the input parameters. The two models we present here are: (1) a "best-fit" model which has appropriate fi- nal masses and metallicities, while also optimally sat- isfying other constraints (§§4.4 -- 4.6); and (2) a "max- imum remnant" model which has the highest possible halo remnant mass without violating the constraints. Parameters of these models are summarized in Table 1. Unless otherwise noted, numerical results will be given for the best-fit model. For all models, the qualitative results are as fol- lows. There is generally a high initial merger rate, which lasts for <∼ 1 Gyr. This reduces the number of protogalactic clouds from 106 to about 100 while producing a burst of star (and hot gas) formation. The star formation rate and cloud number evolution appear in Figure 2. Some of the stars are ejected into the halo by the mergers, and on a longer timescale, a wind is ejected -- first from clouds, then from halo, and ultimately from the Group. When the halo col- lapses, after 5 Gyr (Mathews & Schramm 1993), there is an additional burst of merging and star formation. The remaining clouds coalesce into what will eventu- ally become the galactic bulge and spiral arms. With the collapse there is also heating and a reinvigorated wind. Subsequently, all remaining (hot, metal-rich) halo gas is ejected into the intergalactic medium. The evolution of halo and group masses is given in Figure 3. 4.1. Disk and Bulge Formation The clouds begin with the Jeans mass at recom- bination, 106M⊙. They merge to form the proto- disk+bulge, with a mass 8.1×1010M⊙. It is encourag- ing that, at the end of the halo collapse, most material in the proto-disk is in cold gas which will form disk stars. The evolution of metallicities is given in Figure 4. As seen in the figure, the metallicities grow rapidly in the initial burst, then remain fairly constant un- til the halo collapse at 5 Gyr. Then the metals rise again as the halo collapses and the disk and bulge are formed. Hence, the halo evolution provides an initial metallicity for the gas of the protodisk. This corre- sponds to an "initial enrichment" of the disk material, and so avoids overproduction of metal poor disk stars (i.e., the disk G-dwarf problem). 4.2. Halo Formation In the best-fit model, portrayed in Fig. 3a, the galaxy (=clouds+halo) begins with a mass of 1.35 × 1012M⊙, of which 1.1 × 1012M⊙ is baryonic (all in the clouds). Even higher fractions are allowed, as discussed below (§5). The final galaxy mass is 5.0×1011M⊙, of which 2.5×1011M⊙ is baryonic; thus 50% of the total Galactic mass is in baryons. Some of these baryons constitute the disk+bulge, as described above. The baryonic mass of the halo remnants alone (excluding the disk+bulge) is 1.7 × 1011M⊙, and the halo nonbaryonic mass accounts for 2.5 × 1011M⊙. Hence, 40% of the dark halo is in remnants, consis- tent with the microlensing observations. When combined with mass loss from the other galaxy in the Local Group, there is a total loss of 1.7 × 1012M⊙ of gas from both halos into the intra- group medium. The galactic wind is thus very ef- ficient, as it must be to remove the ejecta that ac- companies the remnant production. The remnants themselves are mostly white dwarfs, with 12% neu- tron stars. 9 4.3. Group Evolution In the best-fit model, the Local Group begins (Fig. 3b)with a mass of 5.6× 1012M⊙, and a baryonic mass of 2.2 × 1012M⊙. At the end of the simulation the group has a total mass 4.3× 1012M⊙, 17% of which is baryonic. Of the baryonic group mass, 2.9 × 1011M⊙ resides in hot gas. The intragroup gas temperature is 0.26 keV, just at the limit of ROSAT sensitivity. The group as a whole loses 1.4 × 1012M⊙ of gas to the intergalactic medium; thus about 64% of the ini- tial baryonic mass in the group is ejected later into intergalactic space. It is remarkable that the remnants in the halos, turn out to represent a small fraction of the initial baryonic matter. The galactic winds required to re- move the stellar ejecta co-produced with the white dwarfs prove strong enough to remove most gas from the group itself. This amount of hot (ionized) ma- terial is consistent with Gunn-Peterson limits on the intergalactic medium if it does not cool further (§5.1). Thus, in the best-fit model we find the dark halo to be 40% microlensing objects. These take the form of stellar remnants, 88% of which are white dwarfs, and the rest neutron stars (and perhaps black holes). In this model the copious production of hot intra- group and intergalactic gas is a natural consequence of white dwarf-dominated halos. We produce a present mass of intragroup gas of 2.9 × 1011M⊙. This cor- responds to about 37% of the baryonic mass of the Local Group. This mass is encouragingly near the value (3.5×1011M⊙) implied by the Suto et al. (1996) model for the excess diffuse X-ray background (§2). (But recall the observational controversy §2.) The X-ray luminosity of the halo is 1.5×1040 erg s−1. Let us assume for simplicity that the earth is at the center of this emission, and that the gas extends ho- mogeneously to the edge of the Local group (R3 = 700 kpc in this model). This leads to a total diffuse back- ground of about 4× 10−2 erg s−1 cm−2 sr−1. This is, at its peak energy, about 3 orders of magnitude be- low the observed diffuse soft X-ray background. We therefore find the diffuse Local Group radiation to be negligible. The contribution is thus much lower than that suggested by Suto et al. (1996), who postulate a similar (but larger) intragroup gas mass, but posit a temperature of 1 keV as opposed to 0.26 keV in this model. 4.4. Mass budget The most basic constraint on our model is the re- quirement that it reproduce the observed mass and metal budget of the Local Group. The final masses determine the initial masses, but the metallicity val- ues constrain how to get these. Consider the to- tal metal mass production in our model. The to- tal initial baryonic mass is 2.2 × 1012. Of this, the mass processed into stars and remnants is Mrem = 4.95 × 1011M⊙. The total mass into ejecta is just R/(1 − R)/Mrem = 3.0 × 1011M⊙. The total metal mass produced is yZ Mej = 1.1 × 1010M⊙, and the total new helium mass is yY Mej = 2.8 × 1010M⊙. Now, the average new metals from a given star rep- resents a high fraction of that star's ejecta: yZ = 0.014 ∼ 2/3 Z⊙. However, this metal-rich ejecta is diluted in several ways. First, it is mixed with η times its mass as it leaves the cold star-forming re- gions. Then it is mixed into the hot cloud gas, which quickly evaporates into the halo and on to the intra- group medium. Much gas is even evaporated from the Local Group itself, becoming part of an intergalactic (extragroup) medium. Thus, the final mean metallic- ity must be averaged over all of the baryons (including the ejecta from the Local Groups). One finds that the global average mass fraction is just 0.0049 ≃ 0.25Z⊙, an acceptably small value. The key point here is that star formation does not occur in all regions containing gas (as is often as- sumed in simple chemical evolution models) but only cold molecular cloud cores. Consequently, the stars represent a small fraction of the total baryonic mass. Furthermore, whereas the stars remain in the galax- ies, the gas is dispersed at all levels. The halo can thus contain many remnants but not a large amount of metals. 4.5. Nucleosynthesis Nucleosynthesis provides an important additional constraint on this scenario (Ryu, Olive, & Silk 1990; Charlot & Silk 1995; Hegyi & Olive 1986). The burst of star formation which produces the white dwarfs must not overproduce metals and helium. As noted above, the metallicities at the different scales are rea- sonable. Furthermore, we may test not only the aver- age metallicities, but also the halo metallicity distri- bution. Figure 5 compares our calculated halo stel- lar metallicity distribution with the observed globu- lar cluster distribution. The agreement is good for the most metal-poor ([Z] < −1), and thus the oldest, portion of the population. The halo metallicity distribution strongly constrains the total halo remnant mass in our model. The posi- tion of the peak constrains the net star formation rate amplitude. If the star formation rate is too large, the peak is shifted too high. Thus, the need to reproduce the peak at [Z] ∼ −1.5 limits the mass processed into stars, and so leaves fewer stars available for the halo. From the χ2 of a fit to the halo metallicity distribution we estimate a 2σ upper limit of a 77% remnant contribution to the halo. Consequently, it seems likely that the halo contains at least some non- baryonic dark matter. In addition, the shape of the distribution is largely sensitive to the relative star for- mation rate, and to the degree of mixing between the stellar ejecta with the cold gas. The IMF shape is of course important in deter- mining the metallicities. However, contrary to Ryu, Olive, & Silk, we find no significant constraint on the IMF upper limit, essentially because the ejection of gas removes some of the metals produced by high- mass stars. This is even more so if there is a cutoff for black hole formation as low as 18 M⊙ (Bethe & Brown 1995). This is 4.6. Halo Luminosity Several sources of luminosity also provide impor- tant constraints. (1) Low mass stars in the halo are long lived and can lead to an unacceptably high halo mass-to-light ratio (Ryu, Olive & Silk 1990; Richstone, Gould, Guhathakurta, & Flynn 1992); however, in our model, the mass to light ratio is comfortably above these au- thors limit M/LB > 500 for Rhalo = 50 kpc. We determine the luminosity contributed by all stars still burning in the halo by integration of stellar luminosi- ties, the past star formation rate, and initial mass function, without the instantaneous recycling approx- imation. We find a mass to total luminosity ratio of M/Ltot = 760, M/LV = 1300, and M/LB = 3500. (2) Light from the bursts of star formation can lead to a large diffuse background (Charlot & Silk 1995; Zepf & Silk 1996). Using a population synthe- sis model we have preliminarily computed this back- ground assuming the starlight to be unscattered after its emission. These will be discussed in Mathews et al. (1996), but results are consistent with observed con- straints. We exceed the Charlot & Silk (1995) limit 10 of a 10% remnant halo because most of the elements are formed at very large redshift (z ∼ 10). (3) The halo white dwarfs will contribute to the white dwarf luminosity function (e.g., Adams & Laugh- lin 1996; Chabrier, Segretain, & M´era 1996). There is a new constraint from Flynn, Gould, & Bahcall (1996), which requires halo white dwarfs to have an luminosity MV >∼ 18.4. for V − I > 1.8. The white dwarf luminosity function in our model is given in Figure 6. The two peaks corresponding to the two bursts of star formation. The strongest con- straint comes from the more recent peak produced at the epoch of halo collapse. These white dwarfs have cooled to present luminosities of ∼ 10−5.5L⊙. This would correspond to a bolometric magnitude of 18.4. However, the bolometric correction (Liebert, Dahn, & Monet 1988), would probably reduce the visible lu- minosity to MV > 20, well below the Flynn, Gould, & Bahcall limit (similar to the findings of Kawaler 1996). The low luminosity of the white dwarfs de- rives in part from their very old age, but also because the population comes from a shifted IMF. Since the progenitors are more massive stars, they died sooner and so the remnants have had longer to cool. Indeed, our adopted IMF centroid of 2.3M⊙ is that of Adams & Laughlin (1996), who chose it specifically to obey the luminosity function constraints in the disk. 4.7. Parameter Sensitivity Our results depend sensitively on several of the model parameters. Parameter values for our best-fit model are summarized in column 3 of Table 1. Among the most important parameters are those that deter- mine the shape of the IMF. We find a good fit for a centroid (c.f. §3.5) mc = 2.3M⊙, and a width σ = 1. Note that this centroid is that of Adams & McLaugh- lin (1996). However, they advocated a tighter width to avoid metallicity problems which we do not require in our model because metals are efficiently ejected in the hot wind. Indeed, the width we use is essentially the present value (Miller & Scalo 1979). Other IMF parameters are allowed, but to obey the observational constraints, these must be near the ones we have adopted. As discussed above (§3.5), the cen- troid must be >∼ 1M⊙ to avoid significant production of long lived, low mass stars still burning in the halo. However, the centroid must not be too high to avoid untenably large metal yields. The width must also be not so large that it allows too many low mass stars, but not so narrow that metal yields are too small and helium yields too large. Thus, the need for a dark halo, and for a reasonable nucleosynthesis, drive the IMF parameters to the range of those we have chosen. The other important parameters involve star for- mation and ejecta. The star formation parameters (eq. 25 and Table 1) are reasonable and consistent with previous results (MS93), derived from requiring consistency among various comoschronometers. The quiescent term dominates, controlling the mass bud- get, but the merger term is also important. The ratio of the merger to quiescent contribution controls the position of the peak in the halo metallicity distribu- tion (Fig. 5). For the ejecta, the most important parameter is the degree of mixing ηmix. A higher value leads to more dilution of the ejecta, which leads to lower metallicity and temperature in the hot gas. The observed lower temperatures and metallicity of the hot gas in groups demands that ηmix > 1. 5. The Remnant Contribution to the Dark Halo Mass Given that (1) there is good evidence for dark mat- ter in the Galactic halo, and that (2) microlensing ob- jects are the first positively detected dark matter in the halo, a key question arises: how much of the dark halo is in MACHOs? In this context it is important to note that our best model does not give an all-remnant halo. Such a model is attractive for its simplicity, but we find that our best fit prefers only ∼ 40% of the dark halo mass in remnants, with the balance in non-baryonic material. The strongest constraint on the white dwarf halo fraction is the halo metal- licity distribution. Increasing the white dwarf frac- tion requires increasing the star formation rate; this leads to halo metallicities whose distribution peaks too high (§4.7). The "maximum remnant" model is that having the 2σ upper limit for the halo fraction (as set by the halo globular cluster metallicity distri- bution). It's parameters are summarized in Table 1; in it the halo remnant mass fraction is 77%. This up- per limit can be raised somewhat if we increase the stellar merger dispersal efficiency κ (see §3.2). While it is difficult to increase the halo remnant fraction, it is easy to decrease it. If the winds are less efficient (e.g., if ζloss < 1), then more material is recycled in the clouds. Thus a lower star formation amplitude is required and fewer stars are available to eject at early 11 Table 1: Model Parameters Adopted Model Feature Parameter Best-Fit Max Remnant IMF Star Formation Ejecta mc σ αmrg βq ηmix ζloss 2.3 M⊙ 1.6 0.007 1 8 1 same same 0.002 2.5 8 1 times. We find a minimum halo remnant fraction, for ζloss <∼ 0.08, of ∼ 1%. Our findings are to be compared to recent analysis of the MACHO data (Alcock et al. 1996). The mi- crolensing results are still very model-dependent, but it is provocative that they find a best fit for ∼ 50% of the halo being made up of 0.5+0.3 −0.2M⊙ objects. The question of whether one can put stronger constraints on the MACHO halo fraction in our model leads to the issues raised in the next section. 5.1. X-ray Groups: a Key Constraint As indicated in §2, there is currently large uncer- tainty regarding X-ray observations of spiral domi- nated groups. However, while the observations are ambiguous, what seems more clear is that spiral-rich groups do not contain hot gas with T ∼ 1 keV. That is, these systems do not evidence the same kind of diffuse emission as groups with early-type galaxies. If hot gas exists in these groups, it must have either low mass (<∼ (1−3)×1010M⊙), or a low temperature, near or below the ROSAT threshold Tth <∼ 0.3 keV. Indeed, gest that these systems do have cool (T <∼ 0.3 keV) Mulchaey, Mushotzky, Burstein, & Davis (1996) sug- gas, and argue that this may have been detected as high-ionization quasar absorption lines. Furthermore, given the morphological similarity of the bulge and halo with early-type galaxies, it would be surprising if no hot gas were found in these systems. At any rate, the observations seem to rule out that the gas is both massive and hot in spiral-rich groups. However, in our model the temperature and the gas mass are related, as follows. After the processing of one stellar generation, the ejecta to remnant ratio is R/(1 − R). Mixing with unprocessed cold material gives a total hot gas to remnant ratio (1 + ηmix)R/(1 − R). In our model, R = 0.38, so Mgas > Mrem for ηmix > 1.7 (which holds in our case). This already means that the mass 12 of ejected gas must be larger than the halo baryon mass (assuming that most hot gas escapes). Con- sequently, if the limits on gas in spiral-dominated groups apply -- i.e., if the gas were to have T >∼ 0.3 keV -- then the mass in halo white dwarfs must be less than the X-ray limits on the gas mass, ∼ 1010M⊙. This would be only a small component of the halo. To allow for a significant gas mass, therefore, one must demand that it avoid the ROSAT limits because it is cool: T <∼ 0.3 keV. However, the gas temperature is related to the gas-to-remnant ratio, as we now show. The total energy in the ejecta is RǫSNMrem. This implies that the energy per unit mass of ejected gas is ǫgas = RǫSNMrem/Mgas = ǫSN(1 − R)/(1 + ηmix). Finally, using the scaling TSN = (2/3) µmpǫSN = 2 keV, we have Tgas = 2 keV(1 − R)/(1 + ηmix). That is, the gas temperature diminishes with ηmix. But we already have seen that having a significant fraction of the halo in remnants, we need the concomitant halo gas to be cool. Using the Ponman et al. (1996) value Tgas ≃ 0.3 keV, and with R = 0.39, one finds ηmix ≃ 3. In our detailed model which includes reprocessing, we find a larger value, η = 8. Thus, cooler gas results if ηmix > 1, but only if the bulk of the baryons are in the gas. 5.2. Cosmological Baryon Fraction In the preceding section we see that our best-fit model finds that a large fraction of baryons, initially in the group, are ejected as an intergalactic medium. This can be reconciled with the total cosmological baryonic budget, as follows. In our scenario, the cos- mic baryonic inventory is ΩB = Ωdisk+bulge + Ωrem halo + Ωgas IGM (writing the uni- versal density in component i in units of the critical density: Ωi ≡ ρi/ρcrit, where ρcrit = 3H 2 0 /8πG). If most baryons reside in the intragroup and intergalac- tic diffuse gas, we therefore require Ωgas IGM ≃ ΩB ≫ Ωrem halo. IGM ≃ Ωrem LG + Ωgas LG + Ωgas halo + Ωgas LG + Ωgas Observationally, galactic rotation curves imply that galaxy halos have Ωhalo ∼ 0.02h(Rhalo/50 kpc), where Rhalo is the (unknown) radius of the dark halo (e.g., Peebles 1993). On the other hand, primordial nucle- osynthesis calculations give ΩB = (0.015 ± 0.005)h−2 (Copi, Schramm, & Turner 1995; Fields et al. 1996). Taking the ratio, we have Ωhalo/ΩB ∼ 1.3h3. Conse- quently, for h <∼ 0.9, Ωhalo < ΩB; that is, for reason- able values of h, some baryons are likely to be non- galactic. Indeed, for a low Hubble constant (h = 0.5), the bulk of the baryons are nongalactic. Furthermore, these relations assume that the halos are entirely baryonic (not the case in our best model). Any non- baryonic component in galaxy halos only strengthens the argument. Thus, we see that the X-ray observations of groups can provide strong constraints on remnants in the halo. Of course, this assumes that the Local Group is like other poor groups. If so, and if the X-ray data is reliable, then these data may provide a key constraint. As we have noted, if there is only a small X-ray gas mass in the Local Group, then there can be few rem- nants in the halo. Taking the more optimistic view, if the Ponman et al. (1996) result is correct, then there could be a large amount of (as yet unobserved) hot gas in the Local Group, as required in our model. 6. Conclusions We have shown that, without violating constraints posed by luminosity and nucleosynthesis considera- tions, one may construct a plausible model in which the dark halo of the Galaxy contains a significant fraction of white dwarfs. These may have already been detected in halo microlensing events towards the LMC, and might also be detected via their luminos- ity function. The same bursts of star formation which produced the white dwarfs also led to hot, metal-rich intergalactic gas, some of which may still reside in the Local Group. This hot gas could be detectable via its X-rays, and by distortions in the cosmic microwave background radiation (Suto et al. 1996). Thus, the predictions of the model are testable. If the halo is comprised of white dwarfs then there must be a background of hot, X-ray emitting gas in the Local Group. Conversely, if there is metal-rich hot gas in the Local Group, then a significant fraction of the halo mass must be in remnants. Clearly, further searches for both of these are warranted. Furthermore, if our galaxy formation scheme is in- deed universal, then hot gas production and ejection should be a ubiquitous aspect of halo formation. Con- sequently, X-ray observations of other systems could provide a key constraint on our model. In partic- ular, our model can be directly tested by observa- tions which can unambiguously confirm or deny the presence of hot gas in other spiral-dominated groups. Also, if white dwarfs are ubiquitous in galactic ha- los, then they may lead to detectable infrared pro- files in edge-on galaxies, which may already have been observed (Barnaby & Thronson 1994; Sackett, Morrison, Harding, & Boroson 1994; Lequeux, Fort, Dantel-Fort, Cuillandre, & Mellier 1996; Lehnert & Heckman 1996). Finally, even if our scenario turns out not to be applicable to spiral-dominated groups, it remains that ellipticals must eject gas in strong winds. Thus, our model may still be valid for clusters, which are elliptical-dominated. This will be explored in a subsequent work. We note as well that in our scenario, just as there is typically a large outflow from the halo, there is also a strong evaporative wind that ejects material from the Local Group. As a result, most baryons eventually reside in hot, intergalactic (as opposed to intragroup) gas. If this gas stays hot, it could perhaps be the ionized intergalactic (as suggested by Gunn-Peterson limits on the neutral intergalactic medium). If it does cool, it presents serious problems, as it would lead to prodigious but unobserved absorption of extragalactic radiation. Finally, we reiterate that stellar remnants and their associated hot ejecta are conservative candidates for both the halo microlensing objects and for the bary- onic dark matter. If these can be ruled out, then we are forced to conclude that the microlensing objects and the dark baryons are something stranger still. We are pleased to acknowledge useful discussions with C. Alcock, D. Bennett, E. Gates, K. Jedamzik, S. Shore, and M. Turner. We are grateful to R. Mushotzky and J. Mulchaey for may useful discus- sions of X-ray observations, and we especially thank R. Pildis for pointing out the limits on spiral-dominated groups. Finally, we are particularly indebted to St´ephane Charlot for advice and assistance on an early version of this work. Work at Notre Dame is supported by DoE Nuclear Theory grant number DE- FG02-95ER40934. Work at the University of Chicago is supported by NSF grant AST 90-22629, DOE grant DEF02-91-ER40606, and NASA grant 1231, and by 13 NASA through grant NAGW 2381 at Fermilab. Graff, D.S., & Freese, K.1996, ApJ, submitted (astro- REFERENCES Adams, F.C., & Fatuzzo, M. 1996, ApJ, 464, 256 Adams, F.C., & Laughlin, G. 1996, ApJ, submitted (astro-ph/9502006) ph/9602051) Hegyi, D.J. & Olive, K.A., 1986, ApJ, 303, 56 Kawaler, S. 1996, ApJ, in press (astro-ph/9606094) Larson, R.B. 1987, Co. Ap., 11, 273 Alcock, C. et al. 1995, ApJ, 461, 84 Lehnert, M.D., & Heckman, T.M. 1996, ApJ, 462, 651 Alcock, C. et al. 1995, ApJ submitted (astro- ph/9606165) Arnauld, M., Rothenflug, R., Boulade, O., Vigroux, L., & Vangioni-Flam, E. 1992, A&A, 254, 49 Aubourg, E. et al. 1993, Nature, 365 623 Bahcall, J.N., Flynn, C., Gould, A., Kirhakos, S. 1994, ApJ, 435, L51 Barnaby D., & Thronson, H.A. 1994, AJ, 107, 1717 Bethe, H.A., & Brown, G.E. 1995, ApJ,445, L129 Binney, J., & Tremaine, S. 1987, Galactic Dynamics, (Princeton: Princeton Univ. Press), 580 Bohringer, H. & Hensler, G. 1988, A&A, 215, 147 Bond, J.R., Arnett, W.D., & Carr, B.J. 1984, ApJ, 280, 825 Lequeux, J., Fort, B., Dantel-Fort, M., Cuillandre, J.C., & Mellier, Y. 1996, A&A, in press (astro- ph/9606172) Liebert, J., Dahn, C.C., & Monet, D.G. 1988, ApJ, 332, 891 Maeder, A. 1992, A&A, 264, 105 Mathews, G.J., Fields, B.D., Charlot, S., & Schramm, D.N. 1996, to be submitted Mathews, G.J., & Schramm, D.N. 1993, ApJ, 404, 468 Miller, G.E., & Scalo, J.M. 1979, ApJS, 41, 513 Mulchaey, J.S., Davis, D.S., Mushotzky, R.J., & Burstein, D. 1993, ApJ, 404, L9 Mulchaey, J.S., Davis, D.S., Mushotzky, R.J., & Burstein, D. 1996, ApJ, 456, 80 Chabrier, G., Segretain, L., & M´era, D., 1996, ApJ, submitted (astro-ph/9606083) Mulchaey, J.S., Mushotzky, R.J., Burstein, D. 1996, & Davis, D.S. ApJ, 456, L8 Charlot, S., & Silk, J. 1995, ApJ, 445, 124 Copi, C.J., Schramm, D.N., & Turner, M.S. 1995, Science, 267, 192 Davis, D.S., Mulchaey, J.S., Mushotzky, R.J., & Burstein, D. 1996, ApJ, 460, 601 Fich, M., & Tremaine, S. 1991, ARAA, 409 Fields, B.D., Kainulainen, K., Olive, K.A., & Thomas, D. 1996, New Astron., in press R.J. 1993, Mushotzky, of Texas/PASCOS 1993, ed. M. Sredniki (New York: Ann. N.Y. Acad. Sci.) Proceedings Pagel, B.E.J. 1988, in Evolutionary Phenomena in Galaxies, eds. J. Beckman & B.E.J. Pagel (Cam- bridge: Cambridge University Press) Peebles, P.J.E., 1993, Principles of Physical Cosmol- ogy, (Princeton: Princeton Univ. Press) Pildis, R., Bregman, J.N., & Evrard, A.E. 1995, ApJ, Flynn, C., Gould, A., & Bahcall, J.N. 1996, ApJ, 466, 443, 514 L55 (astro-ph/9503035) Pildis, R.A., & McGaugh, S.S. 1996, ApJ, submitted Fujimoto, M.Y., Sugiyama, K., Iben, I., Hollowell, D. (astro-ph/9608038) 1995, ApJ, 444, 175 Fukazawa, Y., et al. 1996, Pub. Astr. Soc. Japan, 48, 395 Ponman, T.J., Bourner, P.D.J., Ebeling, H., & in press (astro- Bohringer, H. 1996, MNRAS, ph/9607114) 14 Rich, R.M. 1990, ApJ, 362, 604 FIGURE CAPTIONS Richstone, D., Gould, A., Guhathakurta, P, & Flynn, C. 1992, ApJ, 38, 354 Ryu, D., Olive, K.A., & Silk, J. 1990, ApJ, 353, 81 Sackett, P.D., Morrison, H.L., Harding, P., & Boro- son, T.A. 1994, Nature, 370, 441 Silk, J. 1993, Phys Reports, 227, 143 1. Schematic diagram of model features. 2. The total galactic star formation rate ψ = ncψcloud (solid curve) and number of clouds nc (dashed curve) as a function of time. Two bursts of star formation are evident, one shortly after decoupling when the density is high, and one during halo collapse (5 Gyr). Suto, Y., Makishima, K., Ishisaki, Y., & Ogasaka, Y. 3. Mass evolution for (a) the halo, (b) the local 1996, ApJ, 461, L33 (astro-ph/9602061) Group. Tinsley, B. 1980, Fund. Cosmic Phys., 5, 287 4. Metallicity evolution for the halo and Local Udalski, A. et al. 1994, Acta Astronomica, 44 165 White, S.D.M., Navarro, J.F.,Evrard, A.E., Frenk, C.S. 1993, Nature, 366, 429 Zepf, S.E. & Silk, J. 1996, ApJ, 466, 114 (astro- ph/9501166) Zurek, W.H., Quinn, P.J. & Salmon, J.K. 1988, ApJ, 330. 519 Group. 5. Halo metallicity distribution. The points are for globular clusters (Pagel 1988); only the low metallicity members ([Z] < −1) are included in the analysis. The theory (histogram) is binned in the same way as the data, and normalized to minimize χ2. 6. The halo white dwarf luminosity function, with L in units of L⊙. Only results for halo white dwarfs are shown. The two peaks correspond to the two bursts of star formation; the more luminous peak is the more recent burst at t = 5 Gyr. This second peak has MV >∼ 20, well below current observational limits. This 2-column preprint was prepared with the AAS LATEX macros v4.0. 15 Hot Intergalactic Medium Group wind X-ray emitting intra-group medium Local Group Galactic infall Galactic wind Hot intercloud medium cloud wind infall Hot gas cooling ejecta Star-forming cold molecular gas Protogalactic Clouds Galactic Halo Halo stars and remnants Cloud mergers Survive merging? yes Disk+Bulge 106 105 104 103 102 101 100 10-1 ψ (solar mass yr-1) ncloud 0 5 10 15 t (Gyr) 1013 ) s e s s a m r a l o s ( M 1012 1011 1010 109 0 total mass gas mass remnant mass 5 10 15 t (Gyr) 1014 ) s e s s a m l r a o s ( M 1013 1012 1011 0 total mass baryonic mass gas mass 5 10 15 t (Gyr) 10-2 i Z 10-3 10-4 cold gas hot gas halo cluster cloud stars halo stars 0 5 10 15 t (Gyr) 15 10 5 ] Z [ ∆ / N ∆ 0 -3 -2 -1 0 [Z] ) L d / D W m d ( g o l 11.0 10.5 10.0 9.5 9.0 8.5 8.0 0 -1 -2 -3 -4 log (L) -5 -6 -7
astro-ph/0307212
1
0307
2003-07-10T14:50:06
Weak Gravitational Lensing by Large-Scale Structure
[ "astro-ph" ]
Weak gravitational lensing provides a unique method to map directly the distribution of dark matter in the universe and to measure cosmological parameters. This cosmic-shear technique is based on the measurement of the weak distortions that lensing induces in the shape of background galaxies as photons travel through large-scale structures. This technique is now widely used to measure the mass distribution of galaxy clusters and has recently been detected in random regions of the sky. In this review, we present the theory and observational status of cosmic shear. We describe the principles of weak lensing and the predictions for the shear statistics in favored cosmological models. Next, we review the current measurements of cosmic shear and show how they constrain cosmological parameters. We then describe the prospects offered by upcoming and future cosmic-shear surveys as well as the technical challenges that have to be met for the promises of cosmic shear to be fully realized.
astro-ph
astro-ph
Annual Reviews of Astronomy and Astrophysics 2003 – 41: 645–668 Copyright c(cid:13) 2003 by Annual Reviews (in press) Weak Gravitational Lensing by Large-Scale Structure Alexandre Refregier Service d’Astrophysique, CEA/Saclay, 91191 Gif sur Yvette, France; email: [email protected] KEYWORDS: Cosmology, Dark Matter, Cosmic Shear, Structure Formation, Statistics ABSTRACT: Weak gravitational lensing provides a unique method to map directly the dis- tribution of dark matter in the universe and to measure cosmological parameters. This cosmic- shear technique is based on the measurement of the weak distortions that lensing induces in the shape of background galaxies as photons travel through large-scale structures. This technique is now widely used to measure the mass distribution of galaxy clusters and has recently been detected in random regions of the sky. In this review, we present the theory and observational status of cosmic shear. We describe the principles of weak lensing and the predictions for the shear statistics in favored cosmological models. Next, we review the current measurements of cosmic shear and show how they constrain cosmological parameters. We then describe the prospects offered by upcoming and future cosmic-shear surveys as well as the technical challenges that have to be met for the promises of cosmic shear to be fully realized. 1 INTRODUCTION Gravitational lensing provides a unique method to directly map the distribution of dark matter in the universe. It relies on the measurement of the distortions that lensing induces in the images of background galaxies (for reviews, see Bartel- mann & Schneider 1999, Bernardeau 1999, Kaiser 1999, Mellier 1999, Narayan & Bartelmann 1996, Schneider 1995, Wittman 2002). This method is now widely used to map the mass of clusters of galaxies (see Fort & Mellier 1994 for a review) and has been extended to the study of superclusters (Gray et al. 2002, Kaiser et al. 1998) and groups (Hoekstra et al. 2001). Recently, weak lensing was statisti- cally detected for the first time in random patches of the sky (Brown et al. 2003; Bacon et al. 2002; Bacon, Refregier & Ellis 2000; Hamana et al. 2003; Hammerle et al. 2002; Hoekstra et al. 2002a; Hoekstra, Yee & Gladders 2002a; Jarvis et al. 2002; Kaiser, Wilson & Luppino 2000; Maoli et al. 2001; Refregier, Rhodes, & Groth 2002; Rhodes, Refregier & Groth 2001; van Waerbeke et al. 2000; van Waerbeke et al. 2001a; van Waerbeke et al. 2002; Wittman et al. 2000). These cosmic shear surveys provide direct measurements of large-scale structure in the universe and, therefore, of the distribution of dark matter. Unlike other meth- ods that probe the distribution of light, weak lensing measures the mass and can thus be directly compared to reliable theoretical models of structure forma- tion. Cosmic shear can therefore be used to measure cosmological parameters in the context of these models, thereby opening wide prospects for cosmology (Bernardeau, van Waerbeke & Mellier 1997; Hu & Tegmark 1999; Jain & Seljak 1 2 Refregier Figure 1: Illustration of the effect of weak lensing by large-scale structure. The photon tra jectories from distant galaxies (right) to the observer (left) are deflected by intervening large-scale structure (center). This results in coherent distortions in the observed shapes of the galaxies. These distortions, or shears, are on the order of a few percent in amplitude and can be measured to yield a direct map of the distribution of mass in the universe. 1997; Kaiser 1998). In the present review, we describe the theoretical and observational status of cosmic shear. Earlier reviews of this fast-evolving field can be found in Hoekstra et al. (2002b), Mellier et al. (2001), van Waerbeke et al. (2002), Wittman (2002). Here, we first describe the principles of weak lensing (Section 2). We then summarize the different statistics used to measure cosmic shear and describe how they are used to constrain cosmological parameters (Section 3). In Section 4, we survey the different methods used to derive the lensing shear from the shapes of background galaxies. We present, in Section 5, the current observations and their cosmological significance. Future cosmic-shear surveys and the prospects they offer for cosmology are described in Section 6. In Section 7, we outline how systematic effects present challenges that must be met for the potential of these future surveys to be fully realized. We summarize our conclusions in Section 8. 2 THEORY The idea of cosmic shear can be traced back to a lecture given by Richard Feyn- man at Caltech in 1964 (J.E. Gunn, personal communication). Several theo- rists (e.g., Gunn 1967, Jaroszy´nsky et al. 1990, Kristian & Sachs 1966, Lee & Paczy´nsky 1990, Schneider & Weiss 1988) then studied the propagation of light in an inhomogeneous universe. Predictions for the statistics of the weak-lensing distortions were then computed in a modern cosmological context by several groups (Babul & Lee 1991, Blandford et al. 1991, Kaiser 1992, Miralda-Escud´e 1991, Villumsen 1996). More recently, the power of cosmic shear to measure cosmological parameters was the ob ject of many theoretical studies (Bernardeau, van Waerbeke & Mellier 1997; Jain & Seljak 1997; Hu & Tegmark 1999; Kaiser 1998; Kamionkowski et al. 1997; van Waerbeke, Bernardeau & Mellier 1999). In this section, we briefly describe the principles of weak lensing and show how this technique can be used to map the dark matter in the universe. As they travel from a background galaxy to the observer, photons get deflected Weak Gravitational Lensing 3                       γ 2     ε 2                       ε 1 γ1      Figure 2: Illustration of the geometrical meaning of the shear γi and of the ellipticity ǫi . A positive (negative) shear component γ1 corresponds to an elon- gation (compression) along the x-axis. A positive (negative) value of the shear component γ2 corresponds to an elongation (compression) along the x = y axis. The ellipticity of an ob ject is defined to vanish if the ob ject is circular (center). The ellipticity components ǫ1 and ǫ2 correspond to compression and elongations similar to those for the shear components. by mass fluctuations along the line of sight (see Figure 1). As a result, the appar- ent images of background galaxies are sub ject to a distortion that is characterized by the distortion matrix: Ψij ≡ ∂ (δθi ) ∂ θj ≡ (cid:18) κ + γ1 κ − γ1 (cid:19) , γ2 γ2 where δθi (θ) is the deflection vector produced by lensing on the sky. The conver- gence κ is proportional to the pro jected mass along the line of sight and describes overall dilations and contractions. The shear γ1 (γ2 ) describes stretches and com- pressions along (at 45◦ from) the x-axis. Figure 2 illustrates the geometrical meaning of the two shear components. The distortion matrix is directly related to the matter density fluctuations along the line of sight by Ψij = Z χh 0 where Φ is the Newtonian potential, χ is the comoving distance, χh is the co- moving distance to the horizon, and ∂i is the comoving derivative perpendicular to the line of sight. The radial weight function g(χ) is given by g(χ) = 2 Z χh χ where r = a−1DA , and DA is the angular-diameter distance. The function n(χ) is the distribution of the galaxies as a function of the comoving distance χ from the observer and is assumed to be normalized as R dχn(χ) = 1. As we discuss in Section 4, galaxy shapes can be averaged over a patch of the sky to measure the shear, which is thus an observable. The shear pattern expected in r(χ)r(χ′ − χ) r(χ′ ) dχ g(χ)∂i ∂j Φ, dχ′ n(χ′ ) , (1) (2) (3) 4 Refregier Figure 3: Shear map derived by ray-tracing simulations by Jain, Seljak & White (2000). The size and direction of each line gives the amplitude and position angle of the shear at this location on the sky. The displayed region is 1◦ × 1◦ for an SCDM (Einstein-De Sitter) model. Tangential patterns about the overdensities corresponding to clusters and groups of galaxies are apparent. A more complex network of patterns is also visible outside of these structures. The root-mean- square shear is approximately 2% in this map. (From Jain et al. 2000) a standard Cold Dark Matter (SCDM) model is shown in Figure 3 for a 1 × 1 deg2 region. Jain, Seljak, & White (2000) derived this map from ray tracing through N-body simulations. Tangential patterns around the overdensities corresponding to clusters and groups of galaxies, along with a more complicated network of shear fluctuations, are apparent. By inverting the lensing equation (Equation 2), the shear map can be converted into a map of the pro jected mass κ and, therefore, of the dark matter distribution. 3 COSMIC-SHEAR STATISTICS AND COSMOLOGY The statistical characteristics of the cosmic-shear field can be quantified using a variety of measures, which can then be used to constrain cosmological models. First, we consider the most basic two-point statistic of the shear field, namely the two-dimensional power spectrum (Jain & Seljak 1997, Kaiser 1998, Kamionkowski et al. 1997, Schneider et al. 1998a). The shear power spectrum Cl is defined as a function of multipole moment l (or inverse angular scale) by P2 i=1 h eγi (l) eγi (l′ )i = (2π)2 δ(l − l′ )Cl , where tildes denote Fourier transforms (with the conventions of Bacon, Refregier, & Ellis 2000), δ is the two-dimensional Dirac-delta function, and Weak Gravitational Lensing 5 Figure 4: Shear power spectrum for different cosmological models and for source galaxies at zs = 1. The SCDM model is COBE normalized and thus has a higher amplitude than the three cluster-normalized models ΛCDM, OCDM, and τ CDM. The thin dashed line shows the ΛCDM spectrum for linear evolution of structures. Notice that for l > 1000 (corresponding approximately to angular scales θ < 10′ ) the lensing power spectrum is dominated by nonlinear structures. (4) Cl = the brackets denote an ensemble average. Applying Limber’s equation in Fourier space (e.g., Kaiser 1998) to Equation 2 and using the Poisson equation, one can easily express the shear power spectrum Cl in terms of the three-dimensional power spectrum P (k , χ) of the mass fluctuations δρ/ρ and obtain dχ (cid:20) g(χ) , χ(cid:19) , ar(χ) (cid:21)2 16 (cid:18) H0 c (cid:19)4 P (cid:18) l m Z χh 9 Ω2 r 0 where a is the expansion parameter, and H0 and Ωm are the present value of the Hubble constant and matter density parameter, respectively. The lensing power spectra for four CDM models are shown in Figure 4 (see color insert) (see Bacon, Refregier & Ellis 2000 for the exact cosmological parameter values of each model). They were derived using the fitting formula for the nonlinear matter power spectrum P (k , χ) of Peacock & Dodds (1996). In Figure 4, the galaxies were assumed to lie on a sheet at a redshift zs = 1. A more realistic redshift distribution n(z ) would require corrections of only approximately 10% on these power spectra, as long as the median redshift of the galaxies were kept at zs = 1. The three cluster-normalized models (ΛCDM, OCDM and τ CDM) yield power spectra of similar amplitudes but with different shapes. The COBE-normalized SCDM model has more power on a small scale and thus yields a larger normal- ization. For the ΛCDM model, the power spectrum corresponding to a linear evolution of structures is also shown in Figure 4 for comparison. For l > ∼ 1000 6 Refregier (corresponding to angular scales smaller than approximately 10’), nonlinear cor- rections dominate the power spectrum (Jain & Seljak 1997), making cosmic-shear sensitive to gravitational instability processes. The measurement of the lensing power spectrum can thus be used to measure cosmological parameters, such as Ωm , ΩΛ , σ8 , and Γ. (Bernardeau, van Waerbeke & Mellier 1997; Hu & Tegmark 1999; Jain & Seljak 1997; Kaiser 1998; van Waer- beke, Bernardeau & Mellier 1999). A full-sky cosmic-shear survey would yield a precision of these parameters comparable to that for future cosmic microwave background (CMB) missions. More realistically for the short term, a precision on the order of 10% can be achieved with surveys of approximately 10 square degrees. Such cosmic-shear surveys can also be combined with CMB anisotropy measurements to break degeneracies present when the CMB is considered alone (Hu & Tegmark 1999; Contaldi, Hoekstra & Lewis 2003). This would yield im- provements in the precision of cosmological parameters by approximately one order of magnitude. In addition, the use of photometric redshifts can provide a tomographic measurement of matter fluctuations and improve the precision of cosmological parameters by up to an order of magnitude (Hu 1999, Hu & Keeton 2002, Taylor 2001). In practice, it often is more convenient to measure other two-point statistics. γ ≡ hγ 2 i in randomly placed cells is In particular, the variance of the shear σ2 widely used. It is related to the shear power spectrum by 2π Z ∞ dl lCl (cid:12)(cid:12)(cid:12)fWl (cid:12)(cid:12)(cid:12) 1 2 σ2 , (5) γ = 0 where fWl is the Fourier transform of the cell aperture. The shear two-point cor- relation functions (eg. Kaiser 1998) are Fourier transforms of the power spectra and have also been measured by various groups (eg. Bacon et al. 2002; van Waerbeke et al. 2001a; Hoekstra et al. 2002a). The Map statistics (Schneider et al. 1998a) is another convenient statistic based on the average of the shear within a compensated filter. Its advantages are that its window function in l-space is narrow, that adjacent cells are effectively uncorrelated, and that it can be easily related to the statistics of the pro jected mass κ. In analogy with electromagnetism, a tensor field like the shear field γi can be decomposed into an electric (E, or gradient) component and a magnetic (B, or curl) component (Crittenden et al. 2001b, Kaiser 1992, Kamionkowski et al. 1997, Stebbins 1996). Because the distortions it induces arise from a scalar field (the gravitational potential), weak lensing only produces E-type fluctuations. On the other hand, systematic effects and intrinsic galaxy alignments are likely to produce both E-type and B-type fluctuations (See Heavens 2001 for a review; also see Section 7.6). The presence of B-modes can thus be used as a measure of contaminants to the cosmic-shear signal. In practice, this decomposition can be performed using the Map statistic (Schneider et al. 1998a, van Waerbeke et al. 2001a) or the power spectrum (Hu & White 2000, Padmanabhan, Seljak & Pen 2002, Pen et al. 2001). As is apparent in Figure 3, the shear field is not Gaussian but, instead, has apparent coherent structures. Bernardeau, van Waerbeke, & Mellier (1997) have shown that the skewness hκ3 i hκ2 i2 of the convergence field κ can be used to break the degeneracy between σ8 and S3 = (6) Weak Gravitational Lensing 7 Ωm , which is present when the shear variance alone is considered. The measure- ment of higher-point statistics is thus of great cosmological interest, but their computation is made difficult by the fact that the angular scales accessible to ∼ 10′ ) are in the nonlinear regime. Predictions for higher-point cor- obervation (θ < relation functions have been calculated using perturbation theory, the hierarchi- cal ansatz, (Bernardeau & Valageas 2000; Bernardeau, van Waerbeke & Mellier 1997; Hui 1999; Munshi & Coles 2000; Munshi & Jain 2001; van Waerbeke et al. 2001b) and halo models (Cooray & Hu 2001a; Cooray, Hu, & Miralda-Escud´e 2000). These techniques have also been used to compute the full probability dis- tribution function of the convergence field (Bernardeau & Valageas 2000, Munshi & Jain 2000, Valageas 2000) and the errors in two-point statistics (Cooray & Hu 2001b; Munshi & Coles 2002; Schneider et al. 2002). Another possible way to compute the full nonlinear field is to use ray tracing through N-body simulations (Blandford et al. 1991; Hamana, Colombi & Mellier 2000; Jain, Seljak & White 2000; Premadi et al. 2001; Wambsganss, Cen, & Ostriker 1998; White & Hu 2000) to produce simulated shear maps such as those by Jain, Seljak & White (2000) (shown in Figure 3). These can be used to compute and study other proposed measures of non-Gaussianity such as peak statistics (Jain & van Waerbeke 2000), a generalized maximum likelihood (Taylor & Watts 2000), and cluster counts (Bartelmann, King & Schneider 2001). In practice, the complex geometry of surveys makes it difficult to infer a convergence κ map from the observed shear γi . For this reason, a number of researchers have recently proposed the use of high-order statistics of the shear field rather than the skewness S3 of the convergence (Bernardeau, van Waerbeke, & Mellier 2003; Schneider & Lombardi 2002; Takada & Jain 2002; Zaldarriaga & Scoccimarro 2002). 4 SHEAR MEASUREMENT METHODS Because the sought-after lensing signal is of only a few percent in amplitude, the data acquisition and analysis must be performed carefully, and systematic effects must be tightly controlled. All the current measurements of cosmic shear were derived from deep optical images taken with charged-coupled devices (CCD). It is advantageous for the exposures to be homogeneous in depth and for the ground to be sub ject to as small a seeing as possible. The fields are generally chosen to lie far away from each other to ensure that they are statistically independent and to minimize cosmic variance. The first step in the data analysis is image processing. After flat fielding, the different exposures are co-added to produce the final reduced images. If necessary, any instrumental distortion induced by the telescope optics is corrected for at this stage. This can be done very accurately by measuring the astrometric offsets from several dithered exposures. For a detailed description of the different image-processing steps, see Kaiser et al. (1999). An example of a processed deep image from the cosmic-shear survey of Bacon, Refregier & Ellis (2000) is shown in Figure 5. The next step consists of deriving an estimator for the shear from the shapes of the galaxies in the co-added images. The point-spread function (PSF), which smears the images of galaxies and is generally not circular, complicates this task. In general, the PSF varies spatially and in time, and it must be measured and 8 Refregier Figure 5: Example of an deep image in the cosmic-shear survey by Bacon, Re- fregier & Ellis (2000). This corresponds to a 1 h exposure with the EEV camera on the William Herschel Telescope (WHT). The field of view is 8′ × 16′ and achieves a magnitude depth of R ≃ 26 (5σ detection). The bright ob jects are saturated stars. The faint ob jects comprise approximately 200 stars and approxi- mately 2000 galaxies that are usable for the weak-lensing analysis. (From Bacon, Refregier & Ellis 2000) Weak Gravitational Lensing 9 modelled for each image individually. This can be done by measuring the shape of the stars in the field, whose number can be optimized by tuning the galactic latitudes of the observations. Figure 6 shows the ellipticity pattern of the PSF from one of the William Hershel telescope (WHT) fields of Bacon, Refregier & Ellis 2000. Several methods have been developed to tackle this difficult and crucial task. The more rigorous method of Kaiser, Squires & Broadhurst (KSB; 1995), further developed by Luppino & Kaiser (1997) and Hoekstra et al. (1998), replaced the earlier method by Bonnet & Mellier (1995). The KSB method is now widely used for cluster studies, and it has been used by the ma jority of the groups involved in measuring cosmic shear (see Bartelmann & Schneider 1999 for a detailed review of the KSB method). It is based on the measurement of the quadrupole moment of the galaxy surface brightness I (x), Qij ≡ Z d2xxixj w(x)I (x), where w(x) is a weight function conveniently taken to be Gaussian. These mo- ments capture the lowest-order shape information of the galaxy and can be com- bined to form the ellipticity of the galaxy (7) ǫ1 = Q11 − Q22 Q11 + Q22 , ǫ2 = 2Q12 Q11 + Q22 . (8) The first (second) component of the ellipticity describes compressions and elon- gations along (at 45◦ from) the x and y axes (see Figure 2 for an illustration). The ellipticity vanishes for circular galaxies. The first step in the KSB method is to correct the observed galaxy ellipticity ǫg ′ for the anisotropy of the PSF. The corrected galaxy ellipticity ǫg is given by i ǫg = ǫg ′ − P g sm )−1 ǫ∗ , sm (P ∗ (9) where ǫ∗ is the PSF ellipticity derived from the stars, and P g sm and P ∗ sm are the smear susceptibility tensors for the galaxy and star, respectively, and can be derived from higher moments of the images. The shear in a patch of the sky can then be measured by averaging over the (corrected) ellipticities in the patch of the sky using γ = (Pγ )−1 hǫg i, (10) where the tensor Pγ quantifies the susceptibility to shear acting before isotropic PSF smearing and is given by Pγ = P g sm )−1P g sh (P ∗ sh − P ∗ sm , (11) where the shear susceptibility tensors P g sh and P ∗ sh for the galaxies and the stars can also be measured from higher moments of their respective light distribution. The KSB method was thoroughly tested using realistic simulated images by Erben et al. (2001) and Bacon et al. (2001a). These studies showed that it is accurate to within a few tenths of percent in reconstructing an input shear and is thus sufficient for the current cosmic-shear surveys (see Section 5). However, this accuracy is insufficient for future, more sensitive surveys (see Section 6). Moreover, Kuijken (1999) and Kaiser (2000) have shown that the KSB method is ill-defined mathematically and unstable for PSF’s found in practice. 10 Refregier Figure 6: PSF ellipticity pattern measured in WHT field shown in the Figure 5: (a) PSF ellipticities measured from the stars in Figure 5. (b) Ellipticity residuals for those stars after the KSB corrections. In this survey, the root-mean-square stellar ellipticity is found to be approximately 7% before the corrections and negligible after the correction. (From Bacon, Refregier & Ellis 2000) . This inadequacy has led a number of researchers to develop alternative meth- ods. Rhodes, Refregier & Groth (2000) have modified the KSB method to be better suited for HST images. Kuijken (1999) considered a different approach that consisted of fitting the observed galaxy shape with a smeared and sheared circular model. Kaiser (2000) introduced a new method based on a finite resolu- tion shear operator. Refregier & Bacon (2003) and, independently, Bernstein & Jarvis (2001) developed a new method based on the decomposition of the galaxies into shape components or “shapelets”. The gauss-hermite orthogonal basis func- tions used in this approach allow shears and PSF convolutions to be described as simple matrix operations, using the formalism developped for the quantum har- monic oscillator (Refregier 2003). These new methods are promising but require extensive testing to establish whether they will achieve the required precision. 5 OBSERVATIONS Because the expected distortions are only of a few percent, the measurement of cosmic shear requires large survey areas and excellent image quality. Early searches for cosmic shear signals with photographic plates were unsuccesful (Kris- tian 1967; Valdes, Jarvis & Tyson 1983). Mould et al. (1994), performed the first attempt to detect a cosmic-shear signal with CCDs, but only derived an upper limit. Using the same data, Villumsen (1995) reported a 4.5σ detection. Schneider et al. (1998b) then reported a detection of cosmic shear in one of three QSO fields, an area too small to draw any constraints on cosmology. Within a few weeks, four independent groups (Bacon, Refregier & Ellis 2000; Weak Gravitational Lensing 11 Kaiser, Wilson & Luppino 2000; van Waerbeke et al. 2000; Wittman et al. 2000) reported the first firm statistical detections of cosmic shear using three different 4m-class telescopes: the Cerro Tololo Inter-American Observatory (CTIO), the Canada-France-Hawaii Telescope (CFHT), and the William Herschel Telescope (WHT). These were later confirmed by more precise measurements of the cosmic- shear amplitude from the ground (Bacon et al. 2002; Brown et al. 2003; Maoli et al. 2001; Hamana et al. 2003; Hoekstra et al. 2002a; Hoekstra, Yee & Gladders 2002a; Jarvis et al. 2002; van Waerbeke et al. 2001; van Waerbeke et al. 2002a) using 2-m – (MPG/ESO), 4-m– (WHT, CFHT, CTIO) and 8-m– (Very Large Telescope, Keck, Subaru) class ground-based telescopes. Meanwhile, cosmic shear was also detected (Hammerle et al. 2002; Rhodes, Refregier & Groth 2001) and measured (Refregier, Rhodes & Groth 2002) from space using HST. Space-based surveys are currently limited by the small field of view of HST, but they are deeper and less prone to systematics thanks to the absence of atmospheric seeing. Table 1 summarizes the existing cosmic-shear surveys and highlights the wide range of telescopes, survey areas, and depths. The shear variance σ2 γ (Equation 5) measured recently by several groups is shown in Figure 7 (see color insert) as a function of the radius a circular cell. The results by Hammerle et al. (2002) and Hamana et al. (2002a) are not displayed. Note that, for the shear variance, the data points at different angular scales are not independent. For comparison, the shear variance predicted for a ΛCDM model with σ8 = 1 is also shown in Figure 7. The model is displayed for median galaxy redshifts of zm from 0.8 to 1.0, corresponding approximately to the range of depths of the top five surveys displayed (van Waerbeke et al. 2002a; Brown et al. 2003; Bacon et al. 2002, WHT and Keck; Refregier et al. 2002). These observations are approximately consistent with each other and with the ΛCDM model on angular scales from 0.7 to 20 arcmins. This is compelling given that these were performed with different telescopes (and therefore different instrumental systematics) and independent data-analysis pipelines. The bottom two surveys (Hoekstra et al. 2002b; Jarvis et al. 2002) have a median redshift in the range zm ≃ 0.6–0.7 and yield lower shear variances. As indicated by the theoretical curves in figure 7, this redshift dependence is expected in CDM models and thus confirms the detection of a cosmic shear signal. Although the shear variance was displayed for the purpose of Figure 7, some of the groups have used instead the shear correlation function or Map statistic to quantify their lensing signal. Recently, several groups used the latter statistic to separate their signal into E and B components (see Section 3) and thus to estimate and subtract systematic effects (Jarvis et al. 2002; Hamana et al. 2003; Hoekstra, Yee & Gladders 2002; van Waerbeke et al. 2001, 2002a). Pen, van Waerbeke & Mellier (2001) and Brown et al. (2003) chose instead to measure directly the shear power spectrum for each E and B mode. Figure 8a shows the measurement of the E and B signals using the Map statistic by Hoekstra, Yee & Gladders (2002a), for several Rc -magnitude ranges in the Red-Sequence Cluster (RCS) survey. They measure a clear lensing signal apparent as E modes. However, the significant B modes reveals the presence of residual systematics ∼ 10′ ), especially for bright galaxies. on small scales (θ < In their analysis, these authors use the amplitude of the B modes as a measure systematic uncertainties. Existing cosmic-shear measurements already yield interesting constraints on the amplitude of the matter power spectrum σ8 on which the lensing signal strongly depends. For instance, Figure 8b shows cosmological constraints for 12 Refregier Figure 7: Shear variance σ2 γ as a function of the radius θ of a circular cell. The data points correspond to recent results from different groups: van Waerbeke et al. (2002a), Brown et al. (2003), Bacon et al. (2002, WHT and Keck), Refregier et al. (2002), Hoekstra et al. (2002b), Jarvis et al. (2002). When relevant, the inner error bars correspond to noise only, whereas the outer error bars correspond to the total error (noise + cosmic variance). The measurements by Hammerle et al. (2003) are not displayed. The solid curves (2002) and Hamana et al. show the predictions for a ΛCDM model with Ωm = 0.3, σ8 = 1, and Γ = 0.21. The galaxy median redshift was taken to be zm = 1.0, 0.9, and 0.8, from top to bottom, respectively, corresponding approximately to the range of depth of the top five surveys. The bottom two surveys (Hoekstra et al. 2002b; Jarvis et al. 2002) have a median redshift in the range zm ≃ 0.6–0.7 and, as expected, yield lower shear variances. Weak Gravitational Lensing 13 Figure 8: Cosmic shear measurement by Hoekstra, Yee & Gladders (2002a) from the RCS survey. (a): measurement of the E (top panels) and B (bottom panels) modes of the Map statistics variance as a function of aperture scale θ . Several ranges of the Rc -magnitude are considered. The errors correspond to 1σ sta- tistical uncertainties. The presence of significant B Modes reveal the presence of residual systematics on small scales (θ ≤ 10′ ). (b): Cosmological constraints derived from their measurement of the E -mode Map statistics for 22 < Rc < 24 after accounting for the residual B -modes. The joint constraints are shown for Ωm and σ8 in the ΛCDM model. Priors from CMB and galaxy surveys were used to marginalize over Ωtot , Γ, and the source redshift zs . The contours indicate 68.3%, 95.4%, and 99.9% confidence levels. (From Hoekstra, Yee & Gladders 2002a) 14 Refregier Table 1: current cosmic-shear surveys Area (deg2 ) Mag. limit Telescope Reference R < CTIO 1.0 ∼ 26 Wittman et al. 2000 1.7 CFHT van Waerbeke et al. 2000 0.96 CFHT Kaiser et al. 2000 0.5 WHT Bacon et al. 2000 VLT 0.65 Maoli et al. 2001 0.05 Rhodes et al. 2001 HST/WFPC2 6.5 van Waerbeke et al. 2001a CFHT Hammerle et al. 2002 HST/STIS 0.02 Hoekstra et al. 2002a CFHT, CTIO 24 8.5 van Waerbeke et al. 2002a CFHT 0.36 HST/WFPC2 Refregier et al. 2002 WHT, Keck Bacon et al. 2002 1.6 Hoekstra et al. 2002b CFHT, CTIO 53 1.25 MPG/ESO Brown et al. 2003 2.1 Subaru Hamana et al. 2003 Jarvis et al. 2002 CTIO 75 R < ∼ 24 I < ∼ 24.5 hI i ≃ 23.5 R < ∼ 26 R < ∼ 24 R < ∼ 25 R < ∼ 26 R < ∼ 23 I < ∼ 24, V < ∼ 25 R < ∼ 26 I < ∼ 24.5 I < ∼ 26 I < ∼ 24.5 a σ8 1.50+0.50 −0.50 1.03+0.03 −0.03 0.91+0.25 −0.30 0.88+0.02 −0.02 b c 0.81+0.07 −0.09 0.98+0.06 −0.06 0.94+0.14 −0.14 0.97+0.13 −0.13 0.86+0.04 −0.05 0.72+0.09 −0.09 0.69+0.18 −0.13 0.71+0.06 −0.08 a for ΛCDM model with Ωm = 0.3, ΩΛ = 0.7; Γ is marginalised over or set to 0.21 when possible; errors correspond to 68%CL b for combination of Maoli et al. (2001), van Waerbeke et al. (2000), Bacon et al. (2000), and Wittman et al. (2000); cosmic variance not included. c cosmic variance not included. a ΛCDM model derived by Hoekstra, Yee & Gladders (2002a) from the measure- ment of the Map statistic for their 22 < Rc < 24 galaxy sample. The degeneracy between σ8 and Ωm apparent in the figure is typical of cosmic-shear measurements involving only two-point statistics. Hoekstra, Yee & Gladders (2002a) found that m = 0.46+0.05 the constraints are well described by σ8Ω0.52 −0.07 (95% CL), where priors from CMB and galaxy survey data have been used to marginalize over Γ and ΩΛ . Other surveys find constraints of the same form with similar exponents for Ωm . Table 1 lists the values of σ8 found by the different groups. The errors have been converted to 68% confidence level (CL) when necessary, and values of Ωm = 0.3, ΩΛ = 0.7, and Γ = 0.21 have been assumed, when possible. The most recent values are displayed in Figure 9 (see color insert), along with their average σ8 ≃ 0.83 ± 0.04 (1σ). The values derived by the different cosmic shear groups are in the range σ8 ≃ 0.7–1.0, with the most recent measurements (Brown et al. 2003; Hamana et al. 2003; Jarvis et al. 2002) yielding lower values. The 2–3σ mutual inconsistencies between some of the measurements may be symptomatic of a small level of residual systematics. In particular, calibration errors in the shear measurement method would not be detected via the E -B decomposition and are a likely explanation (see Hirata & Seljak 2003). Another source of discrepancy arises from the different fitting functions for the non-linear evolution of the power spectrum (see section 2). Apart from Brown et al. (2003; see also Contaldi et al. 2003) who used the more accurate results of Smith et al. (2003), the other groups used the Peacock & Dodds (1997) fitting function which yields an underprediction of σ8 by roughly 5–10%. The impact of these systematics will be discussed further in section 7. Weak Gravitational Lensing 15 Figure 9: Comparison of the determination of σ8 by different groups and methods. The errors have all been converted to 1σ , and a ΛCDM model with Ωm = 0.3 and Γ = 0.21 was assumed when possible. The vertical dotted lines show the average σ8 ≃ 0.83 ± 0.04 of the seven cosmic-shear measurements and associated 1σ error. The normalization from cluster abundance (e.g., Pierpaoli et al. 2002) as well as that derived from CMB anisotropies and local large-scale structure (galaxy and Lyman α surveys; Spergel et al. 2003) are also shown. Interestingly, an independent measurement of σ8 is provided by the abun- dance of X-ray clusters and can be directly compared to this value. Initially, this method yielded normalisations of σ8 ∼ 0.9 (Eke et al. 1998; Pierpaoli, Scott & White 2001; Viana & Liddle 1999), in agreement with the early cosmic shear re- sults (see table 1 and figure 9). This value was subsequently revised downward to σ8 ∼ 0.7–0.8 by the use of the observed, rather than simulated, mass-temperature relation in clusters (Borgani et al. 2001; Reiprich & Bohringer 2001; Seljak 2001; Viana, Nichol & Liddle 2001; Pierpaoli, Scott & White 2002). Recently, Spergel et al. (2003) derived a value of σ8 = 0.84 ± 0.04 (68%CL) from a joint analy- sis of CMB anisotropy measurements from the Wilkinson Microwave Anistropy Probe (WMAP) and other experiments, galaxy clustering and the Lyman α for- est. For comparison, these results along with a representative value of the revised cluster-abundance normalization (Pierpaoli, Scott & White 2002) are also shown in figure 9. The average of the cosmic shear results is formally in good agreement with the determination of σ8 using the other techniques. Future surveys are how- ever needed to confirm this, by resolving the discrepancies between the current cosmic shear measurements. Also, a full likelihood analysis would be required to establish the significance of the agreement between the different techniques (see Contaldi et al. 2003). This comparison is important as it provides a strong test of the ΛCDM model, the gravitational instability paradigm, the physics of clusters, and of the biased formation of galaxies. As discussed in Section 3, the degeneracy between σ8 and Ωm can be broken by measuring higher-order correlation functions of the lensing field. Bernardeau, 16 Refregier Mellier, & van Waerbeke (2002) recently reported the first detection of a non- Gaussian shear signal using the 3-point shear correlation function formalism of Bernardeau, van Waerbeke & Mellier (2003). Another measure of non-gaussianity was recently performed by Miyazaki et al. (2002) using peak statistics in their Subaru survey. Although these results are consistent with that expected from structure formation models, larger survey areas are needed to infer cosmological constraints. Another approach to probe the dark matter is to measure the bias between the mass and galaxies by cross-correlating the shear map with that of the light from foreground galaxies in the same region of the sky (Cooray 2002, Schneider 1998, van Waerbeke 1998). Hoekstra, Yee & Gladders (2001a), Hoekstra et al. (2002b), and Wilson, Kaiser & Luppino (2001) have recently measured this cross- correlation on large scales. 6 FUTURE SURVEYS AND PROSPECTS The existing measurements described above are primarily limited by statistics. They will therefore be improved upon by ongoing surveys on existing telescopes, such as the Legacy Survey on CFHT (CFHTLS; Mellier et al. 2000), the Deep Lens Survey (Wittman et al. 2002), surveys with the Subaru telescope (Hiroyasu et al. 2001) and the Sloan Digital Sky Survey (Stebbins, McKay & Frieman 1995). Future instruments dedicated to surveys and for which cosmic shear is a primary science driver are being planned, such as Megacam on CFHT (Boulade et al. 2000), the Visible and Infrared Survey Telescope for Astronomy (VISTA; Taylor et al. 2003), the Large aperture Synoptic Survey Telescope (LSST; Tyson et al. 2002a,b), or the novel Panoramic Survey Telescope and Rapid Response System (Pan-STARRS; Kaiser, Tonry & Luppino 2000). From space, the new Advanced Camera for Surveys (ACS) on HST and, much more ambitiously, the future Supernova Acceleration Probe satellite (SNAP; Perlmutter et al. 2003; Rhodes et al. 2003; Massey et al. 2003; Refregier et al. 2003) also offer exciting prospects. Table 2 lists the characteristics of some of these future surveys. Broadly speaking, ground-based measurements will cover large areas, whereas space-based surveys will yield higher-resolution maps and reduced systematics thanks to the absence of atmospheric seeing. Radio surveys offer another interesting prospect. Ongoing efforts are aimed at detecting cosmic shear with the FIRST radio survey (Chang & Refregier 2002, Kamionkowski et al. 1997, Refregier et al. 1998). The future radio telescopes LOFAR (Low Frequency Array) and SKA (Square Kilometer Array) will yield cosmic-shear measurements of comparable sensitivity to the most ambitious op- tical surveys (Schneider 1999). The advantages of radio surveys are that they cover large solid angles, that the bright radio sources are at a higher redshift, and that the PSF is fully predictable and reproducible. These future surveys will provide very accurate measurements of cosmolog- ical parameters through the measurement of the lensing power spectrum and higher-order statistics. Figure 10 (see color insert) shows, for instance, the ex- cellent accuracy with which the lensing power spectrum will be measured with the SNAP wide survey. They will also allow us to test some of the foundations of the standard cosmological model. For example, the measurement of the power spectrum on nonlinear scales at different redshift slices, and of the hierarchy of Weak Gravitational Lensing 17 Table 2: Future cosmic-shear surveys Area Telescope/Survey Ground/Space Diameter FOV (deg2 ) (deg2 ) (m) 28 2 × 0.3 2 × 4 ground DLS 172 1 3.6 ground CFHTLS x100d 1 2.6 ground VST VISTAe 4 2 10000 ground 31000 4 × 4 4 × 1.8 ground Pan-STARRS 30000 7 8.4 ground LSST SNAP 2 0.7 300 space a planned start date of the cosmic shear surveys b references: 1: Wittman et al. 2002; 2: Mellier et al. 2000; 3: Kuijken et al. 2002; 4: Taylor et al. 2003; 5: Kaiser et al. 2000; 6: Tyson et al. 2002a,b; 7: Perlmutter et al. 2003, Rhodes et al. 2003, Massey et al. 2003, Refregier et al. 2003 c the survey will be complete in 2003 d a survey of several 100 deg2 is planned e assuming the availability of the optical camera 1999c 2003 2004 2007 2008 2012 2011 Starta Ref.b 1 2 3 4 5 6 7 high-order correlation functions, will yield a direct test of the gravitational in- stability paradigm. The lensing power spectrum can also be used to measure the equation of the state of the dark energy w and thus complement supernovae measurements in the constraining of quintessence models (Benabed & Bernardeau 2001; Hui 1999; Huterer 2001; Hu 2001, 2002; Munshi & Wang 2002; Weinberg & Kamionkowski 2002). Figure 10 shows, for instance, that a change of 40% in w can easily be measured by SNAP (see Refregier et al. 2003). Cosmic-shear measurements can also be used to test general relativity (Uzan & Bernardeau 2001). Another promising approach to measure weak lensing is to use the fluctuations of the CMB temperature as the background sources (Bernardeau 1997, 1998; Cooray & Kesden 2002; Hirata & Seljak 2002; Hu 2000; Seljak 1996; Seljak & Zaldarriaga 1999; Zaldarriaga & Seljak 1998). Because these fluctuations are produced at a redshift of approximately 1100, CMB lensing provides a probe of the evolution of mass fluctuations at redshifts larger than those probed by optical galaxies. Lensing indeed produces distinct non-Gaussian signatures that can be used to reconstruct the foreground mass distribution and probe the growth of structures. This approach will become feasible with the advent of future CMB missions such as Planck Surveyor or ground-based instruments with high angular resolutions. 7 CHALLENGE: SYSTEMATIC EFFECTS For the future surveys described above to yield their full potential, a number of challenges must be met. Indeed, these surveys require the measurement of shears on the order of 1% with an accuracy better than 0.1%; thus, they require a very tight control of systematic effects. In the following, we review the main sources of systematics. 18 Refregier Figure 10: Prospects for the measurement of the weak-lensing power spectrum with future weak-lensing surveys. The solid line represents the lensing power spectrum of a ΛCDM model with Ωm = 0.3 (and an equation of state w = −1). The boxes correspond to the band-averaged 1σ errors for the SNAP wide survey (300 deg2 , 100 galaxies per arcmin2 with a root-mean-square intrinsic ellipticity dispersion of σǫ = 0.31). The precision is highest in the nonlinear region [see the linear ΛCDM power spectrum (dotted line)] and will thus provide a test of the gravitational instability paradigm. This model can easily be distinguished from models that differ from it by a 17% change in Ωm (dashed line) or by a 40% change in the dark-energy equation of state w (dot-dashed line). Weak Gravitational Lensing 7.1 PSF Anisotropy 19 The most serious systematic for ground-based surveys is that produced by the rather large PSF ellipticities (∼ 10%) observed by the different groups. The KSB method (see Section 4) provides a correction of the PSF anisotropy by, at most, a factor of approximately 10, which is insufficient for a precision of 0.1% in shear. Although new shear measurement methods may improve upon this (see Section 4 and below), the correction is fundamentally limited by the finite surface density and signal-to-noise ratio of the stars used to measure the PSF shape. It is, therefore, necessary for the PSF ellipticity to be reduced in hardware as well as in software. This can be done by using tighter constraints in the tracking and optical systems of the new telescopes and instruments as well as with interactions between telescope designers, engineers, and scientists. 7.2 Shear Measurement Method Even if the PSF ellipticity were guaranteed to be isotropic, future measurements will be limited by the precision of the shear measurement methods. The KSB method is accurate to measure shears of approximately 1% with only a 10% accuracy (Bacon et al. 2001, Erben et al. 2001). This will soon become an important limiting factor (see Hirata & Seljak 2003). Although new methods (see Section 4) promise to improve upon this, more extensive simulations are required to establish that the same accuracy can be reached for shears of 0.1%. 7.3 Redshift Distribution To convert cosmic-shear measurements into constraints on cosmological param- eters, the redshift distribution of the background galaxies must be known (see Section 4). The uncertainty in the median galaxy redshift is already one of the dominant contributions to the uncertainty in the amplitude of the matter power spectrum from cosmic shear. The determination of the galaxy redshift distribu- tion is made difficult by the depth of the cosmic-shear surveys. In addition, the sample of galaxies is not simply magnitude limited, but it is sub ject to complex selection cuts throughout the shear measurement pipeline. Dedicated spectro- scopic surveys and photometric redshift studies (such as that by Brown et al. 2003) are thus required to overcome this limitation. 7.4 CCD Nonlinearities All the shear measurement methods rely on the assumption that the instrumental response is linear. It is therefore important to test whether CCD cameras do not have subtle pixel-to-pixel nonlinearities that would induce biases in the shear measurements. The mean shear offset and the shear gradient across the chip found by van Waerbeke et al. (2001) could be due to this effect. 7.5 Overlapping Isophotes Two neighboring galaxies in an image yield ”peanut-shaped” isophotal contours and thus appear to have aligned ellipticities. This overlapping-ellipticities effect may produce a spurious ellipticity correlation signal on small scales. van Waer- beke et al. (2000) suggested that this effect may explain the excess shear signal, 20 Refregier which they measured on small scale and which disappeared when close pairs were discarded. Evidence for this effect was also found in the image simulation by Ba- con et al. (2001). More extensive and detailed simulations would need to be performed to ascertain and calibrate this effect. 7.6 Intrinsic Correlations The measurement of the lensing shear relies on the assumptions that, in the ab- sence of lensing, the ellipticities of the galaxies are uncorrelated. However, an intrinsic correlation of galaxy shapes could exist owing to the coupling of the galaxy angular momentum or shape to the tidal field or to galaxy interactions (see Heavens 2001 for a review). Theoretical estimation of the size of this effect has been performed using numerical simulations (Croft & Metzler 2000; Heav- ens, Refregier & Heymans 2000; Jing 2002) and analytical methods (Catalan, Kamionkowski & Blandford 2001; Crittenden et al. 2001a,b; Lee & Pen 2001; Mackey, White & Kamionkowski 2001). Measurements of intrinsic correlations have also been performed (Brown et al. 2000; Pen, Lee & Seljak 2000). Although considerable uncertainty remains regarding the amplitude of this effect, a consen- sus is arising that intrinsic correlations are likely to be small for the deep current < surveys with zm ∼ 1, but may be dominant for shallower surveys with zm ∼ 0.2 (see however the conflicting results of Jing 2002). Although lensing distortions are coherent over a large redshift range, intrinsic alignments are only significant for small physical separations. Photometric redshifts can thus be used to separate and reduce intrinsic correlations from cosmic-shear signals (Heymans & Heavens 2002; King & Schneider 2002a,b). Another approach is to search for the B -type correlation signal produced by intrinsic correlations (Crittenden et al. 2001b). 7.7 Theoretical Uncertainties ∼ 10′ ) and Most of the signal in cosmic-shear surveys arises from small scales (θ < thus from nonlinear structures (Jain & Seljak 1997). The existing prescriptions for computing the nonlinear corrections (Peacock & Dodds 1997; Ma 1998) to the matter power spectrum are only accurate to approximately 10% and disagree at that level with one another (see discussion in Huterer 2001). This theoretical uncertainty will soon become one of the dominating errors in the determination of cosmological parameters from cosmic-shear surveys (see discussion in van Waer- beke et al. 2002). New prescriptions based on more recent N-body simulations such as those by Smith et al. (2003) will help improve the accuracy of the pre- dictions. The problem of the prediction of higher-order statistics is even more difficult, but it is not as pressing given the large uncertainties in the the measure- ments within current surveys (see Bernardeau, Mellier, van Waerbeke 2002). The inclusion of second-order terms in the weak-lensing approximation will eventually also be required (Cooray 2002, Cooray & Hu 2002). These systematic effects must be controlled to match the statistical accuracy of future surveys. Along with the studies suggested above, further measurements with different colors and comparison of various surveys in overlapping regions would help to control systematics and to test data analysis pipelines. On the theoretical side, larger simulations coupled with advanced analytical techniques will be required. Weak Gravitational Lensing 8 CONCLUSIONS 21 Cosmic shear has emerged as a powerful method to measure the large-scale struc- ture in the universe. It can be thought of as the measurement of background fluctuations in the space-time metric. Although other methods rely on assump- tions relating the distribution of light to that of the mass, weak lensing is based on “clean” physics and can be directly compared to theory. The past three years have yielded impressive observational progress, as the first statistical detections and measurements of cosmic shear have been achieved. In analogy with the CMB, cosmic shear has moved from the COBE era to that of the first generation of anisotropy experiments. However, the measurement of cosmic shear differs from that of CMB anisotropies in several respects. First, the fluctuations are on the order of 10−2 as opposed to 10−5 , making them easier to measure while retaining the validity of linear calculations. Second, the cosmic-shear field is non-Gaussian and therefore contains more information than that quantified by the power spectrum. Existing cosmic-shear measurements have started to yield significant constraints on cosmological parameters. The measurement of the amplitude of the matter power spectrum σ8Ω0.5 m from cosmic shear should soon replace that derived from the local abundance of clusters. This latter technique is indeed limited by the finite number of bright clusters and by systematic uncertainties in the physics of clusters. With better statistics, the angular and redshift dependence of the shear signal as well as with higher-order moments of the convergence field will break the degeneracy between σ8 and Ωm and yield constraints on further parameters such as ΩΛ and Γ. The current measurements of cosmic shear are now primarily limited by statis- tics. They will therefore be improved upon by a number of upcoming and future instruments such as Megacam, VST, VISTA, LSST, and Pan-STARRS from the ground, and HST/ACS and SNAP in space. For several of these instruments, a weak-lensing survey has been listed as one of the primary science drivers. These surveys will potentially yield measurements of cosmological parameters that are comparable in precision and complementary to those derived from the CMB. They will also be able to address more far-reaching questions in cosmology by measuring parameters beyond the standard model. For instance, they can be used to provide a test of the gravitational instability paradigm, a measure of the equation of state of the dark energy, and a test of general relativity. For these instruments to yield the full promise of cosmic shear, a number of challenges have to be met. First, observationally, systematic effects, such as the PSF anisotropy and CCD nonlinearities, must be controlled and corrected for. From the theoretical point of view, calculations of the nonlinear power spectrum, of high-order statistics, and of the associated errors must be improved to meet the precision of future measurements. The observational and theoretical efforts re- quired to overcome these difficulties are worthwhile given the remarkable promise that cosmic shear offers to cosmology. Acknowledgments The author thanks his collaborators David Bacon, Richard Massey, Tzu-Ching Chang, Jason Rhodes, and Richard Ellis for numerous fruitful discussions. The author thanks Bhuvnesh Jain, David Bacon, and Henk Hoekstra for their permission to reproduce their figure. He also thanks Tony Tyson, Herv´e Aussel and Andy Taylor for precisions regarding the parameters of future cosmic shear surveys. He is also grateful to Richard Ellis, Henk Hoekstra, Gary Bern- 22 Refregier stein, Yannick Mellier, Richard Massey, and Ivan Valtchanov for useful comments on the manuscript. The Annual Review of Astronomy and Astrophysics is online at http://astro.annualreviews.org Literature Cited Babul, A Lee MH. 1991. MNRAS 250:407 Bartelmann M, King LJ, Schneider P. 2001. Astron. Astrophys. 378:361 Bartelmann M, Schneider P. 1999. astro-ph/9912508 Bacon DJ, Refregier A, Clowe D, Ellis R. 2001. MNRAS 325:1065 Bacon DJ, Refregier A, Ellis R. 2000. MNRAS 318:625 Bacon DJ, Massey R, Refregier A, Ellis R. 2002. astro-ph/0203134 Benabed K, Bernardeau F. 2001. Phys. Rev. D 64:083501 Bennet DP, Rhie SH. 2000. GEST Home Page. http://bustard.phys.nd.edu/GEST. astro-ph/0011466 Bernardeau F. 1997. Astron. Astrophys. 324:15 Bernardeau F. 1998. Astron. Astrophys. 338:767 Bernardeau F. 1999. Theoretical and Observational Cosmology. Proc. Cargese Summer Sch., Cargese, France, ed. M Lachieze-Rey. astro-ph/9901117 Bernardeau F, Mellier Y, van Waerbeke L. 2002. Astron. Astrophys. 389:L28 Bernardeau F, Valageas P. 2000. Astron. Astrophys. 364:1 Bernardeau F, van Waerbeke L, Mellier Y. 1997. Astron. Astrophys. 322:1 Bernardeau F, van Waerbeke L, Mellier Y. 2003. Astron. Astrophys. 397:405 Berstein GM, Jarvis M. 2001. astro-ph/0107431 Blandford RD, Saust AB, Brainerd TG, Villumsen JV. 1991. MNRAS 241:600 Bonnet H, Mellier Y. 1995. Astron. Astrophys. 303:331 Borgani S, Rosati P, Tozzi P, Stanford SA, Eisenhardt PR, et al. 2001. Ap. J. 561:13 Boulade O, Charlot X, Abbon P, Aune S. Borgeaud P. et al. 2000. Proc. SPIE 4008:657 Megacam Home Page. http://www-dapnia.cea.fr/Phys/Sap/Activites/Projets/Megacam/page.shtml Brown M, Taylor AN, Hambly N, Dye S. 2000. astro-ph/0009499 Brown M, Taylor AN, Bacon, D.J., Gray, M.E., Dye S., Meisenheimer, K., Wolf, C. 2003. MNRAS 341:100 Catalan P, Kamionkowksi M, Blandford R. 2001. MNRAS 320:7 Chang T-C, Refregier A. 2002. Ap. J. 570:447 Contaldi, C, Hoekstra, H., Lewis, A. 2003. submitted to Phys. Rev. Letters. astro-ph/0302435 Cooray A. 2002. astro-ph/0206068 Cooray A, Hu W. 2001a. Ap. J. 548:7 Cooray A, Hu W. 2001b. Ap. J. 554:56 Cooray A, Hu W. 2002. Ap. J. 574:19 Cooray A, Hu W, Miralda-Escud´e J. 2000. Ap. J. 535:9 Cooray A, Kesden M. 2002. astro-ph/0204068 Crittenden R, Natara jan P, Pen U, Theuns, T. 2001a. Ap. J. 559:552 Crittenden R, Natara jan P, Pen U, Theuns T. 2001b. astro-ph/0012336 Croft RAC, Metzler CA. 2000. Ap. J. 545:561 Eke V, Cole S, Frenk C, Patrick Henry J. 1998. MNRAS 298:1145 Erben T, van Waerbeke L, Bertin E, Mellier Y, Schneider P. 2001. Astron. Astrophys. 3667:17 Fort B, Mellier Y. 1994. Astron. Astrophys. Rev. 5:239 Gray ME, Taylor AN, Meisenheimer K, Dye S, Wolf C, Thommes E. 2002. Ap. J. 568:141 Gunn JE. 1967. Ap. J. 150:737 Hamana T, Colombi S, Mellier Y. 2000. Cosmological Physics with Gravitational Lensing. Proc. XXth Moriond Astrophys. Meet., Les Arcs, France, ed. J-P Kneib, Y Mellier, M Mon, J Tran Thanh Van. astro-ph/0009459 Hamana T, Miyazaki S, Shimasaku K, Furusawa H, Doi M, et al. 2002, submitted to Ap. J., preprint astro-ph/0210450 Weak Gravitational Lensing 23 Hammerle H, Miralles JM, Schneider P, Erben T, Fosbury RA. 2002. Astron. Astrophys. 385:743 Heavens AF. 2001. Intrinsic Galaxy Alignments and Weak Gravitational Lensing. Yale Worksh. Shapes Galaxies Haloes, May. astro-ph/0109063 Heavens A, Refregier A, Heymans C. 2000. MNRAS 319:649 Heymans C, Heavens A. 2002. astro-ph/0208220 Hirata C, Seljak U. 2002. astro-ph/0209489 Hirata C, Seljak U. 2003. to appear in MNRAS. astro-ph/0301054 Hiroyasu A, et al. 2001. Subaru Home Page. http://www.subaru.naoj.org/ Hoekstra H, Franx M, Kuijken K, Squires G. 1998. Ap. J. 504:636 Hoekstra H, Franx M, Kuijken K, Carlberg RG, Yee HKC, et al. 2001. Ap. J. 548:5 Hoekstra H, Yee HKC, Gladders, M. 2001. Ap. J. 558:11 Hoekstra H, Yee HKC, Gladders M. 2002a. Ap. J. 577:595 Hoekstra H, Yee HKC, Gladders M. 2002b. astro-ph/0205205 Hoekstra H, Yee HKC, Gladders M, Felipe Barrientos L, Hall PB, Infante L. 2002a. Ap. J. 572:55 Hoekstra H, van Waerbeke L, Gladders MD, Mellier Y, Yee HKC. 2002b. Ap. J. 577:604 Hu W. 1999. Ap. J. 522L:21 Hu W. 2000. Phys. Rev. D 63:3504 Hu W. 2001. astro-ph/010890 Hu W. 2002. astro-ph/0208093 Hu W, Keeton CR. 2002. astro-ph/0205412 Hu W, Tegmark M. 1999. Ap. J. 514L:65 Hu W, White M. 2000. Ap. J. 554:67 Hui L. 1999. Ap. J. 519:9 Huterer D. 2001. astro-ph/0106399 Jain B, Seljak U. 1997 Ap. J. 484:560. Jain B, Seljak U, White S. 2000. Ap. J. 530:547 Jain B, van Waerbeke L. 2000. Ap. J. 530:L1 Jaroszy´nsky M, Park C, Paczy´nsky B, Gott JR. 1990. Ap. J. 365:22 Jarvis, M., Bernstein, G.M., fisher, P., Smith, D., Jain, B., Tyson, J.A., Wittman, D. 2002,Ap. J. 125:1014 Jing YP. 2002. MNRAS 335:L89 Kaiser N. 1992. Ap. J. 388:272. Kaiser N. 1998. Ap. J. 498:26. Kaiser N. 1999. Weak Lensing by Galaxy Clusters. Boston Lensing Meet. astro-ph/9912569 Kaiser N. 2000. Ap. J. 537:555 Kaiser N, Squires G, Broadhurst T. 1995. Ap. J. 449:460 Kaiser N, Tonry JL, Luppino GA. 2000. PASP 112:768. Pan-STARRS homepage http://pan-starrs.ifa.hawaii.edu/ Kaiser N, Wilson G, Luppino GA. 2000. astro-ph/0003338 Kaiser N, Wilson G, Luppino GA, Dahle H. 1999. astro-ph/99077229 Kaiser N, Wilson G, Luppino G, Kofman L, Gioia I, et al. 1998. astro-ph/9809268 Kamionkoski M, Babul A, Cress CM, Refregier A. 1997. MNRAS 301:1064. astro-ph/9712030 King L, Schneider P. 2002a. astro-ph/0208256 King L, Schneider P. 2002b. astro-ph/0209474 Kristian J, Sachs RK. 1966. Ap. J. 143:379 Kristian J. 1967. Ap. J. 147:864 Kuijken K. 1999. Astron. Astrophys. 352:355 Kuijken K, Bender R, Cappellaro E, Muschielok B, Baruffolo A et al. 2002. ESO Messenger 110: 15; VST homepage http://www.na.astro.it/vst/ Lee MH, Paczy´nsky B. 1990. Ap. J. 357:32 Lee J, Pen U. 2001. Ap. J. 532:5 Luppino GA, Kaiser N. 1997. Ap. J. 475:20 Ma C-P. 1998. Ap. J. 508:5 Mackey J, White M, Kamionkowksi M. 2001. astro-ph/0106364 Massey R., Rhodes J., Refregier A., Albert J., Bacon D. et al. 2003. astro-ph/0304418 Maoli R, van Waerbeke L, Mellier Y, Schneider P, Jain B, et al. 2001. Astron. Astrophys., 368:766 Mellier Y. 1999. Annu. Rev. Astron. Astrophys. 37:127 24 Refregier Mellier Y, van Waerbeke L, Bertin E, Tereno I, Bernardeau F. 2002. astro-ph/0210091 Mellier Y, van Waerbeke L, Maoli R, Schneider P, Jain B, et al. 2001. Cosmic shear surveys. Deep Fields, Proc. Eur. South. Obs., Oct., Garching, Ger. astro-ph/0101130 Mellier Y, van Waerbeke L, Radovich M, Bertin E, Dantel-Fort M. 2000. ESO Proceedings, Mining the Sky, Garching, July 2000, A.J. Banday et al eds. astro-ph/0012059. CFHTLS homepage http://cdsweb.u-strasbg.fr:2001/Science/CFHLS/ Mould J, Blandford R, Villumsen J, Brainerd T, Smail I. 1994. MNRAS 271:31 Miyazaki S, Hamana T, Shimasaku K, Furusawa H, Doi M, et al. 2002. Ap. J. 580:97 Munshi D, Coles P. 2000. astro-ph/0003354 Munshi D, Coles P. 2002. astro-ph/0003481 Munshi D, Jain B. 2000. MNRAS 318:109 Munshi D. Jain B. 2001. MNRAS 322:107 Munshi D, Wang Y. 2002. astro-ph/0206483 Narayan R, Bartelmann M. 1999. In Formation of Structure in the Universe, ed. A Dekel, JP Ostriker, p. 360. astro-ph/9606001 Padmanabhan N, Seljak U, Pen UL. 2002. New Astronomy 8:581 Peacock J, Dodds SJ. 1997. MNRAS 280:L19 Pen U, Lee J, Seljak U. 2000. Ap. J. 543:L107 Pen U, van Waerbeke L, Mellier Y. 2001. astro-ph/0109182 Perlmutter, et al. 2003. SNAP Home Page. http://snap.lbl.gov Pierpaoli E, Scott D, White M. 2001. MNRAS 325:77 Pierpaoli E, Scott D, White M. 2002. submitted to MNRAS. astro-ph/0210567 Premadi P, Martel H, Matzner R, Futamase T. 2001. Ap. J. Suppl. 135:7 Refregier A. 2003. MNRAS 338:35 Refregier A, Bacon DJ. 2003. MNRAS 338:48 Refregier A, Rhodes J, Groth E. 2002. Ap. J. 572:L131 Refregier A, Brown ST, Kamionkowski M, Helfand DJ, Cress CM, et al. 1998. Wide Field Surveys in Cosmology. Proc. XIVth IAP Meet., ed. Y Mellier, S Colombi. Paris. astro-ph/9810025 Refregier A, Massey M, Rhodes J, Ellis R, Albert J, et al. 2003. astro-ph/0304419 Reiprich TH, Bohringer H. 2001. astro-ph/0111285 Rhodes J, Refregier A. Groth E. 2000. Ap. J. 536:79 Rhodes J, Refregier A, Groth E. 2001. Ap. J. 552:L85 Rhodes J., Refregier A., Massey R., Albert, J., Bacon D., et al. 2003, astro-ph/0304417 Schneider P. 1995. Proc. Laredo Adv. Summer Sch., Sept. astro-ph/9512047 Schneider P. 1998. Ap. J. 498:43 Schneider P. 1999. Proc. Perspec. Radio Astron., April, Amsterdam. astro-ph/9907146 Schneider P, Lombardi M. 2002. astro-ph/0207454 Schneider P, van Waerbeke L, Jain B, Kruse G. 1998a. MNRAS 296:873 Schneider P, van Waerbeke L, Kilbinger M, Mellier Y. 2002. Astron. Astrophys 396:1 Schneider P, Weiss A. 1988. Ap. J. 327:526 Schneider P, van Waerbeke L, Mellier Y, Jain B, Seitz S, Fort B. 1998b. Astron. Astrophys. 333:767 Seljak U. 1996. Ap. J. 463:1 Seljak U. 2001. astro-ph/0111362 Seljak U, Zaldarriaga M. 1999. Phys. Rev. D 60:43504 Smith RE, Peacock JA, Jenkins A, White SDM, Frenk CS, et al. 2003. MNRAS 341:1311 Spergel, D., Verde, L., Peiris, H.V., Komatsu, E., Nolta, M.R., et al. 2003, submitted to Ap. J.. astro-ph/0302209 Stebbins A. 1996. astro-ph/9609149 Stebbins A, McKay T, Frieman JA. 1995. Proc. IAU Symposium 173. astro-ph/9510012 Takada M. Jain B. 2002. astro-ph/0205055 Taylor A. 2001. astro-ph/0111605 Taylor A, Watts P. 2000. astro-ph/0010014 Taylor A, et al. 2003. in preparation VISTA Home Page. http://www.vista.ac.uk Tyson JA, Wittman D, Hennawi JF, Spergel DN. 2002a. Proc. 5th Int. UCLA Symp. Sources Detect. Dark Matter, Feb., Marina del Rey, ed. D Cline. astro-ph/0209632 Tyson JA, & the LSST collaboration 2002b. Proc. SPIE Int.Soc.Opt.Eng. 4836, 10-20. astro- ph/0302102. LSST Home Page http://lsst.org Uzan J-P, Bernardeau F. 2001. Phys. Rev. D 64:083004 Weak Gravitational Lensing 25 Valageas P. 2000. Astron. Astrophys. 356:771 Valdes F, Jarvis JF, Tyson JA. 1983. Ap. J. 271:431 van Waerbeke L. 1998. Astron. Astrophys. 334:1 van Waerbeke L, Bernardeau F, Mellier Y. 1999. Astron. Astrophys. 342:15 van Waerbeke L, Mellier Y., Erben T., Cuillandre JC, Bernardeau F, et al. 2000. Astron. As- trophys. 358:30 van Waerbeke L, Mellier Y., Radovich M., Bertin E., Dantel-Fort M. et al. 2001a. Astron. Astrophys. 374:757 van Waerbeke L, Hamana T, Scoccimarro R, Colombi S, Bernardeau F. 2001b. MNRAS 322:918 van Waerbeke L, Mellier Y, Pell´o R, Pen U-L, McCracken HJ, Jain B. 2002. Astron. Astrophys. 393:369 Viana P, Liddle A. 1999. MNRAS 303:535 Viana P, Nichol RC, Liddle A. 2001. astro-ph/0111394 Villumsen J. 1995. astro-ph/9507007 Villumsen J. 1996. MNRAS 281:369 van Waerbeke L, Mellier Y, Tereno I. 2002. astro-ph/0206245 Wambsganss J, Cen R, Ostriker JP. 1998. Ap. J. 494:29 Weinberg NN, Kamionkowski M. 2002. astro-ph/0210134 White M, Hu W. 2000. Ap. J. 537:1 Wilson G, Kaiser N, Luppino GA. 2001. Ap. J. 556:601 Wittman DM, Tyson J, Kirkman D, Dell’Antonio I, Bernstein G. 2000. Nature 405:143 Wittman DM. 2002. Dark Matter and Gravitational Lensing, LNP Top. Vol., ed. F Courbin, D Minniti. Springer-Verlag. astro-ph/0208063 Wittman DM, Tyson JA, Dell’Antonio IP, Becker AC, Margoniner VE, et al. 2002. Proc. SPIE 4836 v.2. astro-ph/0210118. Deep Lens Survey web page http://dls.bell-labs.com/ Zaldarriaga M, Scoccimarro R. 2002. astro-ph/0208075 Zaldarriaga M, Seljak U. 1998. astro-ph/9810257
astro-ph/0501413
2
0501
2005-01-25T15:49:16
Ultraviolet Dust Grain Properties in Starburst Galaxies: Evidence from Radiative Transfer Modeling and Local Group Extinction Curves
[ "astro-ph" ]
This paper summarizes the evidence of the ultraviolet properties of dust grains found in starburst galaxies. Observations of starburst galaxies clearly show that the 2175 A feature is weak or absent. This can be the result of radiative transfer effects (mixing the dust and stars) or due to dust grains which do not have this feature. Spherical DIRTY radiative transfer models imply that it is not radiative transfer effects, but other radiative transfer models with disk/bulge geometries have found cases where it could be radiative transfer effects. Recent work on the extinction curves in the Magellanic Clouds and Milky Way has revealed that the traditional explanation of low metallicity for the absence of the 2175 A feature in the Small Magellanic Cloud is likely incorrect. The SMC has one sightline with a 2175 A feature and the Milky Way has sightlines without this feature. In addition, where the 2175 A feature is found to be weak or absent in both Magellanic Clouds and the Milky Way, there is evidence for recent star formation. Taking the sum of the radiative transfer modeling of starburst galaxies and the behavior of Local Group extinction curves, it is likely that the dust grains in starburst galaxies intrinsically lack the 2175 A feature.
astro-ph
astro-ph
Ultraviolet Dust Grain Properties in Starburst Galaxies: Evidence from Radiative Transfer Modeling and Local Group Extinction Curves Karl D. Gordon Steward Observatory, University of Arizona, Tucson, AZ 85721, USA email: [email protected] Abstract. This paper summarizes the evidence of the ultraviolet properties of dust grains found in starburst galaxies. Observations of starburst galaxies clearly show that the 2175 Å feature is weak or absent. This can be the result of radiative transfer effects (mixing the dust and stars) or due to dust grains which do not have this feature. Spherical DIRTY radiative transfer models imply that it is not radiative transfer effects, but other radiative transfer models with disk/bulge geometries have found cases where it could be radiative transfer effects. Recent work on the extinction curves in the Magellanic Clouds and Milky Way has revealed that the traditional explanation of low metallicity for the absence of the 2175 Å feature in the Small Magellanic Cloud is likely incorrect. The SMC has one sightline with a 2175 Å feature and the Milky Way has sightlines without this feature. In addition, where the 2175 Å feature is found to be weak or absent in both Magellanic Clouds and the Milky Way, there is evidence for recent star formation. Taking the sum of the radiative transfer modeling of starburst galaxies and the behavior of Local Group extinction curves, it is likely that the dust grains in starburst galaxies intrinsically lack the 2175 Å feature. INTRODUCTION There are two main reasons to study the dust in galaxies where single stars are unresolv- able with current telescopes. 1) To directly study the dust itself as galaxies provide a much wider range of envi- ronments than found in the Local Group. These different environments can be charac- terized by more extreme values of radiation field density, radiative field hardness, shock frequency, and initial conditions (eg., metallicity). Such different environments can sig- nificantly impact the formation and destruction of dust grains. 2) To more accurately account for the effects of dust on observations of the stars and gas present in galaxies. This accounting allows for the study of stars and gas in galaxies to be carried out more accurately. In order to study dust in such galaxies, a model of how the mixing of dust with stars and gas affects to the observations of a galaxy is required. One such model is the DIRTY dust radiative transfer model [1, 2] which uses Monte Carlo techniques to compute the radiative transfer of photons through dust and self-consistently accounts for the dust re-emission in the infrared including equilibrium and non-equilibrium emission. This paper concentrates on the ultraviolet dust grain properties in starburst galaxies determined using the DIRTY model and implied from extinction curves measured in the Local Group. The study of dust in galaxies is much larger than this, especially in light of the wealth of information available in the infrared where the dust grain emission dominated the spectral energy distribution of a galaxy. The connection between ultraviolet and infrared dust grain properties is not direct, but determined using dust grain models [3, 4, 5] which have their own problems. For example, I have found that using empirically determined dust scattering properties in the ultraviolet [6] is crucial to reproducing the ultraviolet colors of starburst galaxies. STARBURST GALAXIES Starburst galaxies provide ideal environments to probe dust grain properties in extreme environments as they are intrinsically bright in the ultraviolet where dust grain properties show large variation [7, 8, 9]. The environments probed by starbursts range from very metal poor (eg., I Zw 18) to very metal rich and most starbursts are likely characterized by high and hard radiation fields and elevated shock frequencies. Starbursts probe the more active environments of those possible in galaxies in general. The study of the type of dust found in starburst galaxies really got going with the work of Calzetti et al. [10] and Calzetti [11]. In these studies, a variant of the standard pair method was used to derive an empirical attenuation curve appropriate for ultraviolet bright starburst galaxies. This empirical curve clearly lacked the strong 2175 Å absorp- tion feature seen in almost all known dust extinction curves. In addition, this curve was grayer than most dust extinction curves. The question was then: Is the lack of this 2175 Å feature and the grayer curve due to different dust grain properties or radiative transfer effects? At the time, only a small number of sightlines in the Small Magellanic Cloud (SMC) were known to lack this feature [12]. This lack was attributed to the low metal- licity of the SMC with the higher metallicities of the Large Magellanic Cloud and the Milky Way resulting in the ubiquity of the 2175 Å feature in their dust [7, 8]. But, the starburst galaxies studied by Calzetti et al. [10] had metallicities spanning the known range from lower than the SMC to higher than the Milky Way. Radiative transfer effects were seen as the probable cause of the lack of the 2175 Å feature. The complicated mix- ing of the dust, gas, and many stars in unresolved observations of galaxies implies that radiative transfer effects are important. Such mixing means that the stars in the observ- ing beam have different dust columns and significant scattered flux is included in any measurement. As a result, the attenuation curve for a galaxy is not dependent only on dust grain properties, like that for the extinction curve towards a single resolved star in the Local Group, but is also dependent on the geometry of the dust, gas, and stars. Motivated by the starburst results and improvements in the DIRTY dust radiative transfer model to include nonhomogeneous dust distributions [13], Gordon et al. [14] investigated the importance of radiative transfer effects on determining the shape of the empirical starburst attenuation curve. After investigating a number of dust/star geome- tries, they found that the lack of the 2175 Å feature could not be explained by radiative transfer effects. The only way to remove the 2175 Å feature from the attenuation curve was for the dust grains themselves to lack the feature. This result coupled with the large range in metallicities in the starburst galaxies implied that it was the starburst environ- ment which was responsible for the lack of the 2175 Å feature. The dust grains were either newly formed without the feature or this feature was destroyed by the hard radia- tion field and elevated shock frequency in starbursts. Using the same or similar analysis, this result was extended to higher redshifts (z ≈ 3) by Gordon et al. [15] and Vijh et al. [16] where it was also found that most high-redshift (and likely starburst) galaxies lack a 2175 Å bump. A detailed analysis of the Calzetti [11] attenuation curve by Witt and Gordon [17] found that a clumpy SHELL geometry with SMC dust, t V ∼ 1.5 DIRTY model best fit both the grayness of the curve and the lack of the 2175 Å feature. Since the work utilizing simple spherical DIRTY models, other studies have found ways to produce attenuation curves with weak or nearly absent 2175 Å features while using dust grains with Milky Way-like dust (i.e., a strong 2175 Å feature). Using the GRASIL code, Granato et al. [18] found that differentially embedding sources of different spectral types (basically by stellar lifetime) in molecular clouds in a disk/bulge geometry with clumpy structure and Milky Way-like dust resulted in attenuation curves with weak or absent 2175 Å features. Fischera et al. [19] studied the attenuation curves produced by a foreground screen with a log-normal density distribution of Milky Way- like dust and was able to reproduce the grayness of the Calzetti [11] curve, but was only able to reduce the strength of the 2175 Å feature. Finally, Pierini et al. [20] used DIRTY models of disk/bulge galaxies and found that for highly inclined, weak bulge galaxies the 2175 Å feature was very weak in the global attenuation curves. As a result of these studies [18, 19, 20], it is clear that there are dust/star geometries which can reproduce most, if not all, of the characteristics of the Calzetti [11] starburst curve. But it is not clear that the necessary geometries (disk/bulge or foreground screen) are appropriate for the starburst galaxies which are usually seen as nuclear dominated or have highly irregular morphologies. LOCAL GROUP EXTINCTION CURVES An important part of the type of dust we expect to find in galaxies is driven by our detailed knowledge of dust in the Local Group. One of the most powerful ways to study the properties of dust is to determine ultraviolet through near-infrared extinction curves towards single stars. As the dust measured by observing a star is distributed in a foreground screen, the derived extinction curve is only dependent on the properties of the dust grains (size, composition, and shape). With current telescopes and instruments, this work is limited to Local Group galaxies. Prior to the work on starburst galaxies, it was thought that most galaxies would have dust with a 2175 Å feature because all known sightlines in the Milky Way [7, 8] and the Large Magellanic Cloud [21, 22] had this feature. Only in the Small Magellanic Cloud was the 2175 Å feature absent [12] and this was understood to be due to the low metallicity of this galaxy. The lack of the 2175 Å feature in starburst galaxies with metallicities like those found in the Milky Way and LMC questioned the explanation for the SMC lacking the 2175 Å feature. In fact, the variation in the LMC extinction curves between those near and far from the 30 Dor star forming region pointed to a different explanation [22]. It was found that the extinction curves near 30 Dor had weaker 2175 Å features and somewhat steeper far-ultraviolet extinctions than those found in the rest of the LMC. The existence of two explanations (low metallicity versus nearby active star forma- tion) for the weakness or lack of a 2175 Å feature motivated new work on extinction curves in the Local Group. In the SMC, Gordon and Clayton [23] searched the IUE archives and found it was possible to generate accurate extinction curves for only four sightlines. Three of the four curves were basically linear with 1/l and had no detectable 2175 Å feature. The fourth curve was much more like that found in the Milky Way with a 2175 Å feature and weaker far-ultraviolet extinction compared to the rest of the SMC curves. This Milky Way-like curve was first published in Lequeux et al. [24] but dismissed by Prevot et al. [12] as anomalous. Gordon and Clayton [23] found that this Milky Way-like curve was located in a more quiescent region than the other three curves which were located in the star forming bar of the SMC. One additional sightline was added to the sample of SMC extinction curves by Gordon et al. [9] using new HST/STIS observations. This new sightline was in the star forming bar and was very similar to the previously known extinction curves in this same region. Thus, the spatial variations in the five known extinction curves in the SMC are consistent with the explanation that the lack of the 2175 Å feature is due to nearby active star formation. In the LMC, Misselt et al. [25] searched the IUE archives and was able to construct 19 accurate extinction curves from the available data. Like previous studies, this work found significant differences between those sightlines near 30 Dor and those in the rest of the LMC. Misselt et al. [25] found that this difference was caused by the sightlines associated with the supergiant shell LMC 2 near 30 Dor, not 30 Dor itself. Thus, they found very significant differences (especially in their 2175 Å feature strengths) when the 19 sightlines were grouped between those associated with the LMC 2 shell and those in the rest of the LMC. Like the SMC, the spatial variation of the LMC extinction curves is consistent with active star formation being the cause. In the Milky Way, the most comprehensive extinction curve study to date has been the recent work of Valencic et al. [26]. Like the studies in the SMC and LMC [23, 25], the IUE archive was searched and a total of 417 extinction curves created. The overwhelming majority of these curves (93%) were well characterized by the Cardelli et al. [7] RV dependent relationship with only 4 curves showing systematic deviation from this relationship. HD 204827 is one of these 4 curves and a detailed study of this sightline [27] found that after subtracting a well measured foreground component, the remaining extinction curve was equivalent within the uncertainties to the extinction curves found in the SMC star forming bar (i.e., linear with 1/l and lacking a 2175 Å bump). The local environment of HD 204827 has evidence for a recent supernova shock. In a study of 30 low reddening, long sightlines in the Milky Way, Clayton et al. [28] found that a subsample which had extinction curves like that found in the LMC 2 shell sample [25]. This subsample were all in the same region of the Milky Way and displayed N(Ca II)/N(Na I) ratios and velocities indicating recent dust destruction. Like the work in Magellanic Clouds, new studies in the Milky Way have revealed evidence that processing of dust can cause the weakening or disappearance of the 2175 Å feature and strengthening of the far-ultraviolet extinction. Examining the sum of work on extinction curves in the Milky Way and Magellanic Clouds, it is clear that there is a continuum of dust extinction curves extending from those represented by the Cardelli et al. [7] relationship with strong 2175 Å features to those like those found in the SMC bar with no detectable 2175 Å feature [9]. This implies that it is more accurate to describe the Cardelli et al. [7] relationship as referring to dust in quiescent environments and SMC bar extinction curves as referring to dust in much more active environments. But even this picture of dust ultraviolet extinction curve variations is incomplete as dust in molecular clouds has not been well measured. Only the Taurus molecular cloud has been probed in the ultraviolet through the HD 29647 and HD 283809 sightlines [29]. By subtracting a foreground extinction curve measurement from these two sightlines, Whittet et al. [30] found that the 2175 Å bump disappeared. This is like the dust found in the SMC bar, but the far-ultraviolet extinction in this dust was much weaker. This is good evidence that the true range of dust extinction curves likely includes another extreme, that found in molecular clouds. DISCUSSION In this paper, I have attempted to summarize work I have been involved in relating to the dust found in starburst galaxies and Local Group extinction curve work motivated in part by the starburst galaxy results. Combining the results for radiative transfer in starburst galaxies for what we now know is a more varied story of dust extinction curves in the Milky Way and Magellanic Clouds, it is most probable that the dust grains in starburst galaxies are truly lacking the 2175 Å feature and possesses strong far-ultraviolet extinctions. It is very important to clearly state that this conclusion only applies to the dust found in or near the starburst regions. This is the dust which is probed by the ultraviolet photons from the starburst regions. Given our knowledge of Local Group dust extinction curves, it is very likely that dust far from sites of active star formation in starburst galaxies would be characterized by Milky Way-like or quiescent dust (i.e., Cardelli et al. [7] relationship). The measurements of starburst galaxies in the ultraviolet necessarily probe only the active regions of these galaxies; there are other measurements which probe the type of dust found in more quiescent regions of galaxies. Studies of the colors of gravitational lens systems [31, 32] have found Cardelli et al. [7] type dust curves. As gravitational lenses probe random sightlines in galaxies, this is evidence that quiescent dust is like that seen in our Galaxy. These gravitational lens difference curves are not direct measures of dust extinction as they probe the difference in two dust extinction curves of unknown dust columns [33, 34], but they can be used to determine the presence of the 2175 Å feature. Finally, measurements of Mg II absorbers also probe random sightlines in galaxies and show evidence for the 2175 Å feature [35, 36]. ACKNOWLEDGMENTS I would like thank the organizers of this conference for inviting me to give a talk on my work. The work described in the paper encompasses a large body of work, which would not have been possible but for the contributions of many of my collaborators. I would like to especially thank Adolf Witt who was my Monte Carlo radiative transfer mentor and Geoff Clayton who was my extinction curve mentor. REFERENCES 1. Gordon, K. D., Misselt, K. A., Witt, A. N., and Clayton, G. C., ApJ, 551, 269 (2001). 2. Misselt, K. A., Gordon, K. D., Clayton, G. C., and Wolff, M. J., ApJ, 551, 277 (2001). 3. Weingartner, J. C., and Draine, B. T., ApJ, 548, 296 (2001). 4. Clayton, G. C., Wolff, M. J., Sofia, U. J., Gordon, K. D., and Misselt, K. A., ApJ, 588, 871 (2003). 5. 6. Gordon, K. D., "Interstellar Dust Scattering Properties," in ASP Conf. Ser. 309: Astrophysics of Dust, Zubko, V., Dwek, E., and Arendt, R. G., ApJS, 152, 211 (2004). Fitzpatrick, E. L., and Massa, D., ApJS, 72, 163 (1990). 7. Cardelli, J. A., Clayton, G. C., and Mathis, J. S., ApJ, 345, 245 (1989). 8. 9. Gordon, K. D., Clayton, G. C., Misselt, K. A., Landolt, A. U., and Wolff, M. J., ApJ, 594, 279 (2003). 10. Calzetti, D., Kinney, A. L., and Storchi-Bergmann, T., ApJ, 429, 582 (1994). 11. Calzetti, D., AJ, 113, 162 (1997). 12. Prevot, M. L., Lequeux, J., Prevot, L., Maurice, E., and Rocca-Volmerange, B., A&A, 132, 389 2004, p. 77. (1984). 13. Witt, A. N., and Gordon, K. D., ApJ, 463, 681 (1996). 14. Gordon, K. D., Calzetti, D., and Witt, A. N., ApJ, 487, 625 (1997). 15. Gordon, K. D., Smith, T. L., and Clayton, G. C., "Dust in High Redshift Starburst Galaxies," in ASP Conf. Ser. 193: The Hy-Redshift Universe: Galaxy Formation and Evolution at High Redshift, 1999, p. 517. 16. Vijh, U. P., Witt, A. N., and Gordon, K. D., ApJ, 587, 533 (2003). 17. Witt, A. N., and Gordon, K. D., ApJ, 528, 799 (2000). 18. Granato, G. L., Silva, L., Bressan, A., Lacey, C. G., Baugh, C. M., Cole, S., and Frenk, C. S., Astrophysics and Space Science Supplement, 277, 79 (2001). 19. Fischera, J., Dopita, M. A., and Sutherland, R. S., ApJL, 599, L21 (2003). 20. Pierini, D., Gordon, K. D., Witt, A. N., and Madsen, G. J., ApJ, in press (2004). 21. Clayton, G. C., and Martin, P. G., ApJ, 288, 558 (1985). 22. Fitzpatrick, E. L., AJ, 92, 1068 (1986). 23. Gordon, K. D., and Clayton, G. C., ApJ, 500, 816 (1998). 24. Lequeux, J., Maurice, E., Prevot-Burnichon, M.-L., Prevot, L., and Rocca-Volmerange, B., A&A, 113, L15 (1982). 25. Misselt, K. A., Clayton, G. C., and Gordon, K. D., ApJ, 515, 128 (1999). 26. Valencic, L. A., Clayton, G. C., and Gordon, K. D., ApJ, 616, 912 (2004). 27. Valencic, L. A., Clayton, G. C., Gordon, K. D., and Smith, T. L., ApJ, 598, 369 (2003). 28. Clayton, G. C., Gordon, K. D., and Wolff, M. J., ApJS, 129, 147 (2000). 29. Clayton, G. C., Gordon, K. D., Salama, F., Allamandola, L. J., Martin, P. G., Snow, T. P., Whittet, D. C. B., Witt, A. N., and Wolff, M. J., ApJ, 592, 947 (2003). 30. Whittet, D. C. B., Shenoy, S. S., Clayton, G. C., and Gordon, K. D., ApJ, 602, 291 (2004). 31. Nadeau, D., Yee, H. K. C., Forrest, W. J., Garnett, J. D., Ninkov, Z., and Pipher, J. L., ApJ, 376, 430 (1991). 32. Falco, E. E., Impey, C. D., Kochanek, C. S., Lehár, J., McLeod, B. A., Rix, H.-W., Keeton, C. R., Muñoz, J. A., and Peng, C. Y., ApJ, 523, 617 (1999). 33. Wucknitz, O., Wisotzki, L., Lopez, S., and Gregg, M. D., A&A, 405, 445 (2003). 34. McGough, C., Clayton, G. C., Gordon, K. D., and Wolff, M. J., ApJ, in press (2005). 35. Malhotra, S., ApJL, 488, L101 (1997). 36. Wang, J., Hall, P. B., Ge, J., Li, A., and Schneider, D. P., ApJ, 609, 589 (2004).
astro-ph/9307005
1
9307
1993-07-05T11:34:00
The Environment of Lyman a Absorbers in the Sightline toward 3C273
[ "astro-ph" ]
We present new ground-based data following up on the HST discovery of low-redshift Lya absorption in the sight-line to the quasar 3C273. Narrow-band filter observations show that there are no H II regions within a 12 kpc radius of the line-of-sight to the quasar, at the velocities of three of the absorbers. Broad-band imaging shows that there are no dwarf galaxies at Virgo distances with absolute magnitude above MB~-13.5 and within a radius of 40 kpc. We present fiber spectroscopy of galaxies within a radius of 1 deg, down to an apparent magnitude of B~19. We show that the absorbers are definitely not distributed at random with respect to the galaxies, but also that the absorber-galaxy correlation function is not as strong as the galaxy-galaxy correlation function on large scales. Our data are consistent with the hypothesis that all galaxies more luminous than 1/10 L* have effective cross-sections (for association with absorbers with Log(NH)>13.0), of between 0.5 and 1 Mpc. We also show a clear case of a Lya absorber which has no galaxy brighter than MB=-18 within a projected distance of 4.8 Mpc, and discuss the possibility that Lya absorbers are destroyed in a rich galaxy environment.
astro-ph
astro-ph
The Environment of Lyman α Absorbers in the Sightline towards 3C273 S. L. Morris1, R. J. Weymann Alan Dressler and P. J. McCarthy The Observatories of the Carnegie Institution of Washington, 813 Santa Barbara St., Pasadena, CA 91101 Institute for Astronomy, University of Hawaii, Honolulu, HI 96822 B. A. Smith R. J. Terrile Jet Propulsion Laboratory, Pasadena, CA 91109 Department of Astronomy and National Astronomy and Ionosphere Center2, Cornell University, Ithaca, NY R. Giovanelli 14853 and M. Irwin Royal Greenwich Observatory, Madingley Rd, Cambridge CB3 0EZ, UK ABSTRACT We present new ground-based data following up on the HST discovery of low-redshift Lyman α absorption in the sight-line to the quasar 3C273. Our goal is to investigate the relationship between the low-column-density absorbers and higher column-density objects such as galaxies or H II regions. Narrow-band filter observations with a corono- graph show that there are no H II regions or other strong Hα line-emitting gas within a 12 kpc radius of the line-of-sight to the quasar, at the velocities of three of the absorbers. Broad-band imaging in Gunn r shows that there are no dwarf galaxies at Virgo distances with absolute magnitude above MB ≈-13.5 and within a radius of 40 kpc from the line- of-sight to the quasar. Finally, we present fiber spectroscopy of a complete sample of galaxies within a radius of 1◦, down to an apparent magnitude of B≈19. Analysis of this sample, combined with galaxies within 10 Mpc of the quasar line-of-sight taken from 1Present Address: Institute of Astronomy, Madingley Rd., Cambridge CB3 0HA, UK 2The National Astronomy and Ionosphere Center is operated by Cornell University under a cooperative agreement with the U.S. National Science Foundation 1 the literature, shows that the absorbers are definitely not distributed at random with respect to the galaxies, but also that the absorber-galaxy correlation function is not as strong as the galaxy-galaxy correlation function on large scales. We show that our data are consistent with the hypothesis that all galaxies more luminous than 1/10 L∗ have effective cross-sections (for association with absorbers whose neutral-hydrogen column- density (Log(NH)) is >13.0), of between 0.5 and 1 Mpc. We also show a clear case of a Lyman α absorber which has no galaxy brighter than MB=-18 within a projected distance of 4.8 Mpc, and discuss the possibility that Lyman α absorbers are destroyed in a rich galaxy environment. Subject headings: cosmology -- interstellar:matter -- galaxies:redshifts 2 1. Introduction Understanding the origin and evolution of struc- ture in the universe remains one of the most funda- mental and active challenges of current astrophysical research. As the evidence in favor of a cosmologi- cal origin for the narrow, displaced absorption lines in QSO spectra became overwhelming, it also became clear that both the metal-line systems and the Lyman α systems are invaluable tools for the study of some aspects of this problem. Since ground-based Lyman α studies refer only to redshifts ∼>1.6, they complement studies of galaxy clustering properties, the majority of which involve redshifts much less than this. How- ever, precisely because the redshift regimes have been so different and because it has not been at all clear what relation exists between the typical low-column- density Lyman α absorbers and galaxies, these two approaches have remained disjoint. It was somewhat unexpected, but pleasing, that low-redshift Lyman α absorbers were found in sufficient numbers to enable meaningful studies of the evolution of the Lyman α absorbers and their relation to galaxies (Morris et al. 1991, Bahcall et al. 1991b). This has presented as- tronomers with the opportunity to join these two lines of investigation. There are two levels at which such such attempts can be carried out: 1) Purely statistical investiga- tions aimed at comparing the clustering properties of galaxies and Lyman α absorbers, and 2) Investi- gation of individual cases in which the possibility of establishing the presence (or absence) of a clear link between a Lyman α absorption line and something we could call a "galaxy" presents itself. Preliminary discussions along these lines may be found in papers by Bahcall et al. 1992a, Bahcall et al. 1992b and by Salzer 1992. The present paper is a first attempt to pursue both these approaches along the sightline to 3C273. The remainder of this paper is organized as follows: In § 2. we describe the different sets of observations we have assembled to investigate the en- vironment of the Lyman α absorbers along the 3C273 sightline. In § 3. we analyze them for possible associ- ations or lack of associations of individual Lyman α absorbers with galaxies, and also give some statisti- cal analysis of the clustering properties of the Lyman α absorbers with galaxies. In § 4. we discuss these results in light of current models for the Lyman α ab- sorbers and provide a brief summary and suggestions for further work. 3 Throughout this paper H0 is taken to be 80 km/s/Mpc, the distance to the Virgo cluster is taken to be 16.0 Mpc (a distance modulus of (m-M)Virgo=31.02) (Ja- coby et al. 1992), and q0 is taken to be 0. 2. Observations and Reduction It has long been realized that imaging of the gas directly responsible for the low-column-density Ly- man α absorbers is well beyond the reach of current technology. Specifically, the neutral-hydrogen column densities of order 1013 − 1014 cm−2 that GHRS de- tected toward 3C273 are about 4 or 5 orders of mag- nitude below what can be imaged in 21 cm emission, even without taking into account the powerful radio background contributed by 3C273 itself. The Hα re- combination surface brightness associated with these neutral-hydrogen column-densities is also several or- ders of magnitude below what is feasible to detect, unless the incident flux of ionizing photons is several orders of magnitude higher than that expected from the integrated background radiation. However, it has frequently be suggested that the Lyman α absorbers are intimately connected with, or are actual extensions of, entities which can be imaged either by means of Hα emission, starlight, or 21 cm emission - e.g. dwarf galaxies (Fransson and Epstein 1982) or shells of expanding gas (Chernomordic and Ozernoy 1983) or the outer regions of galactic disks (Maloney 1992). In the case of dwarf irregulars, for example, a very small episode of recent star formation might betray the presence of a dwarf irregular whose outer envelope produces the Lyman α absorbers. Al- ternatively, expanding shells of gas might produce Hα emission via collisional ionization at a shock front. In addition, of course, possible association of indi- vidual Lyman α absorbers with specific galaxies, as well as statistical studies of absorber-galaxy correla- tion can be carried out with a sample of redshifts for galaxies in the field surrounding 3C273. In this section we describe three such new sets of observations of a region centered on 3C273. These are: § 2.1. Coronagraph observations with narrow- band filters of a 5′ diameter region, § 2.2. Deep broad- band imaging of a 17′ diameter region, and § 2.3. Fiber spectroscopy of a 2.2×1.6◦ region down to a limiting magnitude of B=19. We describe the analysis of these three sets of observations in § 3. 2.1. Coronagraph Observations with Narrow Band Filters Observations of a 5.′3×5.′3 region (radius ≈12 kpc at Virgo) around 3C273 were obtained during 1992 February 3-7, with the University of Hawaii Corona- graph (Vilas and Smith 1987) on the Las Campanas 2.5m duPont Telescope. A thinned 1024×1024 Tek- tronics CCD was used, binned 2×2, giving a scale of 1.23′′/pixel. The coronagraph blocking mask had a diameter of 5′′. Data were obtained with a Gunn r fil- ter, and also 5 specially acquired filters, 3 with width 13.5A, centered at 6586.2A, 6598.0A and 6756.0A (hereafter referred to as VN1, VN2 and HN), and 2 with width 25A, centered at 6643.2A and 6718.9A (hereafter referred to as VB and HB). (The above widths and central wavelengths are quoted for an f7.5 beam and a temperature of 15◦ C). The narrow-band wavelengths were chosen to match the redshifted po- sition of Hα at the velocities of the 2 Lyman α ab- sorbers listed in Morris et al. 1991 at velocities corre- sponding to the Virgo cluster (Bingelli, Sandage and Tammann 1985), and one absorber at 1251A. This is the lowest redshift, strong Lyman α system beyond the Virgo cluster. The observing procedure involved cycling through the 6 different filters, with exposure times of 10 mins for the r and 25A filters, and 20 mins for the 13.5A filters. Seven such cycles were completed over the 5 night run. During the observing run, it was discovered that the narrow-band filter HN had slightly non-parallel faces, resulting in detectable 'ghost' images offset from bright stars and also 3C273. In an attempt to minimise the effect of these, this fil- ter was rotated though 90◦ between each night. For calibration purposes, observations were also obtained of Mkn 49 (an emission line galaxy in the Virgo cluster with radial velocity 1524 km/s, and hence with Hα line-emission within 2A of the peak of VN2) and M87, and also a number of bright stan- dard stars. The images were reduced using IRAF3. The reduc- tion steps were: bias subtraction, division by a flat field taken on the same night as the observations, ro- tation and shifting of the images to match a reference image, sky subtraction and averaging together of im- ages taken with the same filter. It was found after the run that refocusing the coronagraph between taking the flat fields and the data meant that the flat field di- 3IRAF is distributed by NOAO, which is operated by AURA Inc. under contract to the NSF vision left significant features in the data, both at the edges of the coronagraph field and also throughout the data at the location of what are presumed to be dust particles on the coronagraph optics or CCD win- dow. This problem was particularly noticeable for the data taken on the last night of the run. However, the residuals are greatly reduced in the combined data. Continuum sources were removed from the narrow- band images by subtracting off a scaled version of the two 25A filter observations. We investigated scal- ing methods, including measuring stars or galaxies in the images, but found that the scaling derived was consistent with simply subtracting off the average of the 25A filters from each 13.5A observation. That is, there was no evidence for a significant continuum slope or calibration difference across the 150A region of interest, and the factor 2 shorter broad-band expo- sure fairly accurately balanced the higher throughput of the 25A filters. This apparent consistency may be fortuitous: due to the small field (small number of galaxies), and the errors in measuring the flux of faint galaxies. Figure 1 shows the sum of the 25A data, and the continuum subtracted 13.5A data for 3C273. As can be seen, the PSF is a function of angle off-axis, and becomes quite broad and asymmetric at the edge of the field, due to aberrations in the coronagraph optics. 2.2. Wide Field Imaging with COSMIC A mosaic of Gunn r band images of the field around 3C273 was obtained on 1992 February 23, at the Palomar 5m, with the prime focus COSMIC sys- tem. This recently commissioned camera contains a thinned 2048×2048 Tektronics CCD. Due to poor seeing, this was read out with 2×2 binning giving a scale of 0.56′′/pixel. Exposures were taken roughly centered on 4 positions offset from 3C273 by 5.′4 NE, NW, SE and SW. Four exposures of 5 minutes each were obtained at each position. During the observa- tions, the moon rose, and so the background level of the images varies by almost a factor 2. The data were reduced using IRAF. After bias subtraction, the data were flat fielded using images of the dome. This left significant large scale struc- ture in the images, and so a 'skyflat' was constructed from the combined data images. This flat was me- dian smoothed to remove any small scale structure. Because of a region of very low sensitivity near the center of the CCD, and also the location of a very bright star coincidentally at the same place in one set 4 of images, the flat fielded data still shows a weak neg- ative feature near the center of each image. The sky background was determined for each image separately and subtracted. The offsets between the images were determined by measuring the positions of the QSO and 2 bright nearby stars, which were common to all images and the data were then mosaiced together, giving a combined image with diameter 17.′2 centered on 3C273. The resulting image is show as figure 2. 2.3. Fiber Spectroscopy During 1992 February 8-10, spectra of objects in a 2.2×1.6◦ rectangle surrounding 3C273 were obtained with the Fiber Spectrograph at the Las Campanas 2.5m duPont Telescope. (Shectman 1992). This sys- tem has 128 fibers which are manually plugged into an aluminum plate over a 1.5×1.5◦ field. The fibers have a projected diameter of 3.′′5. They feed the slit of a floor-mounted spectrograph. A 600 l/mm grating was used, with the 2D-Frutti photon-counting detec- tor, giving a resolution of 8.6A FWHM. Three fields were observed for about 6000 seconds each, offset EW from each other. Objects observed were chosen from a database pro- duced by scanning an UKST IIIaJ plate of the region with the APM scanning machine at Cambridge. The APM produces a catalog of all the objects on a plate, with estimates of isophotal magnitudes, size and a 'sigma' parameter that measures how much the image parameters differ from those of stars with compara- ble magnitude. We chose objects to observe from this catalog with 'sigma'>3.0. This includes many fairly compact objects, and resulted in a rather high con- tamination by stars, but also means that we found a number of compact galaxies that would otherwise have been missed. For each 1.5×1.5◦ fiber field, a magnitude sorted list of candidate galaxies was pro- duced. Due to restrictions on the minimum fiber sep- aration, 330 out of a possible 336 object fibers were used, with 16 fibers per field set on blank sky. The fiber spectra were extracted and reduced using the IRAF apextract package. The spectra were first traced and extracted. Then fiber-to-fiber throughput differences were corrected with a flat field image, and the fibers were wavelength calibrated and rebinned to a common linear wavelength scale. The wavelength calibration used an arc spectrum to determine the non-linear relationship between wavelength and pixel number, after which a zero point shift was measured for each fiber using the strong sky lines in the data. 5 Finally, the sky emission was subtracted from each fiber using an unscaled template constructed from the 16 sky fibers in each frame. Classification and radial velocity measurements for each spectrum were done in two ways. First each spectrum was inspected by eye, and classified as ei- ther a star, galaxy or unknown. Then a radial veloc- ity was determined by measuring either the position of the 4000A break, or that of the [OII] λ3727.6 fea- ture (actually a doublet, but unresolved in our data). A subjective assessment was also made of the reli- ability of the resulting radial velocity, dividing the sample into 'possible' and 'definite'. All the spectra were also analysed using the fxcor routine in IRAF. Each object was cross-correlated with three different templates: (a) a template made up from the 27 best S/N stars in the data, (b) a template made up from 3 high S/N late-type stars, and (c) a template made up from 16 emission line galaxies (all shifted to their rest frames) in the data. For each spectrum, the correla- tion with the highest peak was then selected. A re- assuringly close match was found between the by-eye classification and velocities and the results from fxcor. The resulting histogram of velocities was inspected. Apart from 4 low S/N or hot stars (for which a good template was not constructed), the stellar velocities found by fxcor had a distribution well represented by a gaussian with zero mean and an approximate dis- persion (σ) of 85 km/s. We take this to be a reason- able estimate for our radial velocity uncertainties. It was also found that the subjective 'possible' category in the by-eye radial velocity measurements matched rather well to a cross-correlation peak height less than 0.3 returned by fxcor. Apart from one outlier, the difference between the fxcor velocity and that mea- sured by eye for the galaxy identifications with cross- correlation peak heights above 0.3 also was fairly well fit by a gaussian with σ of 85 km/s. In the end we obtained 129 definite galaxies, 43 possible galaxies, 86 definite stars, 4 possible stars, 10 fibers that had to be unplugged due to over- illumination (which were hence almost certainly stars), 37 fibers which showed no flux (either very low surface brightness galaxies or positional errors) and 21 spec- tra which showed flux but which were unclassifiable either as stars or galaxies (based on very low cross- correlation peak heights from fxcor and visual inspec- tion). We should reiterate that a much higher 'suc- cess' rate in finding galaxies could have been achieved by raising the cutoff 'sigma' value used in choosing galaxies from the APM scans, at the cost of missing some compact galaxies. We present in table 1 the resulting galaxy red- shifts. For each object, the table lists the RA and Dec, an approximate B magnitude and the heliocen- tric radial velocity. The magnitudes were calculated using the APM isophotal magnitudes measured from the plates, crudely calibrated using the B magnitudes listed by Stockton 1980 and Salzer 1992 for objects in the scanned region. They could be in error by as much as 0.5. A number of the brightest galaxies in the field were not included in the fiber survey. Those with known redshifts within the survey region were added to the sample (4 objects taken from the May 5 1990 version of the CfA Redshift Catalog, Huchra 1990). This gives a total sample of 176 galaxies with redshifts within the survey region which is roughly complete to a B magnitude of 19.0. We plot the results of the survey in a number of projections. Figure 3 shows a redshift histogram of the galaxies, figure 4 shows the distribution of galax- ies on the sky, while figure 5 shows the pie-diagrams obtained. 3. Analysis In this section we go through each of the three observations described in § 2. in turn, deriving con- straints on the absorbers. In § 3.1. we derive limits on Hα line emission at the velocities of the narrow-band filters and in § 3.2. we calculate the maximum abso- lute magnitude a dwarf galaxy could have and remain undetected in our broad-band imaging. A long de- scription of the correlation analysis between the Lyα absorbers and the galaxies found in the fiber survey is given in § 3.3., in which the various available absorber and galaxy sub-samples are discussed, and two al- ternative extreme hypotheses for the absorber-galaxy correlation function are tested. Finally in § 3.4. we discuss some particular aspects of the absorber-galaxy distribution found in the 3C273 sightline. 3.1. Flux Limits for Hα Line Emission The final continuum-subtracted coronagraph im- ages have a measured rms dispersion (away from residuals due to bright stars) of 0.0018 DN/pixel/s. From calibration observations of HD84937, this is equivalent to a 1σ flux limit of 2.10−18 ergs/cm2/s/⊓⊔′′, equivalent to an emission measure of approximately 6 2.8 pc cm−6. By blinking the images, it can be seen that none of the objects visible in the broad-band images have emission line fluxes > 3σ. For compari- son, the VN2 image of Mkn49 showed a peak Hα line flux of 3.10−14 ergs/cm2/s/⊓⊔′′, i.e. 5,000 times higher than our 3σ detection limit. One can also perform the following thought experiment. What would the Orion nebula look like if placed at the distance of the Virgo cluster? As discussed in Kennicutt 1984, the Orion nebula is actually a relatively low luminosity H II re- gion, with an Hα luminosity of only 1037 ergs/s. Nev- ertheless, if placed at a distance of 16 Mpc, it would still have a flux of 1.3 10−15 ergs/cm2/s (and would be unresolved - the nebula has a diameter of 5 pc, while at Virgo, 1′′ corresponds to approximately 80 pc). Thus it would be a factor 220 brighter than our 3σ limit. One can also calculate that the Stromgren sphere around a single main sequence star of spec- tral class ∼B1/⊓⊔′′ would be detectable at the 3σ level (Allen 1973). Unfortunately, the expected surface brightness of an optically thick slab of hydrogen, simply bathed in the local UV background would not be detectable. Taking the limit on the UV background from Songaila, Bryant and Cowie 1989, and using the formulae from Osterbrock 1989, one finds a surface brightness of ≤ 3.10−19 ergs/cm2/s/⊓⊔′′, about a factor 20 below our 3σ limit. Higher spectral and spatial resolution data have been taken by T. Williams (private communication) using the Rutgers Fabry-Perot system at the CTIO 4m telescope. This data has not been fully analysed, but should produce even lower surface brightness lim- its (or detection) than our data. 3.2. Limiting Magnitude for detection of Low Surface Brightness Dwarf Galaxies The main motivation for taking the COSMIC im- ages was to determine whether there were any dwarf galaxies within a 40 kpc radius from the line of sight to 3C273 which are too faint to be visible on the POSS sky survey plates. Examples of low surface brightness dwarfs in the Virgo cluster can be seen in Sandage and Bingelli 1984. In order to determine how faint a dwarf galaxy, which had morphological properties typical of those found in the Virgo cluster, would be detectable in our image, we used the IRAF artdata package to insert artificial dwarf galaxies into the data array. The magnitude scale was derived by assuming that the sky background in the Gunn r band (before the moon rose) was 21.5 (Massey 1990). B absolute magnitudes were converted to Gunn r assuming a B-R color of 1.0 (B-V of 0.7), and a conversion from R to Gunn r of r=R+0.43+0.15(B-V) (Kent 1985). Thus a dwarf galaxy with MB=-15.5 was taken to have Mr=- 16.0. The Virgo distance modulus was taken to be 31.02. Dwarf galaxy properties were taken from Bin- gelli, Sandage and Tammann 1985. In particular, typ- ical exponential scale lengths4 for Virgo dwarfs were taken to be 2-4 kpc. Figure 6 shows the same data as figure 2, but with three dwarf galaxies added. One to the NE with MB=-14.5 and scale length 4.4 kpc, one to the NW with MB=-13.5 and scale length 2.2 kpc, and one to the SW with MB=-13.5 and scale length 4.4 kpc. These experiments demonstrated that we would be able to detect dwarf galaxies as faint as -13.5 at the distance of Virgo, having morphological prop- erties similar to those found near the cluster center. For comparison, the dwarf galaxy illustrated in figure 2, panel 4, labelled '15◦47' of Sandage and Bingelli 1984 has an absolute B magnitude of -14.8. A complementary approach is to search for H I 21 cm emission from Virgo dwarf galaxies. Re- cently, van Gorkom and her collaborators have set extremely low limits with the VLA over a 40×40 arcmin field centered on 3C273 and over a velocity range of about 1000 km/s centered on 1300 km/s to a 1 sigma column-density limit of approximately 1019 cm−2 (van Gorkom 1993, van Gorkom et al. 1993). 3.3. Correlations between Lyman α Absorbers and Galaxies We have shown there are no H II regions, or other strong Hα line-emitting gas, or dwarf galaxies near the Virgo absorbers. Having determined the redshift distribution of galaxies near the sightline to 3C273, (§ 2.3.) we would now like to address the statisti- cal question of the degree to which the Lyman α ab- sorbers are correlated with luminous galaxies. If they are correlated, is the Lyman α absorber-galaxy cor- relation the same as the galaxy-galaxy correlation? A cursory inspection of figure 5 is enough to show that there will not be a simple answer to this ques- tion. While there do seem to be Lyman α absorbers associated with clumps of galaxies (e.g. the Virgo absorbers, or the set of absorbers around z=0.02- 0.03), there are also absorbers in conspicuous 'voids' (at z=0.06-0.07), and there are no absorbers associ- 4Scale Length=R0, with Intensity ∝ exp(−1.6783 × R/R0) 7 ated with the prominent excess of galaxies at z=0.078. We consider statistical tests for various assumptions about the correlation between the Lyman α absorbers and the galaxies near the sightline towards 3C273 for which we have redshift information. Ultimately, the goal should be a quantitative and complete statistical description of the clustering properties of the Lyman α absorbers themselves and their correlation with var- ious types of galaxies, clusters, voids etc. Given both our rather meagre understanding of this problem and the small data set, we shall concentrate on testing the following two extreme null hypotheses about the Lyman α absorber clustering properties: 1. The Lyman α absorbers are uncorrelated with galaxies and are randomly distributed. (The second part of this assumption necessarily im- plies the first, but the converse is not necessarily true: The absorbers could be correlated among themselves but be uncorrelated with galaxies) 2. The Lyman α absorbers are correlated with galaxies in the same way that galaxies are corre- lated. More precise formulations of this hypoth- esis depend upon the particular test applied, as described below. In carrying out most of the tests described below, it is necessary to compute the 3-dimensional distance between every absorber-galaxy or every galaxy-galaxy pair. The question arises as to how to compute the component along the line of sight, since departures from a perfectly smooth Hubble flow distort the map- ping of redshift onto radial distance. With no infor- mation other than the angular coordinates and red- shifts for the objects, we cannot uniquely determine the separation along the line of sight for any individ- ual pair of objects. For purposes of statistical tests we therefore make two different assumptions about this component; the degree to which we do or do not derive similar results will provide some indication of the sensitivity of the test to the uncertainty in the estimation of this component: 1. We simply ignore any departures from Hubble flow 2. We adopt the formalism of Davis and Peebles 1983 to estimate this component. These au- thors show how, knowing the two-point correla- tion function for the projected distance between pairs, in principal one can invert the integral equation relating the projected distance corre- lation function to the three-dimensional spatial correlation function. However, for our limited and rather noisy sample, this is not a very sat- isfactory procedure. Moreover, we would like statistical estimates for the 3-dimensional sepa- ration for each pair for the purposes of carrying out other types of tests (e.g. nearest neighbor tests). We therefore use the Davis-Peebles for- malism as follows: Adopting the assumptions described in their paper, the integrand in their equation (22) represents the probability that a pair with projected separation rp Mpc and ve- locity difference π km/s has a separation in the radial direction of y Mpc. We further assume their functional form for h(r) in their equation (23), with the parameter F = 1, and adopt their functional form for f(V), the probability of the relative velocity difference, as well as their ex- pression for σ, the dispersion in f(V), i.e. their equation (32). For rp >> ro or π >> Ho × ro, where ro is the characteristic correlation length, the probability is strongly peaked at y ∼ Ho ×π, i.e., the pair separation is very likely to be that given by assuming a pure Hubble flow. How- ever, as both of these inequalities fail to be sat- isfied, a second maximum in the probability dis- tribution arises at y ≈ 0 whose height depends upon the strength of the correlation - i.e., a pair with a moderate velocity separation and small projected separation may be separated by their Hubble flow distance, but if the correlation be- tween such pairs is strong, this moderate ve- locity separation is more likely to arise from a pair at about the same distance from us, with strong gravitational interaction between them. As noted at the outset of this discussion there is no way to unambiguously determine the rel- ative separation of any pair along the line of sight, but since we are interested in statistical applications we adopt the expectation value of this probability distribution as our second alter- native algorithm. A problem with this second approach is that evalu- ating the probability distribution for the radial sep- aration of the pair requires that we know in advance the two-point correlation function between the pair. Ideally, we could have dealt with this problem by using an iterative approach: for both the absorber- 8 galaxy and galaxy-galaxy pairs separately, we start with some "fiducial estimates" for the two-point cor- relation (e.g. ro = 5.4h−1 Mpc and γ = 1.77 ; c.f. equation 19 in Davis and Peebles) to compute the ex- pectation value of the radial separation for each pair, and thus compute the two-point correlation functions. With these new, separate best-fit values of γ and ro for the correlation functions for the galaxy-absorber pairs and the galaxy-galaxy pairs, we could then re- peat the process until the parameters for the two- point correlation functions have converged. In fact, since the results of our statistical tests do not appear to be very sensitive to departures from a pure Hubble flow, we have not carried out this iteration, but have simply used the single set of parameters (ro = 5.4h−1 Mpc and γ = 1.77) in calculating the expectation val- ues for both sets of pairs. In the following, we refer to these two algorithms for estimating the radial separations as the "pure Hubble flow" and "perturbed Hubble flow" cases. When listing object separations in the following sec- tions we will give the separations found from the ex- pectation value of the perturbed Hubble flow model in brackets following the value for the pure Hubble flow model. Before carrying out any statistical tests we define the two samples and discuss appropriate corrections for completeness. 3.3.1. The Lyman α Sample A carefully defined list of Lyman α absorption lines is essential to a proper statistical discussion of the cor- relation properties of Lyman α absorbers with galax- ies. The preferred list would obviously be drawn from a homogenous set of observations with the smallest detectable equivalent width covering all or most of the relevant redshift range. Several line lists for the 3C273 sight line have been published (Morris et al. 1991, Bahcall et al. 1991a, Bahcall et al. 1991b, Brandt et al. 1993, Bah- call et al. 1993). These line lists are compared in table 2. However, not only are these lists based upon 3 different HST spectrograph configurations (GHRS G160M, GHRS G140L, FOS G130) but they have also been produced by different reduction procedures and line-finding and measuring algorithms and with differing acceptance criteria for what constitutes a "real" line. To investigate the importance of this lat- ter source of inhomogeneity we have run the same continuum fitting, absorption line-finding and line- measuring software, "JASON", used for the FOS line list, which is described in detail by Schneider et al. 19935 on the data sets of Morris et al. 1991 and Brandt et al. 1993. The JASON software was de- signed to run on FOS data with an approximately gaussian point spread function (PSF). Unfortunately this is not a good representation for the GHRS large aperture PSF, and also in general the lines were re- solved - meaning that the observed line profiles had neither the instrumental PSF nor a gaussian shape. The current version of JASON does not perform such convolutions, and so we have run the search routines assuming a fixed PSF with the correct (non-gaussian) shape for the GHRS, but with no account taken for resolved lines. This means that the EWs output by the JASON software are not accurate, but the detec- tion significance levels are approximately correct (see Schneider et al. 1993). The data set used is that described by Brandt et al. 1993. We tabulate the sig- nificance levels from JASON for the Lyman α lines in column 5 of table 2, after the positions and EWs pub- lished in Brandt et al. 1993. It can be seen that all of the 'reliable' lines listed by Morris et al. 1991 are confirmed by the JASON software, with the notable exception of the line at λ1276.54, which was also not found by Brandt et al. 1993. We checked this line by running the JASON software on the original data used by Morris et al. 1991, obtaining a significance level of 4.5 for the line. The line finding and fitting software used in Morris et al. 1991 was developed by R. Carswell and J. Webb, and is described in that paper. It used the GHRS PSF convolved with a Voit profile with variable width. Two other entries in table 2 require special com- ment: The line at λ=1317.08 is identified as Lyman α in the list of Bahcall et al. 1991b whereas in the list of Morris et al. 1991 it is identified as Ni II. This is- sue has been discussed in detail in Brandt et al. 1993 who give reasons for preferring the Ni II identifica- tion which we adopt. The second case involves the line at λ=1393.86. As discussed in detail by Savage et al. 1993 this line appears to be a blend of one of the members of the galactic Si IV doublet and an- other strong line, whose only plausible indentification is Lyman α. The procedure used by Morris et al. 1991 in estimating the strength of this line (which in- 5We thank D. Schneider for kindly making available the most recent version of JASON and for instruction in its use. volves a detailed comparison of the line profiles of the Si IV doublet) is not incorporated into the JASON formalism. For this reason we cannot assign a formal uncertainty in the line strength. However the resid- uals from an unblended fit to the Si IV doublet are highly significant. As a result of the above considerations, we have decided to adopt the following samples of absorbers for our statistical tests: (1) For the Lyman α ab- sorber "full sample" we adopt the list of 16 Lyman α absorbers (and their redshifts) given in Morris et al. 1991 along with the additional low-redshift line (λ1224.52) given by Bahcall et al. 1991a; as noted in table 2, this last line is visible in the GHRS G140L spectra but was below the significance threshold of Morris et al. 1991. This "full sample" is inhomoge- nous and/or biased in three senses: i) The high resolu- tion GHRS data covers only redshifts above z∼0.016; below this, only the FOS data of Bahcall et al. 1991b and the GHRS G140L low resolution data are avail- able. (Observations with the GHRS G160M grating in the redshift regime from 0.0 to 0.016 are sched- uled for HST Cycle 3). Thus, the full sample may be biased against weak low-redshift lines in this red- shift range that may be detected by these Cycle 3 observations. ii) The line of sight toward 3C273 may be somewhat atypical in that it passes through the southern extension of the Virgo cluster. iii) Some of the weakest lines listed as "possible" in the Morris et al. 1991 list may not be real. Accordingly, along with the full sample of 17 lines, we shall also con- sider two subsamples: (2) A "homogeneous sample" made of the set of 14 lines found only with the GHRS G160M observations using the original Carswell and Webb software (i.e. all but the first 3 lines in column 1 of table 2). (As it happens, this is also equivalent to deleting the 3 low-redshift (z∼<0.016) lines possi- bly associated with the Virgo cluster). (3) A "strong sample" composed of the set of 10 lines from Morris et al. 1991 with -log(P)>7.5 (but including the line at z=0.14658 for which a formal probability estimate was not possible due to blending - i.e. the above sam- ple with the lines marked 'd' in the comments column of table 2 removed). We have listed the sample mem- bership in the final column of table 2. Note that in contrast to the galaxy sample discussed below there is no intrinsic observational selection against the higher redshift absorbers. One could, of course, define further samples. In particular, at the request of the referee, we have also 9 run our statistical tests of the complete set of 5 lines listed in Bahcall et al. 1993 (i.e. the line list in col- umn 8 of table 2). Unfortunately, the number of lines in this list is so small that neither of the two null hypotheses considered below in connection with cloud-galaxy association can be rejected with any sig- nificance. For this reason, and in order to keep the various combinations of absorption line and galaxy samples to managable proportions, we limit our ta- bles of statistical results to consideration of the three samples defined above. 3.3.2. The Galaxy Sample An appropriate sample of galaxies with which to carry out the correlation analysis would be one which is complete to some limiting absolute magnitude through- out a cylindrical volume centered on the 3C273 sight- line (i.e. out to a constant impact parameter) with a radius large enough to sample most of the expected power in the correlation function and length over (and beyond) the full redshift range covered by the Ly- man α line sample. The observed sample described in § 2.3. fails this requirement in two obvious respects: i) It detects only the more luminous galaxies at the higher redshifts, ii) It contains no galaxies with large impact parameters at low-redshifts. For some tests these deficiencies are probably not important but for others they are. Accordingly, we will consider two galaxy samples. The first is the sample described in § 2.3., which we will refer to as the 'cone' sample. The second is the cone sample together with all galaxies from the May 5 1990 version of the CfA Redshift Cat- alog, (Huchra 1990) within 10 Mpc of the 3C273 line- of-sight. This gives a heterogeneous sample with an unknown selection function, but is closer to the ideal 'filled cylinder' than the cone sample. It contains 1498 galaxies, the vast majority being at Virgo distances, and will be referred to as the 'cylinder' sample. Thus for each hypothesis tested below there are 6 galaxy-absorber sample combinations, and 2 possible estimates for the distance between every pair of ob- jects giving a total of 12 data sets for each statistical test. 3.3.3. Tests of the 1st Null Hypothesis: The Lyman α absorbers are uncorrelated with galaxies It is obvious that this hypothesis cannot literally be true - every sightline that passes close to any galaxy, except those utterly devoid of gas, will surely pro- 10 duce a detectable Lyman α line, and indeed examples of this are already known (Bahcall et al. 1992a, Bah- call et al. 1992b). Nevertheless, in light of the fact that the high-redshift Lyman α absorbers show al- most no power in their two-point correlation function (c.f. Rauch et al. 1992 and references therein) it is of interest to see if the present data set does or does not exclude this hypothesis, and if it does, how strongly. Having formulated this null hypothesis we consider two statistics as measures of correlation (or lack of it): 1. The average over all the Lyman α absorbers of the distance between a given Lyman α absorber and the N nearest galaxies in the sample, with N=1,3 and 5. 2. The total number of absorber-galaxy pairs within a fixed radius R, with R=500 kpc and 10 Mpc. In order to see whether the values of these observed statistics are such that the null hypothesis can be re- jected, we must find the distribution of these same statistics for many realizations of the null hypothesis. To do this, we carry out 1000 Monte Carlo simula- tions in which the same number of absorbers as that of the particular Lyman α sample under considera- tion are laid down randomly (but follow the "global" distribution dN/dz ∼ (1+z)0.3 as determined by Bah- call et al. 1993). Redshift limits for the random ab- sorbers were 0.03<z<0.151 for the total sample, and 0.016<z<0.151 for the homogeneous and strong sam- ples. The results for the nearest-neighbour tests are summarised in table 3. We show results for the single nearest neighbour galaxy, the mean of the 3 near- est and the mean of the 5 nearest. For each case, the three columns list the observed mean nearest- neighbour(s) distance, the average of the mean nearest- neighbour(s) distances produced by the Monte Carlo simulations, and the number of the 1000 Monte Carlo simulations that had a mean nearest-neighbour dis- tance less than the observed one. Thus this last col- umn, divided by 1000, can be taken as the probability that the observed value could arise from a sample of absorbers distributed at random with respect to the galaxies. The most striking result is that there is a less than 0.1% probability that the average nearest-neighbour distance to the single closest galaxy could arise from a randomly distributed set of absorbers. This is true for all sample combinations, and for either pure or perturbed Hubble flow. For all samples, as one in- cludes more galaxies in the nearest-neighbour aver- age, the significance drops. After seeing this result, we wanted to test whether all the significance came from the nearest galaxy, and so ran tests on the sec- ond nearest galaxy only (also given in table 3). One can see that the observed mean distance is still signif- icantly lower than that expected for a randomly dis- tributed set of absorbers, but this could be explained by a combination of the highly significant correlation with the nearest galaxy and the strong galaxy-galaxy two-point correlation. This point is discussed in some detail by Phillips, Disney and Davies 1993 in the con- text of bright galaxies found near quasar MgII ab- sorbers. We also give in table 4 the RA, Dec and redshift of the nearest galaxy in the 'cylinder' sample to each Lyα absorbers of the 'total' sample (assuming pure Hubble flow). The final column in this table is the minimum absolute magnitude that could have been detected in the fiber survey. This shows that for one absorber there is no known galaxy with absolute mag- nitude above -17.8 within nearly 10 Mpc, (the same distance for the perturbed Hubble flow model) and that even the nearest absorber-galaxy pair in our sam- ple are separated by 350 kpc (240 kpc). We will return to this in § 3.4. The results for the number of galaxies within a fixed radius of each absorber are given in table 5. These tests are essentially a comparison of the in- tegrated two-point correlation function out to the given radius (see Mo et al. 1992 for a discussion of this point). The numbers given in the table are the observed number of absorber-galaxy pairs within the given radius, the average of the number of pairs found in 1000 Monte Carlo simulations with randomly distributed absorbers, and the number of the Monte Carlo simulations with a larger number of pairs than that observed. Thus the final column divided by 1000 is the probability that the observed numbers of pairs or more would arise from a random distribution of absorbers. Table 5 shows that there is no significant excess of galaxies within volumes of radius 10 Mpc centered on the absorbers compared to a random distribution, apart from the cylinder/total subsample. For this combination there are a large number of pairs between the Virgo absorbers and the many Virgo galaxies in the CfA catalog. This result may be interpreted as saying that it is surprising to find 3 out of 17 absorbers below z=0.008 (although see § 3.4.3.). There is a marginally significant excess of absorber-galaxy pairs within 500 kpc, over that expected for a random dis- tribution of absorbers, although the inclusion of the Virgo velocity range removes the significance of this result. In summary, these tests seem to be consistent with the nearest-neighbour distance results, showing that there is an excess of close pairs of absorbers and galaxies, but that this result vanishes if the averaging is done over several galaxies or large radii. 3.3.4. Tests of the 2nd Null Hypothesis: Identi- cal Lyman α absorber-Galaxy, Galaxy-Galaxy Correlations In order to test this hypothesis, one would like to use the same tests as were used in § 3.3.3. A diffi- culty arises though in generating a large number of Monte Carlo samples. We are loathe to compare the observed distributions with simulations involving any- thing other than the actual observed galaxy distribu- tion since differences between the observations and simulations (based, for example on n-body or other galaxy clustering models) may result simply from in- adequacies of such models, and it is not clear how to create simulations of (fake) absorber-(real) galaxy dis- tributions having cross correlation properties which are the same as the observed galaxy-galaxy correla- tion properties. One way to deal with this difficulty is to use a test which does not require the generation of Monte Carlo samples: If the absorbers are distributed in the same way as the galaxies, then, given the pencil-beam na- ture of the galaxy sample, the redshift distributions of the absorbers and the galaxies should be identical, after correction for differing selection effects in the two samples. Table 6 shows the Kolmogorov-Smirnov D-values and probabilities that the absorber redshift distribution is the same as that of the galaxies, after the galaxy distribution is corrected using the selec- tion function shown in figure 3. This selection func- tion was derived by assuming a Schechter luminos- ity function with M∗=-19.5 and α=-0.97 (Loveday et al. 1992). Because this test requires a known selec- tion function, it can only be run for the 'cone' galaxy sample. Also, as it directly compares the redshifts, it does not require any assumptions about pure or per- turbed Hubble flow. It can be seen from table 6 that when all the absorption lines are included there is a highly significant difference in the redshift distribu- tions. This significance level becomes marginal when 11 the Virgo absorbers are removed, and vanishes when only strong absorbers beyond Virgo are considered. Our other test of the hypothesis that the absorbers and galaxies have identical correlation functions uses the observed galaxy sample to generate our Monte Carlo 'absorber' sample. For each realization a num- ber of the actual galaxy redshifts were selected at random, and were treated as absorbers on the 3C273 sightline. The actual algorithm involved selecting at random a number of the observed galaxy redshifts equal to the number in the absorber sample (making no correction for the galaxy selection function, and with no restriction on how close together the chosen galaxies were), and treating these redshifts as if they were measured absorber redshifts on the 3C273 line of sight. The galaxies which provided these redshifts were removed from the galaxy sample for each test, to avoid an excess of 'spurious' pairs6. The number of absorber-galaxy pairs within a given radius were determined for both the real and 1000 Monte Carlo samples, in an identical manner to the second test in § 3.3.3. above. This procedure should be valid as long as the radius within which the pair counts are being made is significantly larger than the typical distance between a galaxy and the 3C273 line-of-sight. Be- cause of this, tests were not run for a 500 kpc radius. The results from these tests are given in table 7. The columns are the same as for table 5, except that the final column lists the number of Monte Carlo runs with fewer absorber-galaxy pairs than the real sam- ple. Thus this number divided by 1000 is the prob- ability that the absorbers could have a correlation function as strong as that between galaxies. As can be seen, the absorber-galaxy correlation function av- eraged over 10 Mpc is significantly weaker than that between galaxies. In summary, the two tests above seem to show that the absorber-galaxy correlation function is sig- nificantly weaker than the galaxy-galaxy correlation function over large scales (10 Mpc). Even though there is clear evidence for galaxy-absorber clustering, there is a significant difference between the strength with which the galaxies are clustered with respect to each other and the strength with which the Lyman α 6In fact leaving in the galaxies would allow one to model a situ- ation where the absorbers were actually part of the halo of one of the observed galaxies, although some maximum halo size would also have to be imposed to make the simulation realis- tic. We are investigating this point in more detail, and a paper is in preparation (Mo, Morris and Weymann 1993) absorbers are associated with galaxies. Some of the results of the two preceding sections can be inferred directly by inspection of the actual two-point correlation functions themselves. A log- arithmically binned version of the correlation func- tions for pure and perturbed Hubble flows are shown in figure 7. The absorber-galaxy correlation function was generated using the total Lyman α sample and the cone galaxy sample. The correlation functions were normalised in the usual way using random sam- ples with the selection function shown in figure 3. The error bars were estimated using the formulae in Mo, Jing and Borner 1992. As expected, the pure and perturbed Hubble flow models agree fairly well for separations larger than one or two Mpc; the per- turbed Hubble flow model produces more very close pairs, since even small velocity differences wipe out small separations for the pure Hubble flow model. In both cases, the absorber-galaxy correlation is clearly weaker than the galaxy-galaxy correlation on scales from about 1-10 Mpc. However, although visual in- spection of the absorber-galaxy correlation function may suggest that the correlation is significant out to about 10 Mpc, in fact the pair tests summarized in table 5 and the nearest neighbor tests summarized in table 3 both indicate that a statistically significant absorber-galaxy correlations can be detected in our data sets only over volumes which are smaller than this. 3.4. Some Particular Cases of Interest 3.4.1. The Closest Absorber-Galaxy Associations Actual associations between individual observed galaxies and absorbers in the 3C273 sight line are dif- ficult to prove for a number of reasons: (1) The small- est projected distance to the 3C273 line-of-sight of all the galaxies in our sample is still 160 kpc. (2) Galaxy rotation or velocity dispersions could produce veloc- ity differences as large as 200 km/s between the mean galaxy velocity and an actually associated absorber (also comparable to the 3σ error in our fiber data velocity measurements) (3) In regions of high galaxy density along the line-of-sight, peculiar motions of the galaxies and the absorbers in cluster potential wells may make the velocity-distance relationship complex. This is especially true of the Virgo region, and for the z=0.0034 and 0.0053 absorbers (which also lie on the steep portion of the damping wings of the galactic Lyman α, and have only been observed at low resolu- 12 tion). With these caveats, it can be seen from table 4 that the closest absorber-galaxy pair has a separa- tion of 350 kpc (240 kpc). Outside of the Virgo re- gion (where the velocity-distance relationship may be less complex, but also where our galaxy sample goes to much less faint absolute magnitudes), the smallest separation is 410 kpc (350 kpc). The best published example of an association be- tween a Lyman α absorber and a galaxy is given in Bahcall et al. 1992a, where they find a galaxy within 90 kpc of the sightline to H1821+643 which has a strong Lyman α absorber (EW 950 mA) within 400 km/s. There are no galaxies in our sample this close to the line of sight (or indeed any absorption systems this strong). One can (somewhat arbitrarily) break the absorbers in our sample into two groups: (1) those with a galaxy within 1 Mpc of the line-of sight, and with a velocity difference of less than 400 km/s (all of the Virgo sys- tems and 6 of the higher-redshift systems. Despite its entry in table 4, the absorber at z=0.02622 actually has 2 galaxies within 400 km/s and 1 Mpc projected separation. They do not appear in the table, as their large velocity differences (>250 km/s) make their sep- aration large, assuming pure Hubble flow), and (2) the rest (8 systems). There is no significant difference in the EW distribution of these two samples. Our data can also be used to consider the ques- tion: What is the average galaxy diameter within which one would see a neutral-hydrogen column den- sity of at least 1013 cm−2? In practice this is rather a naive question, as the cross-section almost certainly depends on galaxy luminosity and probably also mor- phology. One might also expect a patchy distribu- tion of neutral-hydrogen in the outer parts of galaxies leading to a covering factor not equal to one. Never- theless, ignoring these complications, and excluding the Virgo velocity range due to (a) the possibility of large peculiar velocities (b) a wish to avoid the large range of intrinsic galaxy luminosities in the sample and (c) because our absorber EW limit is higher in this region, we find that of the 12 galaxies with pro- jected separation to the line-of-sight less than 1 Mpc, 8 show Lyman absorption systems within 400 km/s. A velocity difference of up to about 400 km/s could possibly be attributed to internal motions within a large galaxy, coupled with our measuring error, or al- ternatively to a small group. Having 8 or more such matches would occur 0.6% of the time if the absorbers were randomly distributed in velocity space. How- 13 ever, of these 8 galaxies, two are associated with the absorber at z=0.02622, and three with the absorber at z=0.2933. Clearly several distinct galaxies can not be producing the same absorber, and so in fact only 5 of the 12 systems within 1 Mpc of the line-of-sight to the quasar can be legitimately associated with indi- vidual absorbers. Reducing the projected separation to 500 kpc, one finds 3 galaxies, with velocity sep- arations of 20, 120 and 0 km/s (see table 3). This may be interpreted as saying that the cross-section of galaxies with luminosities greater than 1/10 of L∗ for log(NH)> 1013 is between 0.5 and 1 Mpc, although by "cross-section" we do not mean that the absorbers are neccesarily associated with the actual galaxy in question. Because of the 3 Virgo absorbers, all of the Virgo galaxies have an absorber within 400 km/s. However, it is clearly unreasonable to suggest that more than one galaxy is associated with a given absorber. Thus of the 19 Virgo galaxies within 500 kpc of the line- of-sight, 16 must not be producing absorption. Of these 19, 2 have absolute magnitudes brighter than 1/10 L∗, and so are consistent with the above state- ment, despite the higher absorber EW limit, as long as only one of the galaxies with luminosity <1/10 L∗ is producing observed absorption. 3.4.2. The Most Isolated Lyman α Absorbers One of the stronger lines in our absorber sample is also the most isolated. Any galaxy brighter than -18 would be in our sample at the redshift of the iso- lated z=0.06655 absorber. However, the nearest such galaxy has a projected separation of 4.75 Mpc and a velocity difference of 710 km/s, which corresponds to a spatial separation of 10 Mpc for both pure and perturbed Hubble flow models. Indeed studying the galaxy distribution in figure 5, the three absorbers with 0.060<z<0.07 seem to lie in a 'void' in the galaxy distribution. While it is certainly possible that these absorbers are associated with galaxies below our ab- solute magnitude limit, these isolated absorbers are a clear demonstration that one can not associate ALL low-redshift Lyman α absorbers with luminous (L∗) galaxies. 3.4.3. The Absence of Lyman α Absorbers in the z ∼ 0.078 Galaxy Concentration The most striking feature of the observed galaxy redshift distribution is a concentration of galaxies cen- tered at a redshift of z ∼ 0.078. This excess is almost certainly associated with a structure which includes the galaxy cluster Abell 1564. This cluster has a tab- ulated center at 12:32:25 +02:07:11 (1950), giving it an offset of 89′ SW of the 3C273 sightline. This angle corresponds to a separation of 7.3 Mpc at the clus- ter distance. Abell 1564 is the closest Abell cluster to the 3C273 sightline, but is only richness class 0. Redshifts for two member galaxies are given in Met- calfe et al. 1989, giving an average cluster redshift of 0.0793. Selecting all galaxies in our fiber sample with redshifts between 0.07 and 0.085 (54 galaxies), one derives a mean redshift of 0.0781 with rms of 0.0021 (630 km/s). For comparison, selecting galaxies be- tween redshifts of 0.15 and 0.17 (the cluster around 3C273, see figure 3), one gets 24 galaxies with mean redshift 0.1581 and rms 0.0024 (720 km/s). Despite the location of Abell 1564 to the SW, the galaxies in the redshift slice 0.07-0.085 show a weak concentration to the NE of the fiber field, although there are galaxies in all parts of the fiber area sur- veyed. If they are really part of a structure includ- ing A1564, the structure must have a size of at least 8 Mpc. However, the nearest Lyman α absorber is 0.0095 from the peak in redshift space (or 2850 km/s, which is 4.5 times the rms from the mean of the con- centration) and is at a distance of 36 Mpc from the center of the concentration. It thus seems unlikely that there are any observed Lyman α absorbers in our line of sight which are physically associated with this concentration. We can test a slightly different form of the hypoth- esis considered in § 3.3.4., (though we grant its appli- cation to the z ∼ 0.078 galaxy concentration involves post facto statistics), namely: What is the probabil- ity that, if the Lyman α absorbers are distributed in space "in the same way" as the galaxies in our sam- ple, we should find the observed number of Lyman α absorbers (none, in our case). By "in the same way" we mean that: (cid:18) ρLyα(~r) ρLyα (cid:19) = (cid:18) ρgalaxy(~r) ρgalaxy (cid:19) and by "in the vicinity of" we shall mean within ±2.5 × rms of the redshift peak of the concentration. We consider that our selection function can be mean- ingfully applied over the redshift range from about z= 0.016 out to our adopted "proximity cutoff" at z= 0.151. Carrying out the integration of our sam- ple (i.e. the galaxy numbers weighted by the selection 14 function) between ±2.5×rms of the peak of the galaxy concentration and over the redshift range from 0.016 to 0.151 we find that after correction for the selection function, about 0.33 of an absolute-magnitude-limited cylindrical sample of galaxies should be found within 2.5σ of the velocity peak of this structure. Since the Lyman α absorbers are not subject to such a selec- tion function, we can apply this equation to all the volume elements along the line of sight and, assuming further that the effective cross-section of the Lyman α absorbers (σ) is constant over the redshift range 0.016-0.151, one can write: n × σ × [(zpeak + 2.5 × rms) − (zpeak − 2.5 × rms)] = 0.33 × n × σ × [0.151 − 0.016] This equation represents the number of Lyman α ab- sorbers in the vicinity of the peak that we should ex- pect to see if our hypothesis is correct. Over the red- shift interval 0.016-0.151 we have observed up to 14 Lyman α absorbers so that we expect 0.33 × 14 ≈ 4.7 absorbers, and we observe none. The probability of this occuring is thus exp(−4.7) ≈ 0.01. This is not a strong result due to the rather small number of ab- sorbers expected, and the post facto nature of the test as we acknowledged above, but is additional sug- gestive evidence that the Lyman α absorbers do not follow the galaxy distribution and in particular that they may avoid strong concentrations of galaxies. It could be strengthened, of course, by observing addi- tional lines of sight through dense concentrations. It would be nice to repeat the above calculation for the case of the three absorbers found in the Virgo velocity region. However, there are a number of rea- sons why any estimate of the probability of finding three absorbers in this region with an apparent over- density of galaxies is highly uncertain. Firstly, our absorber sample is incomplete in this velocity range, as described in § 3.3.1. Secondly, the area of sky sur- veyed is so small that the selection function correc- tion for this velocity range is very large. Taking our selection function at face value, the 8 galaxies actu- ally in our survey at Virgo velocities imply that 60% of an absolute-magnitude-limited cylindrical galaxy sample from z=0 to 0.151 would be found in that velocity range. Our hypothesis above would then predict that (neglecting the incompleteness of the absorber sample) one would expect 0.6×17≈10 ab- sorbers, when we only see 3. Unfortunately, it is not possible to place a meaningful uncertainty on this ap- parent under-density of absorbers in the Virgo region, for the reasons listed above. In summary, there is marginal evidence, both from a clump of galaxies at z=0.078 and also from the Virgo velocity range, that Lyman α absorbers are less common in regions of high galaxy density. 4. Summary and Discussion We have assembled several types of observations in an attempt to find objects with which the Lyman α absorbers along the line of sight to 3C273 might be associated, and in order to carry out statistical tests of galaxy-Lyman α absorber association. In particu- lar, we obtained narrow-band images centered on and off the expected position of any Hα emission which might be associated with three of the low redshift Ly- man α absorbers, and obtained deep broad-band im- ages of a 17′×17′ field centered on 3C273. Both these searches were negative. Our failure to identify any broad-band or Hα emission from plausible "galaxy- like" objects a few tens of kpc from the 3C273 sight line at the approximate distance of the Virgo clus- ter will be checked by more sensitive and extensive searches for Hα by T. Williams (private communica- tion) and 21 cm emission by van Gorkom 1993 and van Gorkom et al. 1993. We have also obtained redshifts for a large num- ber of galaxies in the vicinity of the 3C273 sightline. Again, we find no unambiguous instance of associa- tion of any of the Lyman α absorbers with individual galaxies. We define a number of samples for both the Lyman α absorbers and the galaxies and estimate the 3-dimensional separation between each galaxy-galaxy pair and each Lyman α absorber-galaxy pair based upon two models for converting the observed redshift difference between any pair into a radial separation, viz. i) the assumption of a pure Hubble flow and ii) a statistical model of "perturbed" Hubble flow based upon work of Davis and Peebles 1983. The result- ing data base is used to carry out statistical tests to confirm or reject two null hypotheses about the asso- ciation of galaxies and Lyman α absorbers, namely: i) The Lyman α absorbers show no tendency to clus- ter around galaxies ii) The Lyman α absorbers cluster around galaxies exactly as the galaxies cluster about each other. While neither of these two hypotheses can be unambiguously rejected in the sense that ev- ery combination of samples and flow hypotheses reject both of them at significant levels, the evidence from these tests, and from the galaxy-galaxy and Lyman α absorber-galaxy two-point correlations themselves, points quite strongly to the conclusion that both hy- potheses are false. In particular, over length scales from about 1 to 10 Mpc there seems little doubt that the Lyman α ab- sorbers cluster around galaxies less strongly than the galaxies themselves cluster. This is born out by an examination of a redshift interval centered at about z=0.078 at which a strong concentration of galaxies occurs but in the neighborhood of which there are no Lyman α absorbers. Additionally, we find at least one Lyman α absorber for which no galaxy with abso- lute magnitude brighter than about -18 can be found closer than about 5 Mpc. Taken together, all this ev- idence suggests that the most significant conclusion we have reached is that the majority of low-redshift Lyman α absorbers are not intimately associated with normal luminous galaxies. In view of the fact that it has long been realized that at high redshifts there is an absence of power in the Lyman α absorber two-point correlation function in redshift space, except possibly at the very smallest velocity separations, this conclusion is not too sur- prising. On the other hand, it is also fairly clear that there is some tendency for the Lyman α absorbers to cluster around galaxies, and even weak evidence that this clustering becomes strong at very small separa- tions. Also Bahcall et al. 1992a have investigated the auto-correlation function of the low-redshift ab- sorbers seen in the line-of-sight to H1821+643, and show that there is only a 4% probability that the ob- served 'clumping' arose from a randomly distributed sample. None of the foregoing points unambiguously, in our estimation, to a particular model for the formation and evolution of the Lyman α absorbers. As have oth- ers, we simply offer the following speculations which appear to be compatible with the facts as they are presently understood: At high redshifts, the Lyman α absorbers consist primarily of entities which are only very loosely asso- ciated with larger mass objects (e.g. proto galaxies), and which are evolving fairly rapidly. Possibly this dominant population consists of absorbers in which gravitational binding (eg by dark matter) plays no significant role. In addition to this group, there is a smaller population of absorbers which are evolving less rapidly, possibly stabilized by dark matter and which are clustered more strongly about galaxies. At 15 new Palomar Sky-survey plates, Alan Sandage for al- lowing us to inspect one of his older IIIaJ plates, Mark Davis and Margaret Geller for advice on generating Monte Carlo samples with known correlation func- tions, Don Schneider for help with the JASON soft- ware and Xavier Barcons, Donald Lynden-Bell, Mario Mateo, Houjon Mo, Michael Rauch, John Webb, Jeff Willick and Dennis Zaritzky for useful conversations. We would also like to thank the referee, J. N. Bah- call, for his helpful and constructive comments which improved the content and presentation of the paper. SLM and RJW acknowledge support through NASA contract NAS5-30101 and NSF grant AST-9005117. very low redshifts, this latter population is beginning to constitute a large enough fraction of the absorbers that power in the two-point correlation function is detectable. Thus, the present mix of Lyman α ab- sorbers appears to have clustering properties inter- mediate between present-epoch normal galaxies and a random non-clustered population. This property also appears to be shared by the low-luminosity moderate redshift "blue galaxies" (Pritchet and Infante 1992), leading to the plausible conjecture that the Lyman α absorbers are more closely related to low mass, low lu- minosity galaxies than they are to L∗ galaxies, though the relation is clearly not one-to-one. We have no good way at present of estimating the characteristic scale or masses of the low-redshift Ly- man α absorbers. A guess at a diameter of 30 kpc is as plausible as any. In particular, consider a pan- cake whose diameter is 30 kpc and whose thickness is 10 kpc. In this case, at the present epoch, a hydro- gen column-density of 1013-1014 cm−2 normal to the face of the absorber, coupled with estimates for the present-epoch energy-density of ionizating radiation leads to a total gas mass of order 107 M⊙, but a mass of only a few hundred solar masses of neutral hydro- gen. It is of interest that the mass function of neutral- hydrogen gas clouds appears to be truncated below about 108 M⊙ (Weinberg et al. 1991). As Maloney (1992,1993) has shown, the gas in a flaring galaxy with decreasing column-density will undergo a rather sudden transition along its face from being mostly neutral to mostly ionized, with the consequence that few if any contours with neutral-hydrogen column- density of order 1018 cm−2 are known. Similarly, it is conceivable that, given the appropriate run of length scale with mass, a sequence of masses would have the property of making a sudden transition in the mass function of neutral-hydrogen starting at about 108 M⊙, leading to a dearth of objects with total HI masses for several orders of magnitude below this. If these speculations have any connection with re- ality then one might expect to see some similarity in the clustering properties of the Lyman α absorbers and low mass galaxies. We are currently attempt- ing to obtain redshifts of galaxies of lower luminosity along the 3C273 sightline in order to investigate this possibility. We would like to thank Greg Aldring and Steve Shectman for help and advice in using the LCO fiber system, Neil Reid for giving us access to one of the 16 Table 1: Definite Galaxy redshifts in 3C273 Field 17 18 19 Table 2: Comparison of 3C273 Extragalactic Absorption Line Lists 20 Table 3: Nearest Neighbour Monte-Carlo tests Table 4: Nearest Galaxies to 3C273 Lyman α absorbers 21 Table 5: Pair Count Monte-Carlo tests using random absorbers Table 6: KS comparison of z-distribution 22 Table 7: Pair Count Monte-Carlo tests using the galaxy distribution 23 REFERENCES Maloney, P., 1992, ApJ, 389, L89 Allen, C. W. 1973, "Astrophysical Quantities 3rd Edi- Maloney, P., 1993, ApJ, In Press tion", Athlone Press, London, Pg 267 Bahcall, J. N., Jannuzi, B. T., Schneider, D. P., Har- tig, G. F., Bohlin, R. and Junkkarinen, V., 1991a, in "The First Year of HST Observations", ed A. L. Kinney and J. C. Blades, Pg 46 Bahcall, J. N., Jannuzi, B. T., Schneider, D. P., Har- tig, G. F., Bohlin, R. and Junkkarinen, V., 1991b, ApJ, 377, L5 Bahcall, J. N., Jannuzi, B. T., Schneider, D. P., Har- tig, G. F. and Green, R. F., 1992a, ApJ, 397, 68 Bahcall, J. N., Jannuzi, B. T., Schneider, D. P., Har- tig, G. F. and Jenkins, E. B., 1992b, ApJ, 398, 495 Bahcall, J. N., Bergeron, J., Boksenberg, A., Hartig, G. F., Jannuzi, B. T., Kirhakos, S., Sargent, W. L. W., Savage, B. D., Schneider, D. P., Turnshek, D. A., Weymann, R. J. and Wolfe, A. M., 1993, ApJS, In Press Binggeli, B., Sandage, A. and Tammann, G. A., 1985, AJ, 90, 1681 Brandt, J. C. et al., 1993, AJ, 105, 831 Chernomordic, V. V. and Ozernoy, L. L., 1983, Na- ture, 303, 153 Davis, M. and Peebles, P. J. E. 1983, ApJ, 267, 465 Fransson, C. and Epstein, R., 1982, MNRAS, 198, 1127 Haynes, M.P. and Giovanelli, R. 1984, AJ, 89, 758 Huchra, J. P. 1990 May 5 1990 version of the CfA Red- shift Catalog, obtained through the Astronomical Data Center. Jacoby, G. H., Branch, D., Ciardullo, R., Davies, R. L., Harris, W. E., Pierce, M. J., Pritchet, C. T., Tonry, J. L. and Welch, D. L., 1992, PASP, 104, 599 Kennicutt, R. C., 1984, ApJ, 287, 116 Kent, S. M., 1985, PASP, 97, 165 Loveday, J., Peterson, B. A., Efstathiou, G. and Mad- dox, S. J., 1992, ApJ, 390, 338 Massey, P., 1990, NOAO Newsletter # 21, March 1990 Mo, H. J., Jing, Y. P. and Borner, G., 1992, ApJ, 392, 452 Mo, H. J., Einasto, M., Xia, X. Y. and Deng, Z. G., 1992, MNRAS, 255, 382 Mo, H. J., Morris, S. L. and Weymann, R. J., In Preparation Metcalfe, N., Fong, R., Shanks, T. and Kilkenny, D. 1989, MNRAS, 236, 207 Morris, S. L., Weymann, R. J., Savage, B. D. and Gilliland, R. L., 1991, ApJ, 377, L21 Osterbrock, D. E., "Astrophysics of Gaseous Nebu- lae and Active Galactic Nuclei", 1989, University Science Books, CA Phillips, S., Disney, M. J. and Davies, J. I., 1993, MNRAS, 260, 453 Pritchet, C. J. and Infante, L., 1992, ApJ, 399, L35 Rauch, M, Carswell, R. F., Chaffee, F. H., Foltz, C. B., Webb, J. K., Weymann, R. J., Bechtold, J. and Green, R. F., 1992, ApJ, 390, 387 Salzer, J. J. 1992, AJ, 103, 385 Sandage, A. and Bingelli, B., 1984, AJ, 89, 919 Savage, B. D., Lu, L., Weymann, R. J, Morris, S. L. and Gilliland, R. L., 1993, ApJ, 404, 124 Schneider, D. P., Hartig, G. F., Jannuzi, B. T., Kirhakos, S., Saxe, D. H., Weymann, R. J., Bah- call, J. N., Bergeron, J., Boksenberg, A., Sargent, W. L. W., Savage, B. D., Turnshek, D. A. and Wolfe, A. M., 1993, ApJS, In Press Shectman, S. A., 1992, in "Fiber Optics in Astronomy II", ed P. M. Gray, ASP Volume 37, Pg 26 Songaila, A., Bryant, W. and Cowie, L. L., 1989, ApJ, 345, L71 Stockton, A. 1980, In IAU symposium 92, "Objects of High Redshift", ed G. O. Abell and P. J. E. Peebles, Pg 87 24 van Gorkom, J., H., 1993, Teton Conference Proceed- ings, In Press, Ed M. Shull and H. Thronson van Gorkom, J., H., Bahcall, J.N., Jannuzi, B., and Schneider.D. 1993, AJ, In Press Vilas, F. and Smith, B. A., 1987, Appl.Optics, 26, 4 Weinberg, D. H., Szomoru, A., Guhathakurta, P. and van Gorkom, J. H., 1991, ApJ, 372, L13 This 2-column preprint was prepared with the AAS LATEX macros v3.0. 25 Fig. 1. -- Coronagraph Data of 3C273 field. (a) Sum of 25A Filters, the horizontal bar in the lower left is 1 armin long, (b) Continuum Subtracted VN1, (c) Con- tinuum Subtracted VN2, (d) Continuum Subtracted HN. See text in § 2.1. Fig. 2. -- Mosaic of COSMIC Images in Gunn r of 3C273 field. See text in § 2.2. Region shown in figure 1 is circled. Fig. 3. -- Histogram of redshifts for Galaxies found in the 3C273 sightline. Also plotted are the locations of Lyman α absorbers from table 3 (where 'strong' refers to lines with EW≥55 mA, and 'weak' refers to the remainder), and the (arbitrarily normalised) galaxy selection function for the fiber survey. See text in § 2.3. Fig. 4. -- Locations on the sky of the galaxies with redshifts in the field of 3C273. See text in § 2.3. Fig. 5. -- Pie-diagrams for the Galaxies observed with the LCO fiber system. Angles have been exagerated by a factor 15 to prevent overcrowding of the symbols. Please note that this results in a highly distorted plot with initially spherical structures (such as the 3C273 cluster of galaxies) appearing elongated transverse to the line of sight. Note also that the star marking the position of 3C273, while readily visible in the pro- jection in RA, is partially obscured by a clump of galaxies in the projection in Dec Fig. 6. -- Mosaic of COSMIC Images in Gunn r of 3C273 field, with 3 artificial dwarf galaxies added. NE with MB=-14.5 and scale length 15kpc, NW with MB=-13.5 and scale length 7.5kpc, and SW with MB=-13.5 and scale length 15kpc. See text in § 3.2. Fig. 7. -- Two point correlation functions for galaxy- galaxy and absorber-galaxy (a) assuming pure Hubble flow (b) assuming perturbed Hubble flow. See text in § 3.3.4. 26 This figure "fig1-1.png" is available in "png"(cid:10) format from: http://arxiv.org/ps/astro-ph/9307005v1 This figure "fig1-2.png" is available in "png"(cid:10) format from: http://arxiv.org/ps/astro-ph/9307005v1 This figure "fig1-3.png" is available in "png"(cid:10) format from: http://arxiv.org/ps/astro-ph/9307005v1 This figure "fig1-4.png" is available in "png"(cid:10) format from: http://arxiv.org/ps/astro-ph/9307005v1 This figure "fig1-5.png" is available in "png"(cid:10) format from: http://arxiv.org/ps/astro-ph/9307005v1 This figure "fig1-6.png" is available in "png"(cid:10) format from: http://arxiv.org/ps/astro-ph/9307005v1 This figure "fig1-7.png" is available in "png"(cid:10) format from: http://arxiv.org/ps/astro-ph/9307005v1
0710.4094
1
0710
2007-10-22T16:30:02
Deep Optical Imaging of Starbursting "Transition" Dwarf Galaxies
[ "astro-ph" ]
A subgroup of dwarf galaxies have characteristics of a possible evolutionary transition between star-forming systems and dwarf ellipticals. These systems host significant starbursts in combination with smooth, elliptical outer envelopes and small HI content; they are low on gas and unlikely to sustain high star formation rates over significant cosmic time spans. We explore possible origins of such starburst "transition" dwarfs using moderately deep optical images. While galaxy-galaxy interactions could produce these galaxies, no optical evidence exists for tidal debris or other outer disturbances, and they also lack nearby giant neighbors which could supply recent perturbations. Colors of the outer regions indicate that star formation ceased > 1 Gyr in the past, a longer time span than can be reasonably associated with the current starbursts. We consider mechanisms where the starbursts are tied either to interactions with other dwarfs or to the state of the interstellar medium, and discuss the possibility of episodic star formation events associated with gas heating and cooling in low specific angular momentum galaxies.
astro-ph
astro-ph
Deep Optical Imaging of Starbursting "Transition" Dwarf Galaxies Kate E. Dellenbusch1 Department of Astronomy, University of Wisconsin, 475 N. Charter Street, Madison, WI 53706 [email protected] John S. Gallagher, III Department of Astronomy, University of Wisconsin, 475 N. Charter Street, Madison, WI 53706 [email protected] Patricia M. Knezek WIYN Consortium Inc., PO Box 26732, Tucson, AZ 85726 [email protected] Allison G. Noble2 Department of Astronomy, University of Wisconsin, 475 N. Charter Street, Madison, WI 53706 [email protected] ABSTRACT A subgroup of dwarf galaxies have characteristics of a possible evolutionary transition between star-forming systems and dwarf ellipticals. These systems host significant starbursts in combination with smooth, elliptical outer envelopes 1Visiting Astronomer, The 0.9m telescope is operated by WIYN Inc. on behalf of a Consortium of ten partner Universities and Organizations (see http://www.noao.edu/0.9m/partners.html). WIYN is a joint partnership of the University of Wisconsin at Madison, Indiana University, Yale University, and the National Optical Astronomical Observatory. 2Now at McGill University, 3600 rue University, Montreal, QC, Canada; [email protected] -- 2 -- and small HI content; they are low on gas and unlikely to sustain high star for- mation rates over significant cosmic time spans. We explore possible origins of such starburst "transition" dwarfs using moderately deep optical images. While galaxy-galaxy interactions could produce these galaxies, no optical evidence ex- ists for tidal debris or other outer disturbances, and they also lack nearby giant neighbors which could supply recent perturbations. Colors of the outer regions indicate that star formation ceased >1 Gyr in the past, a longer time span than can be reasonably associated with the current starbursts. We consider mecha- nisms where the starbursts are tied either to interactions with other dwarfs or to the state of the interstellar medium, and discuss the possibility of episodic star formation events associated with gas heating and cooling in low specific angular momentum galaxies. Subject headings: galaxies: dwarf -- galaxies: starburst -- galaxies: evolution -- galaxies: structure 1. Introduction The formation and evolution of dwarf galaxies remains poorly understood. How these processes produce both rotationally supported (dIrr) as well as dynamically warm (dE/dS0s) dwarf systems remain puzzles, despite considerable effort. An outstanding question in dwarf galaxy evolution is whether there is an evolutionary connection between the various mor- phological classes of dwarf galaxies (e.g. van den Bergh 1977, van Zee et al. 1998, van Zee et al. 2001, Lisker et al. 2007). Specifically, can dwarfs of one morphological type evolve by some process into a different type? For example, do some dIrrs evolve into dE/dS0s or did the dE/dS0s we see in the Universe originally form as that type? These questions regarding the formation and evolution of dwarf galaxies are of particular importance in light of hier- archical cold dark matter (CDM) galaxy formation models which predict large numbers of dwarf galaxies and suggests that larger galaxies are built up from the accretion of many low mass halos (e.g. White & Frenk, 1991). A related issue concerns the impact of star formation on dwarf galaxy evolution and particularly the role of episodic star formation (e.g. Searle, Sargent & Bagnuollo 1973, Lee et al. 2002). How does star formation relate to morphology in low mass galaxies? For example, where do blue compact dwarfs (BCDs) fit in the dwarf galaxy zoo as starbursting objects and are they related to either dIrrs or dE/dS0s (e.g. Sung et al. 1998, Gil de Paz & Madore 2005, Noeske et al. 2005)? In particular, it has been suggested that BCD galaxies may evolve into dE/dS0 galaxies after they lose their gas, either through supernovae-driven -- 3 -- winds during episodes of intense star formation (Marlowe et al. 1999, Tajiri & Kamaya 2002), or through stripping processes induced by galaxy-galaxy interactions (Drinkwater & Hardy 1991). Or, are some of these objects dEs where star formation has been renewed through gas capture as suggested by Silk, Wyse, & Shields (1987)? Previous studies, which have concentrated mainly on optical morphological differences between dwarf galaxies and to some extent on their neutral gas content, have not firmly established a direct evolutionary connection between dIs and dEs in galaxy groups and other low density environments (e.g. van Zee, Salzer & Skillman 2001, Noeske et al. 2005). Even in the Local Group the situation is unclear. Many Local Group dwarfs have complex star formation histories and kinematic peculiarities, and at least five of the faint dwarfs possess optical morphologies which place them in a "transition" or "mixed-type" structural category between dI and dE/dS0 (e.g. Mateo 1998, Grebel, Gallagher, & Harbeck 2003). While nearby dE galaxies, such as NGC 185 and NGC 205 in the local group, support star formation, the rates are extremely low and consistent with these objects' small gas supplies. More distant galaxy samples also contain dwarfs with early-type structures that also host HI gas and often also some young stars. As in the Local Group, these transition dwarfs frequently are small, low luminosity objects containing . 106 M⊙ of HI and having correspondingly low star formation rates (e.g. Bouchard et al. 2005). On the other hand, the Virgo Cluster of galaxies appears to contain a complete sequence of more luminous dwarfs with declining and increasingly concentrated star formation, extending from typical BCDs through to blue core dEs (Gallagher & Hunter 1989, Lisker et al. 2006,2007). While moderate luminosity galaxies with early-type outer structures and young stellar populations in their centers also exist in less dense environments (e.g. NGC 404, del Rio et al. 2004; NGC 5102; Deharveng et al. 1997), they are sufficiently rare that it is difficult to determine if evolutionary sequences exist. In this paper we explore possible histories for a sample of actively star forming luminous dwarf galaxies with transitional properties between BCDs and dEs residing in loose group environments. These systems are fairly isolated and show little indication of a recent inter- action or merger. They have an intriguing combination of characteristics which do not allow them to easily be categorized as either dI/BCD or dE systems, but rather show a mixture of the two. The key properties which indicate these low luminosity galaxies may be in the midst of an evolutionary transition from dI/BCD to dE include: • They are actively forming stars with star formation rates between 0.1 and 1 M⊙/yr. • The star formation is currently centrally concentrated, with the outer regions composed -- 4 -- of an older stellar population. • Although actively forming stars, the starburst is fueled by very little HI gas, with MHI/LB . 0.1. This compares with MHI/LB > 0.2 for typical dIs (Roberts & Haynes 1994) and BCDs (Pustilnik et al. 2002) and MHI/LB < 0.1 for dEs. Star formation can continue in our sample galaxies only for about another 109 years at their current rate, based on their HI content; in a few cases, including molecular gas will at most double this time (e.g. Jackson et al. 1989, Gordon 1991). • Unlike many other star forming dwarf galaxies including BCDs they have high oxygen abundance ratios, with 12 + log(O/H) > 8.4 as we found in Dellenbusch et al. (2007). • Their outer optical colors are similar to those of typical BCDs with (B-R)0 of order 1 (Gil de Paz et al. 2005), but bluer than the (B-R)0 ≈ 1.3 - 1.4 expected for slightly fainter dEs (Conselice et al. 2003). In this regard our sample resembles the Gil de Paz et al. "E-type" BCDs. • These objects have smooth outer isophotes which are much more indicative of early- rather than late-type galaxy structures. Table 1 quantifies several of these properties. In this paper we examine the significance of these final two characteristics. In §2 we discuss the observations and data reduction. In §3 we describe our analysis and modeling processes used to examine the data for evidence of tidal debris and fine structures. In this section we also describe the results of this analysis. We discuss two possible evolutionary pictures to explain the galaxies' star formation histories as a means of transition in §4. §5 contains a brief summary and our conclusions. 2. Observations and Data Reduction For this study, we obtained imaging data with the WIYN 0.9-m telescope on Kitt Peak in Arizona. We used the S2KB CCD camera to obtain deep R-band imaging of five transition dwarf galaxies. The S2KB camera provides a 20' × 20' field of view with a plate scale of 0.′′6. The data were taken on 24-26 and 28-29 April 2006; conditions were not photometric. We obtained total exposure times of 50 to 62 minutes for each galaxy by combining 4 or 5 750 second exposures. The data were reduced using standard methods in IRAF1. Data reduction 1IRAF 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 -- 5 -- included subtraction of combined bias frames and flat fielding using combined dome flats. The multiple dithered frames were aligned and combined. We also use short B- and R-band imaging observations made at Michigan-Dartmouth- MIT (MDM) Observatory on the 1.3 m McGraw-Hill telescope during 10-12 April 1996. Typical integration times were 20-39 minutes in B, and 10-15 minutes in R under generally photometric conditions. Details of the data reduction performed on these images can be found in Knezek, Sembach, & Gallagher, 1999). The sample of transition galaxies studied here, consisting of NGC 3265, NGC 3353, NGC 3773, NGC 3928, and IC 745, was selected to have a mixture of early- and late-type dwarf galaxy characteristics. That is, an underlying dE or dS0 with a high rate of current star formation. They thus were chosen in the spirit of Sandage & Hoffman (1991) and were initially observed as transition dwarf candidates with MDM along with NGC 3377A, NGC 4286, and IC 3475, which were discussed in Knezek, Sembach, & Gallagher (1999). 3. Results and analysis 3.1. Elliptical Modeling and Residual images In order to study the morphology and fine structure of our galaxies, we produced residual images, in other words the difference of the original images minus galaxy models. This allows weak features which are otherwise dominated by the overall smooth and symmetric light distribution to be more easily discernable. These residual images are created using several routines in the IRAF package ISOPHOTE. First, elliptical isophotes are fit to the galaxies using the ELLIPSE routine. The results from such a model of a galaxy's surface brightness distribution is then input into the BMODEL task to create a two-dimensional model of the galaxy (e.g. Forebes & Thomson 1992, Lisker et al. 2006), representing the smooth elliptical structure of each system. Because of intense star formation in the very centers of these galaxies, ELLIPSE does a poor job of fitting isophotes at small radii. Ellipses were fit starting in the outer galaxy and proceeding in- wards. These fits were stopped inside the 5 arcsecond isophote. By allowing the ellipticity and position angle of the model-fit to vary, any large-scale structure such as a bar, are in- cluded in the model. When the model is then subtracted from the original galaxy image, the residual more clearly reveals any fine structure (e.g. dust, tidal tails, shells) which had Science Foundation. been hidden beneath the dominant elliptical light distribution. -- 6 -- 3.2. Dwarf Galaxy Fine Structure Features The procedure described above was used to examine the deep R-band images for ev- idence of tidal interaction and clues to their evolutionary histories. We do not find tidal tails or extended tidal debris in any of the galaxies studied. The residual images do however reveal interesting fine structures near the galaxy centers where we find the light distribution is not smooth. Several of them have clear dust lanes, in particular NGC 3773 and IC 745 (See Figure 1). The residual images in Figure 1 are presented as negative images. The dust features therefore appear as lighter regions. These dust features are also visible in the original images. In addition, most of the objects do not show significantly boxy or disky isophotes as derived from a4 coefficients of the ellipse fits. NGC 3353 does show evidence of being boxy with -0.07 . a4/a . 0 out to a radius of 30 arcsec, while NGC 3265 is weakly disky with 0 . a4/a . 0.02 inside of 10 arcseconds. IC 745, NGC 3928, and NGC 3773 are not predominantly boxy or disky with a4/a not trending toward positive or negative values on average (See Figure 2). Thus, the conservative result is that the outer isophotes are quite close to ellipses. It is important to note however, that these trends are likely influenced by the prominent dust features present in the inner parts of these objects. In addition to dust features, several of the galaxies have complex fine structures in their inner regions (See Figure 1), indicating the light distribution is not smooth in the inner parts of these galaxies. Galaxies that have not experienced mergers in the past several Gyr generally have symmetric stellar bodies. In systems where mergers or stellar mass transfer events are more recent, distinctive features can be seen, including tidal tails. As these systems age, structural features associated with past events become increasingly subtle, passing from stellar shells and ripples to possibly tidal bars (Miwa & Noguchi 1998) or boxy isophotes (see Barnes & Hernquist 1992). If the galaxies we observe as starbursting transition dwarfs formed via near equal mass mergers, then these events were in the distant past or involved a companion with very few stars. The fine structure features as described above, are apparently associated with gas and are similar to those others have found in larger isolated elliptical galaxies. For example, Reda et al. (2004) studied a sample of nearby isolated early-type galaxies and found a variety of fine structure features in several of them such as dust lanes and shells as well as both disky and boxy isophotes. For several of the E's studied, Reda et al. suggest they are the remnant of a galaxy merger, as their light distributions deviate from perfect ellipticity. Thus the structure of the ISM in our sample allows for the possibility of past galaxy mergers; however -- 7 -- if these events occurred, they took place many outer-galaxy dynamical time scales in the past. 4. Discussion The galaxies discussed in this paper exhibit a puzzling combination of properties. Al- though these types of galaxies are not ubiquitous in the local universe, they may represent an important snapshot in dwarf galaxy evolution. It is therefore useful to understand their possible evolutionary histories and likely future states. Two potential scenarios are explored here; an internal ISM instability leading to multiple starburst episodes and a past interaction or dwarf-dwarf merger event. 4.1. Internal Chemo-Dynamical Star Formation Regulation The pattern of star formation in these galaxies indicates that producing a concentration of low angular momentum gas in the central regions plays an important role in their evolution. In this regard the problem is similar to that generally seen in BCD galaxies with high star formation rates: how is the central region of a starbursting dwarf fueled with gas (e.g. Taylor et al. 1993, van Zee et al. 2001)? In galaxies with low specific angular momenta, random motions and thermal energy could play important roles in supporting a spatially extended ISM. In the extreme situation where these are the only factors, numerical simulations suggest that the system can become unstable against bursts of star formation. This idea was first explored by Loose et al. (1982) who modeled the Galactic center. Stars energize the ISM which then expands, shutting off star formation, leading to collapse and another round of star formation; a cycle that resembles a kind of cosmic diesel engine. This type of galactic "star formation dieseling" also is seen in some of the more sophisticated 1-dimensional simulations of non-rotating dE galaxies by Hensler et al. (2004; see also Pelupessy et al. 2004). The basic principle is again the ability of mechanical energy from stars to expand the radial extent of a low mass ISM to the point where star formation declines, leading to a collapse on a gas dissipational time scale and another round of star formation, or gas exhaustion It is not clear whether the kinds of star formation rate oscillations found in these models occur in galaxies. A connection between mechanical power and the overall radial structure of the ISM is one key feature. For example, Tajiri & Kamaya (2002) estimate conditions under which BCD HI envelopes could be blown away, and the galaxies in this sample sit near their -- 8 -- boundary SFR for retaining their HI. Thus we expect the envelopes may be influenced by the current starbursts. A full assessment, however, requires knowledge of the gravitational potentials and structural form of the system. Unfortunately the stellar kinematics of the galaxies in our sample are not known and thus more detailed modeling is not yet warranted. Studies of HI in galaxies like those in our sample (central starburst, moderate luminosity, low MHI /L), however, show that some of the gas is in a rotating disk surrounding the central gas concentration (Lake et al. 1987, Sadler et al. 2000, van Zee et al. 2001). In some cases the gas kinematics are not regular; e.g. Lake et al. note the HI in the outer parts of NGC 3265 has confused kinematics, which they summarize as "infall, outflow, no-flow". Lake et al. and Sadler et al. consider that some of the HI could have been recently accreted, but this now seems unlikely given the high gas phase chemical abundances that we find in Dellenbusch et al. (2007). The possibility of an internal instability that leads to an episodic central concentration of gas is consistent with the chemical abundances and therefore merits closer examination. An alternative internal model for feeding gas into the centers of low mass galaxies via viscosity of gas in their disks has been developed by Noguchi (2001). In this picture low density gas disks in small galaxies fail to efficiently form stars, thereby allowing gas to accrete into the centers of the systems. This model, however, produces long-lived central star forming zones and thus it is not clear if it can be applied to the low HI content, high SFR objects studied here. 4.2. Old Interaction/Merger Event An alternative mechanism to explain the evolution of these galaxies is a past interaction with another galaxy or possibly a dwarf-dwarf merger. The amorphous dwarf galaxy NGC 5253 is a well studied nearby starburst galaxy which is similar in several ways to the galaxies discussed in this paper. It has been suggested that the central starburst in NGC 5253 may have been triggered by an interaction with M83 (e.g. van den Bergh 1980). NGC 5253 is an amorphous dwarf galaxy with an underlying elliptical structure and bright HII regions in the core. Like the transition dwarfs in this study, it has a low neutral gas mass-to-light ratio of MHI/LB = 0.04 and an Hα derived star formation rate of 0.12 M⊙/yr (Ott et al. 2005). It does however have a somewhat lower oxygen abundance with 12+log(O/H) = 8.23 (Martin 1997). Kobulnicky & Skillman (1995) have suggested the current star formation activity of NGC -- 9 -- 5253 may be due to the accretion of a gas-rich companion or the result of an interaction with nearby M83 about 109 years ago. M83 and NGC 5253 are separated by only 0.15 Mpc thus making an interaction between them ∼ 1 Gyr ago possible. This is supported by the galaxy's unusual gas dynamics (Kobulnicky & Skillman 1995) and an elliptical young halo stellar population with ages of 108 - 109 yrs (Caldwell & Phillips 1989). Could similar interactions be responsible for triggering the star formation we see in our transition dwarf sample? First, are there any galaxies nearby with which the transition dwarfs could have interacted? To examine this we must first estimate when stars last formed in the outer regions of these galaxies. We currently only have one color, (B-R), from short exposures obtained with the 1.3 m McGraw-Hill telescope at Michigan-Dartmouth-MIT Ob- servatory in addition to the deep R-band data presented here. Although more colors would be beneficial, we can roughly estimate the minimum time since stars formed in the red outer regions of the galaxies using these data and spectral evolutionary synthesis models. An estimate of the minimum time passed since the population of stars in the outer disk formed was made using models from the Gottingen Galaxy Evolution (GALEV) Group courtesy of Dr. Uta Fritze-v. Alvensleben. These are spectral evolutionary synthesis models for several metallicities and include gaseous emission (Schulz et al. 2002, Anders & Fritze-v. Alvensleben 2003). We use data from these models to plot the variation of (B-R) color with time for a galaxy which has experienced a burst of star formation. An estimate of the time since stars last formed in the outer galaxy envelope can then be made based on the time elapsed since the model burst for the color of the simulated galaxy to fade to the observed outer region color. This was done using GALEV models with a metallicity of z=0.008 for the underlying galaxy and a burst of solar metallicity (z=0.02). Constant and exponential star formation rates were explored for the underlying galaxy model. In both cases a burst with a simple stellar population was placed on top of the underlying galaxy population. In addition, models for different burst strengths were also examined. We find the model post-burst colors to be too blue for the case of constant star formation rate. Observed integrated (B-R)0 colors for our galaxies range from 0.84 to 1.2 (see Table 1), while the model post-burst B-R color only gets as red as B-R=0.7 by 6 Gyr after the burst. Modeling the underlying galaxy with an exponential star formation rate of the form (SF R ∝ e−t/1Gyr) produces post-burst colors which are a much better fit to our outer colors data. Changing the strength of the burst however, does not significantly affect our age estimate. As a result of these models, we find that at least between 2 and 6 Gyr have passed since star formation ceased in the outer regions of the galaxies studied here. An example of one of our models is shown in Figure 3 with (B-R) indicated for the envelope of NGC 3353. These results are consistent with ages found for the underlying stellar population of NGC 3353 by -- 10 -- Cair´os et al. (2007). Although these galaxies have no obvious close companions with which they are clearly interacting, they are not completely isolated. The galaxy which visually appears near IC 745 in Figure 1 is actually a background object. Assuming a relative velocity of 200 km/s between galaxies in the loose group environments in which the transition dwarfs reside, in the estimated time elapsed since the outer regions last experienced star formation, they could now be separated by about ∼0.5 Mpc. Several other galaxies can be found within 1 Mpc of each of our sample galaxies, and smaller gas-rich companions may be relatively common but difficult to locate in optical data (Taylor et al. 1995, 1996, Hogg et al. 1998, Noeske et al. 2001, Pustilnik et al. 2001). Any proposed mechanism for producing these dwarfs must be able to explain the ob- served low gas content and centrally concentrated star formation as well as envelope structure and colors. A likely scenario therefore is a mechanism through which some gas can be re- moved from the galaxy and much of the remaining gas falls into the center. An interaction could be a way to achieve this. For example, Bushouse (1987) found that the majority of interaction induced star formation is concentrated near the nuclei of disk galaxies and that gas depletion times are lower than for isolated spirals. For the gas to be moved toward the center, it must lose angular momentum and dissi- pate energy. Numerical simulations of spiral arms and bars tidally induced in disk galaxies (e.g. Noguchi & Ishibashi 1986, Noguchi 1987) and Noguchi (1988) showed that the gas is also affected, whereby the stellar bar induces the infall of gas toward the galactic nucleus. However, most of the transition dwarfs examined here do not appear to have such features, although it is uncertain what the bar lifetime might be for such objects. An exception to this is NGC 3353, which has star formation occurring in an elongated arc, as seen in Hα imaging (Dellenbusch et al. 2007 in prep.), that is not as centrally concentrated as in the other sample galaxies. This may be a bar-like structure. In addition, van Zee et al. (2001) found the HI in many BCDs, which have characteristics in common with transition dwarfs, to have intrinsically low angular momenta. Another related scenario to explain the evolution of transition galaxies is a dwarf-dwarf merger. Although dwarf galaxy mergers with larger systems have been modeled (e.g. Hern- quist 1989, Mihos & Hernquist 1994), the results of which can also describe a dwarf galaxy accreting a cloud for example, there is little discussion of dwarf-dwarf merger simulations in the literature. This is no doubt at least in part because they occur much less frequently than other types of merger events. However, there is evidence that they take place. One example is II Zw 40, a BCD which is thought to be an on-going merger of two gas-rich dwarf galaxies. Optically, II Zw 40 is dominated by a starburst with two faint tails extending outward (Sar- -- 11 -- gent & Searle 1970, Baldwin, Spinrad, & Terlevich 1982). It's HI extends several times the optical diameter with distinct tidal tails extending even further. As van Zee, Skillman, & Salzer (1998) note, the morphology of the system, a strong color gradient indicating the low surface brightness optical tidal tails are from an older stellar population, and no evidence for a second optical galaxy along the HI tidal tails indicate II Zw 40 is likely a merger rather than just a tidal interaction. Tidal interactions have often been invoked as a mechanism by which star formation can be triggered in a galaxy as well as an explanation for the high star formation rates found in BCD's (e.g Brinks & Klein 1988). Another possible dwarf-dwarf merger is the galaxy He 2-10 which is a Wolf-Rayet galaxy with a primarily elliptical appearance and two bright star clusters. It also exhibits a tail-like feature in both HI and CO. Kobulnicky et al. (1995) suggest, based on HI and CO kinematics, that He 2-10 is an advanced merger of two dwarf galaxies. A third well known example of two dwarfs interacting are the Small and Large Magellanic Clouds. Could the transition dwarfs which we are studying be a more evolved version of a dwarf-dwarf merger where it is no longer morphologically apparent a merger has taken place? Although dwarf-dwarf mergers occur in nature and may provide an explanation for at least some transition dwarfs, they are rare. This is to be expected because dwarfs have small interaction cross-sections. Also, for a merger to actually take place, the two interactors must have very low relative velocities. As galaxies with transition dwarf characteristics also seem to be rare, it is possible we are seeing an uncommon form of dwarf-dwarf merger in isolated environments with energy dissipation playing an important role in the subsequent evolution of these systems. Although the transition galaxies discussed here do not appear very disturbed and have regular outer isophotes, we cannot rule out an interaction or merger as playing an important role in their evolutionary and star formation histories. Several dynamical crossing times have passed since star formation occurred in the outer envelope, and it is therefore reasonable for minimal evidence of tidal interaction to remain today. On the other hand, the very long time scales that we find seem to be at odds with the intrinsically short time scales expected for starburst development in small galaxies (e.g. Recchi et al. 2002). 5. Summary & Conclusions In summary, we have presented results of deep R-band imaging for five starbursting transition dwarf galaxies. Our observational results show: -- 12 -- 1. All five galaxies exhibit smooth elliptical outer isophotes. 2. We find no evidence of extended tidal debris or other indications of a recent major interaction (e.g. tails, shells or ripples). 3. Fine structure exists in the central regions that are associated with the ISM and star formation, with dust structures and HII regions being prominent in several of the galaxies. 4. Outer envelope colors are consistent with having no star formation in the outer regions for the past several Gyr. This timescale could be longer if star formation slowly declined rather than stopping suddenly. We consider two possible classes of evolutionary histories which could be consistent with the observed properties and locations of transition dwarf galaxies: an internal ISM instability and a past interaction or merger event. The interaction or dwarf-dwarf merger scenario remains a possibility, although we find no strong evidence to support it. After several billion years it is not clear if we should still find evidence of tidal debris in the outer regions of the galaxies. However, long term products of dwarf interactions and dwarf-dwarf mergers in particular have not been well modeled (e.g. the effects of large dark matter contents on merger products). The other possibility of an unstable ISM seems promising to explain the central star formation in these objects. In this case the eventual evolution to an inactive state will be slower, possibly requiring several starburst cycles. Observationally the model can be tested by looking for the young stellar populations in fading starburst phases in dE-like dwarfs. A remaining question, which is particularly puzzling, If these objects last experienced significant star formation in their outer regions & 1 Gyr ago, why are they currently undergoing bursts of centrally concentrated star formation? What mechanisms control the structure of the ISM such that these galaxies now support rapid star formation, and will these circumstances lead to galaxies with little or no star formation, i.e. objects resembling dEs, within the next ∼ 1 Gyr? is the issue of timescales. We would like to thank P. Mucciarelli for assisting with the observations. In addition we wish to thank H. Schweiker and J. Davies for their assistance and dedication to the WIYN 0.9m telescope. We also thank U. Fritze-von Alvensleben for the use of the GALEV models and helpful discussions regarding them. This research is supported in part by the University of Wisconsin-Madison Graduate School. -- 13 -- REFERENCES Anders, P. & Fritze-v. Alvensleben, U. 2003, A&A, 401, 1063 Baldwin, J. A., Spinrad, H., & Terlevich, R. 1982, MNRAS, 198, 535 Barnes, J. E. & Hernquist, L. 1992, ARA&A, 30, 705 Bouchard, A., Jerjen, H., Da Costa, G. S., Ott, J. 2005, AJ, 130, 2058 Brinks, E. & Klein, U. 1988, MNRAS, 231, 63 Burstein, D, Krumm, N., & Salpeter, E. E. 1987, AJ, 94, 883 Bushouse, H. A. 1987, ApJ, 320, 49 Cairos, L. M., Caon, N., Garcia Lorenzo, B., Monreal-Ibero, A., Amorin, R., Weilbacher, P., Papaderos, P., 2007, ApJ, in press. Caldwell, N. & Phillips, M. M. 1989, ApJ, 338, 789 Conselice, C. J., Gallagher, J. S., Wyse, R. F. G. 2003, AJ, 125, 66 Deharveng, J.-M., Jedrzejewski, R., Crane, P., Disney, M. J., Rocca-Volmerange, B. 1997, A&A, 326, 528 Dellenbusch, K. E., Gallagher, III, J. S., & Knezek, P. M. 2007, ApJ, 655, L29 del Rio, M. S., Brinks, E., Cepa, J. 2004, AJ, 128, 89 Drinkwater, M. & Hardy, E. 1991, AJ, 101, 94 Forbes, D. A. & Thomson, R. C. 1992, MNRAS, 249, 779 Gallagher, J. S. & Hunter, D. A. 1989, AJ, 98, 806 Gil de Paz, A. & Madore, B. F. 2005, ApJS, 156, 345 Gordon, M. A. 1991, ApJ, 371, 563 Gordon, D. & Gottesman, S.T. 1981, AJ, 86, 161 Grebel, E. K., Gallagher, III, J. S., & Harbeck, D. 2003, AJ, 125, 1926 Hensler, G., Theis, C., Gallagher, III, J. S. 2004, A&A, 426, 25 Hernquist, L. 1989, Nature, 340, 687 -- 14 -- Hogg, D. E., Roberts, M. S., & Knezek, P. M. 1998, AJ, 115, 502 Knezek, P. M., Sembach, K. R., & Gallagher, III, J. S. 1999, ApJ, 514, 119 Kobulnicky, H.A., Dickey, J.M, Sargent, A.I., Hogg, D.E., & Conti, P.S. 1995, AJ, 110, 116 Kobulnicky, H. A. & Skillman, E. D. 1995, ApJ, 454, L121 Jackson, J., Snell, R. L., Ho, P. T. P., & Barrett, A. H. 1989, ApJ, 337, 680 Lake, G. & Schommer, R. A. 1984, ApJ, 280, 107 Lee, J. C., Salzer, J. J., Impey, C., Thuan, T. X., & Gronwall, C. 2002, AJ, 124, 3088 Li, J. G., Seaquist, E. R., Wang, Z., Sage, J. J. 1994, AJ, 107, 90 Lisker, T., Grebel, E. K., & Binggeli, B. 2006, AJ, 132, 497 Lisker, T., Grebel, E. K., Binggeli, B., & Glatt, K. 2007, ApJ, 660, 1186 Loose, H. H., Krugel, E., Tututkov. A. V. 1982, A&A, 105, 342 Marlowe, A. T., Meurer, G. R., & Heckman, T. M. 1999, ApJ, 522, 183 Martin, C. L. 1997, ApJ, 491, 561 Martin, M. 1998, A&AS, 131, 77 Mateo, M. L. 1998, ARA&A, 36, 435 Mihos, J. C. & Hernquist, L. 1994, ApJ, 425, L13 Miwa, T. & Noguchi, M. 1998, ApJ, 499, 149 Noeske, K. G., Iglesias-P´aamo, J., V´ıchez, J. M., Papaderos, P. & Fricke, K. J. 2001, A&A. 371, 806 Noeske, K. G., Papaderos, P., Cair´os, L. M., & Fricke, K. J. 2005, A&A, 429, 115 Noguchi, M. 1987, MNRAS, 228, 635 Noguchi, M. 1988, A&A, 203, 259 Noguchi, M. 2001, ApJ, 555, 289 Noguchi, M. & Ishibashi, S. 1986, MNRAS, 219, 305 -- 15 -- Ott, J., Walter, F., & Brinks, E. 2005, MNRAS, 358, 1453 Pelupessy, F. I., van der Werf, P. P., & Icke, V. 2004, A&A, 422, 55 Pustilnik, S. A., Kniazev, A. Y., Lipovetsky, V. A., & Ugryumov, A. V. 2001, A&A, 373, 24 Pustilnik, S. A., Martin, J.-M., Huchtmeier, W. K., Brosch, N, Lipovetsky, V. A., Richter, G. M. 2002, A&A, 389, 405 Recchi, S., Matteucci, F., D'Ercole, A., & Tosi, M. 2002, A&A, 384, 799 Reda, F. M., Forbes, D. A., Beasley, M. A., O'Sullivan, E. J., & Goudfrooij, P. 2004, MNRAS, 354, 851 Roberts, M. S. & Haynes, M. P. 1994, ARA&A, 32, 115 Sadler, E. M., Oosterloo, T. A., Morganti, R., Karalas, A. 2000, AJ, 119, 1180 Sandage, A. & Hoffman, G. L. 1991, ApJ, 379, L45 Sargent, W.L.W. & Searle, L. 1970, ApJ, 162, 155 Schulz, J., Fritze-v. Alvensleben, U., & Fricke, K. J. 2002, A&A, 392, 1 Searle, L., Sargent, W. L. W., & Bagnuollo, W. 1973, ApJ, 179, 427 Silk, J., Wyse, R. F. G., & Shields, G. A. 1987, ApJ, 322, L59 Sung, E.-C., Han, C., Ryden, B. S., Chun, M.-S., & Kim, H.-I. 1998, ApJ, 499, 140 Tajiri, Y. Y. & Kamaya, H. 2002, A&A, 389, 367 Taylor, C. L., Brinks, E., & Skillman, E. D. 1993, AJ, 105, 128 Taylor, C. L., Brinks, E., Grashuis, R. M. & Skillman, E. D. 1995, ApJS, 99, 427. Taylor, C. L., Thomas, D. L., Brinks, E. & Skillman, E. D. 1996, ApJS, 107, 143 van den Bergh, S. 1977, Evolution of Galaxies and Stellar Populations, Proceeding of a Conference at Yale University, 19 van den Bergh, S. 1980, PASP, 92, 122 van Zee, L., Salzer, J. J., Skillman, E. 2001, AJ, 122, 121 van Zee, L., Skillman, E.D., & Salzer, J.J. 1998, AJ, 116, 1186 -- 16 -- White, S. D. M. & Frenk, C. S. 1991, ApJ, 379, 52 This preprint was prepared with the AAS LATEX macros v5.2. -- 17 -- Table 1. Galaxy Properties Galaxy D(Mpc)† SFR (M⊙yr−1) LB ( x 108L⊙) MHI /LB 12 + log(O/H) (B-R)0 ‡ IC 745 NGC 3265 NGC 3353 NGC 3928 NGC 3773 15.3 17.6 12.6 13.2 13.2 0.2 0.1 0.8 0.1 0.1 4.9 7.8 7.8 3.6 6.5 0.05a 0.10b 0.30c 0.33d 0.08e 9.0 9.2 8.4 9.1 8.8 0.92 1.19 0.84 1.19 0.88 †A value of H0 = 75 km/s/Mpc is used to calculate distances ‡Outer region color; integrated colors will be bluer aMartin 1998 bLake & Schommer 1984 cGordon & Gottesman 1981 dLi et al. 1994 eBurstein, Krumm, & Salpeter 1987 -- 18 -- Fig. 1. -- Residual negative images in R-band of galaxies: A) IC 745, B) NGC 3265, C) NGC 3353, D) NGC 3928, and E) NGC 3773. Dust features appear as lighter regions in these negative images. Because of the intense star formation there, the inner ∼ 5 arcseconds of each galaxy were not included in the model and therefore appear as dark regions. -- 19 -- Fig. 2. -- Deviation from perfect ellipticity, described by the Fourier a4 parameter, as a function of distance along the semi-major axis for galaxies A) IC 745, B) NGC 3265, C) NGC 3353, D) NGC 3928, and E) NGC 3773. a4 > 0 describe disky isophotal shapes while a4 < 0 describe boxy isophotes. Although our images are moderately deep, S/N becomes low at outer isophotes and therefore we terminate the plot at ∼ 30 arcseconds for the two smaller galaxies. -- 20 -- Fig. 3. -- Illustrative spectral evolution synthesis model including an underlying elliptical galaxy model having an exponential star formation rate with a simple stellar population starburst occurring after 9 Gyr (Solid Curve). Horizontal dashed line indicates B-R color of the quiescent outer envelope of NGC 3353. The model indicates that for the color to have faded to B-R=0.84, after an instantaneous burst of star formation, 2.5 Gyrs have passed (Vertical Dashed Line).
astro-ph/0408248
1
0408
2004-08-13T01:50:38
The Anatomy of Star Formation in NGC 300
[ "astro-ph" ]
The Spitzer Space Telescope was used to study the mid-infrared to far-infrared properties of NGC 300, and to compare dust emission to Halpha to elucidate the heating of the ISM and the star formation cycle at scales < 100 pc. The new data allow us to discern clear differences in the spatial distribution of 8 micron dust emission with respect to 24 micron dust and to HII regions traced by the Halpha light. The 8 micron emission highlights the rims of HII regions, and the 24 micron emission is more strongly peaked in star forming regions than at 8 microns. We confirm the existence and approximate amplitude of interstellar dust emission at 4.5 microns, detected statistically in Infrared Space Observatory (ISO) data, and conclude it arises in star forming regions. When averaging over regions larger than ~ 1 kpc, the ratio of Halpha to Aromatic Feature emission in NGC 300 is consistent with the values observed in disks of spiral galaxies. The mid-to-far-infrared spectral energy distribution of dust emission is generally consistent with pre-Spitzer models.
astro-ph
astro-ph
The Anatomy of Star Formation in NGC 300 G. Helou,1 H. Roussel,2 P. Appleton,1 D. Frayer,1 S. Stolovy,1 L. Storrie-Lombardi,1 R. Hurt,1 P. Lowrance,1 D. Makovoz,1 F. Masci,1 J. Surace,1 K.D. Gordon,3 A. Alonso-Herrero,3 C.W. Engelbracht,3 K. Misselt,3 G. Rieke,3 M. Rieke,3 S.P. Willner,4 M. Pahre,4 M.L.N. Ashby,4 G.G. Fazio,4 H.A. Smith4 ABSTRACT The Spitzer Space Telescope was used to study the mid-infrared to far-infrared properties of NGC 300, and to compare dust emission to Hα to elucidate the heating of the ISM and the star formation cycle at scales < 100 pc. The new data allow us to discern clear differences in the spatial distribution of 8 µm dust emission with respect to 24 µm dust and to H II regions traced by the Hα light. The 8 µm emission highlights the rims of H II regions, and the 24 µm emission is more strongly peaked in star forming regions than 8 µm. We confirm the existence and approximate amplitude of interstellar dust emission at 4.5 µm, detected statistically in Infrared Space Observatory (ISO) data, and conclude it arises in star forming regions. When averaging over regions larger than ∼ 1 kpc, the ratio of Hα to Aromatic Feature emission in NGC 300 is consistent with the values observed in disks of spiral galaxies. The mid-to-far-infrared spectral energy distribution of dust emission is generally consistent with pre-Spitzer models. Subject headings: galaxies: (NGC 300) ISM - infrared: galaxies - individual: galaxy 4 0 0 2 g u A 3 1 1 v 8 4 2 8 0 4 0 / h p - o r t s a : v i X r a 1. Introduction NGC 300 is a SA(s)d galaxy in the Sculptor Group of galaxies, at a distance of about 2.1 Mpc (Freedman et al. 1992), viewed at an inclination of about 50◦ (Puche et al. 1990). 1 2 3 4 SIRTF Science Center, California Institute of Technology, M.S. 220-6, Pasadena, CA 91125 California Institute of Technology, M.S. 320-47, Pasadena, CA 91125 Steward Observatory, University of Arizona, Tucson, AZ 85721 Harvard-Smithsonian Center for Astrophysics, 60 Garden St., Cambridge, MA 02138 – 2 – Its total luminosities in the blue band (3855–4985A) and in the far-infrared (42.5–122.5 µm, estimated from Spitzer data) are ∼ 3.3 × 108 L⊙ bol and 2.2 × 108 L⊙ bol respectively. Its LFIR/LB ratio is thus very close to the average ratio for the blue magnitude-limited sample studied by Thuan & Sauvage (1992). Due to its large angular extent and low surface bright- ness, no reliable total radio continuum flux measurement exists (only discrete sources are detected). NGC 300 has a striking appearance in the visible and in Hα due to several H II regions with nearly circular shapes and various degrees of filling-in (Deharveng et al. 1988; Hoopes et al. 1996; see Figure 1). Its large HI envelope (Puche et al. 1990) extends well beyond the visible image. Pannuti et al. (2000) reported a total of 44 SNR candidates, evi- dence that the current star formation activity manifested as H II regions has been on-going for tens of millions of years (see also Butler et al. 2004). Because of its proximity, NGC 300 allows the Spitzer Space Telescope (Werner et al. 2004) to compare dust emission to other ISM components and discern the interplay between the ISM and the star formation cycle at scales < 100 pc. 2. Observations and Data Reduction The IRAC (Fazio et al. 2004) observations of NGC 300 are 12-second frames and map the galaxy in approximately half array spacings, yielding a total time per sky position of 48 s. They were reduced with the standard Spitzer Science Center data reduction pipeline (version 9.5). Due to the readout of bright point sources, the signal was reduced (pulled-down) in some columns of the array1; a correction measured from the map was applied. Persistent images left by bright sources were found by median-combining all the dithered positions together and identifying any remaining sources. These objects were removed from individual frames. The data were then combined into a mosaic using a cosmic ray rejection and a background matching applied between overlapping fields of view. The relative photometric uncertainty is of the order of 5%, and the absolute uncertainty is 10%; uncertainty tied to the angular sizes of source and measuring aperture contributes up to 15%. Images of NGC 300 at 24, 70, and 160 µm were obtained with the MIPS Instrument (Rieke et al. 2004) in the scan-map mode. The final mosaics have a total exposure time of approximately 160, 80, and 16 s per point at 24, 70, and 160 µm. The MIPS images were reduced using the MIPS Instrument Team Data Analysis Tool (Gordon et al. 2004) as described by Engelbracht et al. (2004). The uncertainties on the final absolute calibrations are estimated at 10%, 20%, and 20% for the 24, 70, and 160 µm data, respectively. The 1see the Spitzer Observer's Manual at http://ssc.spitzer.caltech.edu/documents/som/. – 3 – 70 and 160 µm images exhibit linear streaks along the scan direction which are residual instrumental artifacts due to the time dependent responsivity of the Ge detectors, and affect the photometry in large apertures. The Hα (6563A) + [N II] (6583 and 6548A) map was derived from images posted in NED by Larsen & Richtler (1999). The narrowband image containing Hα emission and the R-band image were aligned, and rescaled to subtract the stellar continuum, deriving the scaling factor from aperture photometry on 32 bright stars. Residuals from saturated or improperly subtracted stars were masked out, and the final image was regridded and rotated to match the IRAC reference frame. The Spitzer and Hα+[N II] images are shown in Figure 1. The images contain three very bright Milky Way stars directly superposed on the disk. We masked them out in IRAC and 24 µm images, as well as additional point sources as described in Section 3.1 below (Figure 4b). Foreground stars will be removed more rigorously in a later paper. Fainter, indistinguishable stars are unlikely to affect significantly the results presented here. 3. Physical Content in Mid-Infrared Images Whereas at λ ≥ 24 µm the emission is dominated by interstellar dust, the mid-infrared marks the transition from stellar to interstellar emission. The appearance of the galaxy shifts from resembling the stellar disk in the 3.6 µm image to an ISM-like appearance at 8 µm. We assume conservatively the 3.6 µm emission is all due to stars, and extrapolate this component to longer wavelengths using stellar population modeling from Starburst99 (Leitherer et al. 1999); the stellar Spectral Energy Distribution (SED) longward of 2 µm depends very little on star formation history or metallicity, as illustrated in Figure 2. The decomposition amounts to scaling the 3.6 µm map by 0.596, 0.399, 0.232 and 0.032 respectively for 4.5, 5.8, 8 and 24 µm, and subtracting these from the observed maps pixel by pixel to yield what we will refer to as the "dust maps" shown in Figure 1. Based on this decomposition, the 24 µm map is probably ≥ 98% interstellar emission globally, and ≥ 93% interstellar in the inner arcmin, where the stellar contribution is the greatest. The 8 µm map is ≥ 81% interstellar emission. This method potentially overestimates the dust component by underestimating the red- der SED of very low mass stars. However the model reproduces adequately the 4.5 µm/3.6 µm color of the nuclear stellar cluster, of the emission averaged in the inner arcmin (Figure 2), and of the diffuse emission outside several disk H II regions. On the other hand, the method probably underestimates the dust component by as- suming that the 3.6 µm map contains only photospheric emission, ignoring evidence that the – 4 – hot dust component stretches down to λ < 3 µm (Bernard et al. 1994; Lu et al. 2003; Hunt et al. 2002). From the ISO data in Figure 6 of Lu et al. (2003), this underestimate may reach a factor of up to 2.6, 10–20% and 3–4%, respectively at 4.5, 5.8 and 8 µm. In view of these opposing potential biases, our conservative approach should be reliable qualitatively for NGC 300, in spite of localized departures due to variations in stellar population colors or dust extinction. 3.1. Interstellar Dust Emission In order to constrain the origin of the 4.5µm non-stellar emission, we identified pix- els where the excess is above the 3σ significance threshold in the 4.5µm "dust map", and examined their colors and locations. In scatter plots of these pixel values vs. Hα or vs. 8 µm dust emission, a correlation branch is evident in both cases with moderate slopes in fν(4.5 µm)/fν(8 µm) and fν(4.5 µm)/Hα (Figure 3), as well as a branch with very steep, al- most vertical slopes. The pixels in the latter branch are strongly associated with the 3.6 µm emission, including the brightest foreground star and a large number of point sources spread almost uniformly accross the whole map. We use this excess map as a mask to efficiently remove foreground stars, with additional hand masking of residues around very bright stars. On the other hand, the branches with moderate 4.5 µm/8 µm and 4.5 µm/Hα ratios map into bright star formation regions, as might be expected (Figure 4a). The global ratio νfν(4.5 µm)/νfν(8 µm) for the selected dust emission amounts to ∼ 4%. However, summing the pixels in star forming regions with large 4.5 µm excess, dust alone accounts for 17% of the total 4.5 µm emission, probably more if dust contribution to the 3.6 µm emission were accounted for. Two uncertainties remain which will be addressed in future papers: Whether redder SEDs of young stars affect significantly the results in star forming regions; and whether the 4.5µm dust associates more closely with ionized regions or with Photo- Dissociation Regions (PDR). The dust maps at 5.8 and 8 µm should be dominated by the Aromatic Features in Emission (AFE) proposed by Puget & L´eger (1989)) to explain IRAS data, and well studied with ISO (Boulanger et al. 1998; Helou et al. 2000; Lu et al. 2003). The value and variability of the ratios between these bands are of considerable interest in constraining the physical state of the emitters (e.g. de Frees et al. 1993). Lu et al. (2003) found the AFE profiles among galaxies constant to within measurement uncertainty of 15 to 25% in the range 5 to 12µm; the Spitzer images can extend this finding to much greater accuracy, and could point to variations within disks on a variety of scales. – 5 – We find that νfν(5.8 µm)/νfν(8 µm) for dust emission alone is constant to within the uncertainties, with a value of 0.50± 0.09 (3σ) over most of the area where surface brightness is above the 6σ level in each of the 5.8 µm and 8 µm dust maps. 3.2. Comparison to Hα Figure 4c shows a comparison between Hα and 8 µm maps at a resolution of 2×FWHM ∼ 5.15′′ ∼ 50 pc. In the outer disk, several H II regions as well as more complex structures are visible in the Hα map, with the corresponding 8 µm profile flatter and sometimes significantly offset from the Hα peak. In spite of the complex structure, Figure 4c-d clearly shows the 8 µm emission highlighting the rims of H II regions. The ratio 8 µm/Hα goes up by a factor five to ten from inside to just outside the H II region. This ratio may increase by another factor of 2–4 in diffuse regions, but this is uncertain because of limited sensitivity and inaccurate continuum subtraction in the Hα map. While this behavior is not a complete surprise, the Spitzer data are the first to show it so clearly in a nearby galaxy. Boulanger et al. (1990), Sellgren et al. (1990), Giard et al. (1994) and others have reported sharp boundaries to the AFE, with peaks on molecular cloud surfaces outside ionized regions, and related this to a combination of UV excitation and photo-processing of the AFE carriers. Helou et al. (2001) also speculated that AFE arise primarily in PDRs based on physical arguments rather than direct observations. Inside the H II regions, the Aromatics are depressed probably because of destruction by the ionizing UV (Boulanger et al. 1990). The sharp contrasts near H II regions between 8 µm and Hα are only a small-scale property. Summing over scales larger than about 1 kpc, the ratio between the two tracers is constant to about 40%. The global ratio of 8 µm to unextincted Hα is roughly 30% smaller than the average measured in disks of spiral galaxies by Roussel et al. (2001a), which had itself a dispersion of 50% among disks. For this comparison, the total Hα+[N II] flux from Hoopes et al. (1996) was corrected for an extinction A(Hα) = 0.37 and an Hα/(Hα+[N II]) ratio of 0.91, averaged from values of individual H II regions in Webster & Smith (1983). The 8 µm IRAC flux was converted to 7 µm flux in the ISO filter LW2 assuming the same spectral shape as in the disk of M 83 (Roussel et al. 2001b), which was verified to reproduce accurately the 5.8 µm/8 µm dust ratio of NGC 300. The star formation rate in the disk of NGC 300 is 0.08–0.11 M⊙ yr−1 from the 8 µm emission (as calibrated for solar-metallicity spiral disks by Roussel et al. 2001a), and ∼ 0.14 M⊙ yr−1 from Hα (as calibrated by Kennicutt 1998). The agreement points to 8 µm under these conditions as a viable proxy for ionizing flux, and therefore for on-going massive star formation rate. Since the association between 8 µm and Hα is not at the atomic process – 6 – level however, it will be subject to geometric and other effects, and the use of 8 µm as such a proxy is unlikely to be valid under all conditions. Figure 4e-f also shows the 24 µm emission is more peaked than 8 µm in star forming regions. The spatial resolution is insufficient to distinguish whether PDRs or ionized regions are more closely associated with the 24 µm peaks. This confirms earlier suggestions based on global SED analysis that 24 µm traces heating by the youngest and most massive stars in a galaxy (Helou 2000). 4. Mid-IR to Far-IR SED In order to study the SED at λ ≥ 8 µm, we identified nineteen emission regions in the dust maps, most centered on a local peak (Figure 5l). After smoothing all maps to the resolution of 160 µm, we measured the emission of each region in a circular aperture 2.5 × FWHM(160 µm) = 95′′ ∼ 1 kpc in diameter. The resulting photometry is plotted in Figure 5 as various color ratios against fν(8 µm)/fν(24 µm), and compared to the ratios expected from pre-Spitzer model SEDs of galaxies (Dale & Helou 2002). These are single- parameter models of dust emission integrated over whole galaxies, and incorporate a power- law distribution of dust over heating intensities. The data and model agree in the trends, though there are systematic offsets in the values. These offsets are most simply understood as a systematic discrepancy of ∼ 30% at 160 µm and ∼ 15% at 70 µm between model and measurement, assuming reliable photometry at 8 and 24 µm. In view of the simplicity of the models, photometric uncertainties (especially non-linearity effects) and source extent, the agreement between model and data is quite satisfactory, and suggests that beyond a certain scale portions of galaxies may approach full galaxies in their behavior. The 160 µm and less so the 70 µm maps display a striking large halo which echoes several features of the HI envelope. It is unclear at this point whether this reflects unusually cold material in the outer disk, or whether the larger 160 µm beam is simply more sensitive to low surface brightness features. 5. Summary The Spitzer Space Telescope data allow us more clearly than ever to dissect the inter- stellar dust emission components in NGC 300 and to relate them to star formation. The 8 µm AFE highlights the rims of the H II regions defined by Hα, confirming it is more closely associated with PDRs than with ionized regions. When averaging over regions larger than – 7 – ∼ 1 kpc, the ratio of Hα to AFE emission in NGC 300 is consistent with the disk value observed in other galaxies. This confirms AFE as a convenient massive star formation rate estimator in disks of galaxies. The mid-IR spectral signature of the AFE is invariant as measured by the ratio fν(5.8 µm)/fν(8 µm), which is essentially constant at ∼ 50 pc reso- lution at the ∼ 20% (3σ) measurement accuracy over most of the disk of NGC 300. The 24 µm emission is more strongly peaked than 8 µm in star forming regions. Interstellar dust emission at 4.5 µm is confirmed at about the expected amplitude; for the first time, its spa- tial distribution associates it clearly with star forming regions. The data cannot distinguish between PDRs and H II regions as the primary origin for either 4.5 or 24 µm dust emission. The SED models of (Dale & Helou 2002) apply to ∼kpc-sized portions in the disk to within the current photometry and calibration uncertainties of 20–30%. The Spitzer Space Telescope is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Adminis- tration. This research has made use of the NASA/IPAC Extragalactic Database which is operated by JPL/Caltech, under contract with NASA. REFERENCES Bernard, J.P., Boulanger, F., D´esert, F.X., Giard, M., Helou, G., & Puget, J.L. 1994, A&A, 291, L5 Boulanger, F., Boissel, P., Cesarsky, D., & Ryter, C. 1998, A&A, 339, 194 Boulanger, F., Falgarone, E., Puget. J.-L., & Helou, G. 1990, ApJ, 346, 136 Butler, D.J., Martinez-Delgado, D. & Brandner, W. 2004, AJ127, 1472 Dale, D.A., Helou, G., Contursi, A., Silbermann, N., & Kolhatkar, S. 2001, ApJ, 549, 215 Dale, D.A., & Helou, G. 2002, ApJ, 159, 576 de Frees, D.J., Miller, M.D., Talbi, D., et al. 1993, ApJ, 408, 530 Deharveng, L., Caplan, J., Lequeux, J., et al. 1988, A&AS, 73, 407 Engelbracht, C.W. et al. 2004, ApJS, this volume Fazio, G. et al. 2004, ApJS, this volume Freedman, W.L., Madore, B.F., Hawley, S.L., et al. 1992, ApJ, 396, 80 – 8 – Giard, M., Bernard, J. P., Lacombe, F., Normand, P., & Rouan, D. 1994, A&A, 291, 239 Gordon, K.D. et al. 2004, PASP, submitted Helou, G. 2000, in "Infrared Astronomy: Today and Tomorrow", ed. F. Casoli, J. Lequeux & F. David (Springer Verlag: Les Ulis) p. 337 Helou, G., Lu, N.Y., Werner, M.W., Malhotra, S., & Silbermann, N.A. 2000, ApJ, 532, L21 Helou G., Malhotra S., Hollenbach D.J., Dale D.A. & Contursi A. 2001, ApJ, 548, L73 Hoopes, C.G., Walterbos, R.A., & Greenawalt, B.E. 1996, AJ, 112, 1429 Hunt, L.K., Giovanardi, C., & Helou, G. 2002, A&A, 394, 873 Kennicutt, R.C. 1998, ARA&A, 36, 189 Larsen, S.S., & Richtler, T. 1999, A&A, 345, 59 Leitherer, C., Schaerer, D., Goldader, J.D., et al. 1999, ApJS, 123, 3 Lu, N., Helou, G., Werner, M.W., et al. 2003 ApJ, 588, 199 Pannuti, T.G., Duric, N., Lacey, C., et al. 2000, ApJ, 544, 780 Puche, D., & Carignan, C., & Bosma, A. 1990, AJ, 100, 1468 Puget, J.L. & L´eger, A. 1989, ARA&A, 27, 161 Rieke, G. H. et al. 2004, ApJS, this volume Roussel, H., Sauvage, M., Vigroux, L., & Bosma, A. 2001a, A&A, 372, 427 Roussel, H., Vigroux, L., & Bosma, A., et al. 2001b, A&A, 369, 473 Sellgren, K., Tokunaga, A.T., & Nakada, Y. 1990, ApJ, 349, 120 Thuan, T.X., & Sauvage, M. 1992, A&AS, 92, 749 Webster, B.L., & Smith, M.G. 1983, MNRAS, 204, 743 Werner, M. W. et al. 2004, ApJS, this volume This preprint was prepared with the AAS LATEX macros v5.2. – 9 – Fig. 1.- Images of NGC 300 obtained with the Spitzer Space Telescope and other sources. The left-most column shows observed maps at (a) 3.6 µm, (b) 4.5 µm, (c) 5.8 µm, (d) 8 µm. The center column shows (e) BVRHα composite image (credit MPG/ESO), where B is coded in blue, V in green, R and Hα in red; (f) Hα map described in Section 2; (g) dust map at 5.8 µm; (h) dust map at 8 µm. The right-most column shows (i) H I map from Puche et al. (1990); and observed maps at (j) 24 µm; (k) 70 µm; (l) 160 µm, with photometric apertures used in Figure 5 superposed. Each frame contains the peak surface brightness in the background-subtracted image; all images are in logarithmic scale, except the 160 µm image which is linear. North is about 40◦ clockwise from vertical up, and each frame is approximately 20′ on a side. – 10 – Fig. 2.- SED decomposition: The SEDs of NGC 300 in three different apertures are shown, with all fluxes normalized to 3.6 µm: the nuclear stellar cluster (clearly contaminated by dust emission judging from morphology in the IRAC maps), the inner arcmin, where dust contribution is relatively small, and the whole galaxy. The thick dashed lines trace model photospheric SEDs from 2 to 24 µm, derived from Starburst99 for a range of star formation histories and two metallicities, and transformed to IRAC filter fluxes. The horizontal error bars represent the filter widths at half maximum transmission. The Ks fluxes are lower limits, because the 2MASS map is not sensitive to the diffuse emission. Fig. 3.- Scatter plot of excess fν(4.5 µm) versus dust-only fν(8 µm), normalized by the noise in the original maps. Individual pixels are plotted, after convolution of all maps to a common angular resolution. The correlation branch was isolated by selecting pixels below the solid line (corresponding to Figure 4a). The uncorrelated emission, extending to much higher brightnesses, was cut here for clarity. The scatter plot versus Hα, not included, shows a similar correlation branch, most pixels of which map into the 8 µm correlation branch. – 11 – Fig. 4.- Top: Distribution of excess 4.5 µm emission, f e ν (4.5 µm), in units of the noise in the initial 4.5 µm map, σ4.5. (a): Part of f e ν (4.5 µm) which is correlated with 8 µm and Hα (950 pixels ∈ [3; 19] σ4.5. (b): Part of f e ν (4.5 µm) which is not correlated with 8 µm nor Hα, in logarithmic scale (4583 pixels ∈ [3; 8150] σ4.5). The contours are Hα isophotes (a) and the 4σ isophote of the 8 µm dust map (b). Bottom: Comparison of Hα and dust emission: (c) log(fν(8 µm) / Hα) (95% of pixels span 1.2 dex), showing low ratios inside prominent H II regions; (e) log(fν(8 µm) / fν(24 µm)) (95% of pixels span the linear range [0.6; 2.2]), showing again the distinctly low color ratios of H II regions; (d) 8 µm (red) and Hα (green) composite image; (f) 8 µm (red) and 24 µm (green) composite image. – 12 – 100.00 10.00 ) 8 / 0 7 ( R , ) 0 7 / , ) 4 2 / 0 7 ( R , ) 8 / 0 6 1 ( R , ) 4 2 / 0 6 1 ( R 0 6 1 ( R 1.00 0.50 1.00 1.50 2.00 2.50 R(24/8)=fnu(24um)/fnu(8um) Fig. 5.- Color-color plot for all Spitzer bands at λ ≥ 8 µm, for the regions measured with an aperture corresponding to ∼ 1 kpc in the plane of NGC 300, with the Dale & Helou (2002) model predictions superposed as curves. Each curve has one symbol at x=2.28 to indicate the corresponding color ratio. We corrected the 8 µm fluxes by a factor of √0.69 since the aperture is intermediate in size between the nominal point-source aperture and infinitely extended scales. The one point with error bars illustrates the uncertainties of 25% on the x- axis and 35% on the y-axis. Squares are for 160 µm/8 µm; X signs are for 160 µm/24 µm; plus signs are for 70 µm/8 µm; triangles are for 70 µm/24 µm; and circles are for 160 µm/70 µm. The systematic offset at the longer wavelength bands is discussed in the text.
astro-ph/0604492
2
0604
2007-09-19T08:08:07
Einstein and Jordan frames reconciled: a frame-invariant approach to scalar-tensor cosmology
[ "astro-ph", "gr-qc", "hep-ph", "hep-th" ]
Scalar-Tensor theories of gravity can be formulated in different frames, most notably, the Einstein and the Jordan one. While some debate still persists in the literature on the physical status of the different frames, a frame transformation in Scalar-Tensor theories amounts to a local redefinition of the metric, and then should not affect physical results. We analyze the issue in a cosmological context. In particular, we define all the relevant observables (redshift, distances, cross-sections, ...) in terms of frame-independent quantities. Then, we give a frame-independent formulation of the Boltzmann equation, and outline its use in relevant examples such as particle freeze-out and the evolution of the CMB photon distribution function. Finally, we derive the gravitational equations for the frame-independent quantities at first order in perturbation theory. From a practical point of view, the present approach allows the simultaneous implementation of the good aspects of the two frames in a clear and straightforward way.
astro-ph
astro-ph
Einstein and Jordan frames reconciled: a frame-invariant approach to scalar-tensor DESY 06 -- 076 cosmology Riccardo Catena Deutsches Elektronen-Syncrotron DESY, 22603 Hamburg, Germany INFN, Sezione di Padova, via Marzolo 8, I-35131, Padova, Italy Massimo Pietroni Luca Scarabello Dipartimento di Fisica Universit`a di Padova and INFN, Sezione di Padova, via Marzolo 8, I-35131, Padova, Italy Scalar-Tensor theories of gravity can be formulated in different frames, most notably, the Einstein and the Jordan one. While some debate still persists in the literature on the physical status of the different frames, a frame transformation in Scalar-Tensor theories amounts to a local redefinition of the metric, and then should not affect physical results. We analyze the issue in a cosmological context. In particular, we define all the relevant observables (redshift, distances, cross-sections, ...) in terms of frame-independent quantities. Then, we give a frame-independent formulation of the Boltzmann equation, and outline its use in relevant examples such as particle freeze-out and the evolution of the CMB photon distribution function. Finally, we derive the gravitational equations for the frame-independent quantities at first order in perturbation theory. From a practical point of view, the present approach allows the simultaneous implementation of the good aspects of the two frames in a clear and straightforward way. I. INTRODUCTION The evidence for Dark Energy has revived the interest in modifications to General Relativity (GR). Among these theories, Scalar-Tensor gravity (ST) [1] represents a good benchmark to accommodate new ultra-light degrees of freedom possibly responsible for the accelerated expansion of the universe. From a phenomenological point of view, it respects local Lorentz invariance and the universality of free fall of test bodies. Moreover, the post-Newtonian parameters γ − 1 and β − 1, parameterizing the deviations from GR, are expressed in terms of a single function, thus allowing a straightforward confrontation of the theory with solar system tests of gravity [2, 3, 4]. From a theoretical point of view, this class of theories is large enough to accommodate a vast range of possible extensions of GR in which new scalar fields are present in the gravitational sector; from extra-dimensional radions and string theory moduli, to f (R) theories of gravity. Moreover, in a ST context, ultralight scalar fields are technically natural. Indeed, general covariance implies that the contribution of radiative corrections from the (visible and dark) matter sector to the scalar field mass is at most of order Λ4/M 2 p , Λ being the cosmological constant. Thus, the lightness of the scalar field is just a manifestation of the smallness of the cosmological constant, or of the curvature of the universe [5]. Finally, in a cosmological setting, it has been pointed out that an intriguing attraction mechanism towards GR [6] could be operative under very generic conditions, including the case of a runaway potential suitable for DE [7, 8, 9]. Therefore, these theories may differ considerably from GR at high redshifts and at the same time fulfill the stringent bounds coming from solar system tests [4] today. ST theories can be formulated in different guises. In the so-called 'Jordan frame', the Einstein-Hilbert action of GR is modified by the introduction of a scalar field1 with a non-canonical kinetic term and a potential. This field replaces the Planck mass, which becomes a dynamical quantity. On the other hand, the matter part of the action is just the standard one. By Weyl-rescaling the metric, one can express the ST action in the so called 'Einstein Frame'. In these new variables, the gravitational action is just the Einstein-Hilbert one plus a scalar field with canonically normalized kinetic terms and an effective potential. On the other hand, in the matter action the scalar field appears, through the rescaling factor multiplying the metric tensor everywhere. As a consequence, the matter energy-momentum tensor is 1 ST theory can be generalized with the introduction of many scalar fields [1]. In order not to overload the notation, in this paper we will consider a single field, but our results are easily generalizable to the multi-field case. 2 not covariantly conserved, and particle physics parameters, like masses and dimensionful coupling constants appearing in the lagrangian are space-time dependent. It is a general fact that physics is invariant under a local redefinition of field variables -- in this case, the Weyl- rescaled metric. Nevertheless, this invariance is not fully exploited in the literature, where some confusion also exists about the physical status of the different frames. Most authors prefer to work in the Jordan frame, which is also referred to as the 'physical' one. The advantage of this frame is that all the particle physics' properties, i.e masses, coupling constants, decay rates, cross sections, etc. can be computed straightforwardly, since the matter action is just the standard one. On the other hand, the gravitational equations are more involved than in GR, since the scalar is non-trivially mixed to the metric tensor. Working in the Einstein frame is easier for what concerns the gravitational equations, but the connection with particle physics is not so direct as in the Jordan frame, since, for instance, the electron mass appearing in the lagrangian is space-time dependent. So, many authors use the Einstein frame as a mathematical tool to solve the field equations, and then translate back the results in the Jordan frame to compare with observations. For some recent applications along these lines, see, for instance [8, 10, 11, 12]. Non-linear approaches have also been discussed [13, 14]. While both these procedures are correct, a general discussion of the frame-invariance of physics in the ST theories in a cosmological setting is still missing. In particular, while it is quite straightforward to go back and forth from one frame to the other when the barotropic fluid approximation for matter holds, it is not so clear how to do it when Boltzmann equations have to be employed. For instance, the epoch of decoupling of some interaction in the expanding universe, expressed in conformal time, should be derived independently on the frame. But the standard rule of thumb, that is, Γ <∼ H/a , (1) with Γ the reaction rate and H the Hubble parameter, is not a frame-invariant relation. The purpose of this paper is to formulate a frame-invariant approach to discuss ST cosmology. Following Dicke [15, 24], we will start by the observation that the physical observables are dimensionless ratios between physical quantities and the appropriate units of measure2. These numbers are frame-invariant, and should therefore be expressible in terms of frame-invariant quantities. We will discuss the main observables in cosmology (redshift, distances, CMB temperature perturbations, . . . ) and particle physics (masses, cross sections, rates, . . . ) and express them in terms of frame-invariant combinations of the theory parameters and variables, and the units. We will discuss metric perturbations and define a frame-invariant phase space and distribution function. This will enable us to write down a frame-invariant Boltzmann equation to discuss processes relevant for cosmology, like particles freeze-out, the Sachs-Wolfe effect, and matter-radiation decoupling. Finally, we will derive the equations of motion for the frame invariant quantities at first order in metric perturbations. The plan of the paper is as follows. In sect. II we will introduce frame-transformations. In sect. III we will discuss how frame-invariant results can be obtained for rates, cross sections, and so on. The cosmological background observables, that is redshift and (angular and luminosity) distance will be discussed in sect. IV. Then, we will consider scalar perturbations in the Newtonian gauge. We will see that a frame transformation amounts to a change of gauge, and will therefore define frame-invariant scalar metric perturbations. This will enable us to define a frame-invariant phase space and energy-momentum tensor (sect. V). The Boltzmann equation will be introduced in sect. VI, where its application to particles' freeze-out and CMB photons will be outlined. Finally, in sect. VII, we will discuss the ST dynamics. We will write down the ST action in terms of frame-invariant quantities only, and write down the corresponding equations of motion at first order. In the appendix, the case of the synchronous gauge and of a generic gauge will be discussed. II. FRAME TRANSFORMATIONS In general, a frame transformation is a rescaling of the metric gµν, of the form with f = f (xµ) and real, and the space-time coordinates xµ (µ = 0, . . . , 3) are kept fixed. The diffeomorphism- invariant space-time interval then gets transformed as ds2 = gµνdxµdxν = e−2f gµνdxµdxν = e−2f ds2 . (3) gµν = e−2f gµν , (2) 2 Strictly speaking, this argument applies only to local physical quantities independent of metric derivatives. 3 Considering time-like and space-like intervals, we see that proper times, dt0 = ds(timelike), and proper lengths, dl0 = ds(spacelike), transform as dt0 = e−f dt0 dl0 = e−f dl0 . (4) The above transformations can be seen as the relations between the clocks and rods in two different systems of units [15, 16]. Notice that the above transformation laws are, in general, space-time dependent, since so is the function f (xµ). Such transformations are not common in GR and in standard physics in general, where they are of no practical use. However, in ST the situation changes dramatically. For instance, one could consider two different units of length, say, the size of an atom and the radius of a small Schwarzschild black hole. The ratio between these two physical lengths, which is constant in GR, is in general space-time dependent in ST. Therefore, in this kind of theories, the operational definition of units implies local transformations. Since ds2 = 0 is a frame-invariant condition, the speed of light is also invariant under the transformation (4). Moreover, we will impose that the Planck constant ¯h is also invariant. This amounts to transforming units of mass (and energy) as M = ef M , (5) and in leaving actions invariant (but, in general not covariant!). It should be emphasized that this is by no means a unique choice. One could for instance consider frame transformations in which ¯h varies and the Newton constant is kept fixed. However, our prescription is the one realized in ST theories, which are the main focus of this work. Since -- in the particular class of local transformations we are considering -- units of time, length, and inverse masses transform in the same way, we will consider a single dimensionful unit, length. The generic unit length will be indicated by lR. It transforms according to the space-time dependent relation of eq. (4) lR = e−f lR . (6) Then, a generic local physical quantity, Q, having the dimensions [Mass]a [Length]b, and [Time]c, will transform according to Q = ef (a−b−c) Q . (7) The scaling above holds as long as the quantity Q does not depend on derivatives of the metric. The choice of the unit length is dictated, as usual, by practical convenience. For instance, one could use a reference atomic wavelength, the inverse physical mass of some particle, or the Planck length. Physics does not depend on which particular clock or rod one adopts. In the following, we will discuss this independence in the framework of Friedmann-Robertson-Walker cosmology, but of course this is a general property [15]. III. FRAME-INVARIANT PARTICLE PHYSICS The particle physics action (the underlying theory being the Standard Model, or any of its extensions) has the form Z d4x√−g L =Z d4x √−g l4 R l4 R L . (8) Since both the action and the combination √−g/l4 construct a lagrangian L such that R are frame-invariant, so is the product l4 R L. It is convenient to l4 R L = l4 R L + ··· , (9) where the dots represents terms containing space-time derivatives of lR and lR. This is achieved transforming the parameters and fields in L as follows: R λn = ln ln R Aµ = Aµ , (10) where λn is a generic coupling of canonical dimension n (e.g. m2 φ2 ≡ λ2 φ2, h φ4 ≡ λ4 φ4, and so on), φ, ψ, and Aµ are scalar, spinor and vector fields, respectively, and γµ are the Dirac gamma matrices. lR γµ = lR γµ , lR φ = lR φ , l3/2 R ψ = l3/2 R ψ , λn , 4 In particular, the mass parameters appearing in the lagrangian can be constant in one frame and space-time dependent in all the other ones, as m = lR/lR m = ef m. In ST gravity masses and couplings are constant in the Jordan frame and space-time dependent in the Einstein one. In general, the space and time scales of variation of the function f = log(lR/lR) depend on the model and can be determined by solving the full set of equations of motion. We will make the assumption that these scales are of cosmological -- or at least astrophysical -- size, and in any case much larger than particle physics interaction times and effective ranges [13, 14]. So, in computing transition amplitudes, decay rates and cross sections, the particle physics parameters adiabatically adjust their relations in eq. (10) to the local values of the functions f , lR and lR, up to corrections of O(λP P /L), λP P being a typical particle physics interaction range and L the typical scale of variation of f . Then, one can compute all the relevant observables in a frame-independent way, following the usual rules of quantum field theory and using the frame-invariant combinations of eq. (10). The results are frame-independent decay rates, cross sections, etc., given by The above quantities are the true observables, that is, in any frame, the dimensionless combinations between the Γ's, σ's, ..., and the appropriate powers of the standard rod length. lR Γ = lR Γ , R σ = l−2 l−2 R σ , ··· . (11) IV. FRAME-INVARIANT FRW COSMOLOGY: BACKGROUND OBSERVABLES We will consider the background FRW metric where τ = x0 is the conformal time, δij is the delta-function, and latin indices run from 1 to 3. We will assume that the function f defining the frame transformation (2) can be expanded as ds2 = −a2(τ )(dτ 2 − δijdxidxj) , (12) f (τ, xi) = ¯f (τ ) + δf (τ, xi) , (13) where δf can be treated as a perturbation of the same order as the metric perturbations. Then, the scale factor in the other frame is given by a(τ ) = e− ¯f(τ )a(τ ) . (14) A. Redshift and temperature One of the basic cosmological observables is the redshift of photon wavelengths. Using the metric (12) one gets the standard result that a photon traveling through the cosmos, which, at time τi had wavelength λ(τi), at a later time τf would have a wavelength λ(τf ) = λ(τi) a(τf ) a(τi) . (15) Looking at the transformation (14) we see that the ratio λ(τf )/λ(τi), which is usually defined as the cosmological redshift, is not a frame-invariant quantity. This should be no surprise, since this ratio is not what is actually measured. Instead, the physical quantity is the dimensionless ratio between the wavelength of the -- emitted or absorbed -- photon and some reference length, measured in the laboratory. Then, the frame-invariant redshift can be defined using a frame-invariant combination such as λ(τ0) ¯lR(τ0) ¯lR(τ ) λ(τ ) = a(τ0) a(τ ) ¯lR(τ ) ¯lR(τ0) , (16) where the bar denotes the spatial average. In order to give an operative definition of redshift, the unit lR has to be specified. In practice, a reference atomic wavelength is chosen, which we will indicate with lR = lat. In principle, different reference wavelengths could have different space-time dependences, thus leading each to a different definition of redshift. However, in ST theories this is not the case, as they all have constant ratios one another. Therefore, in these theories, the redshift can be defined unambiguously as 1 + z(τ ) ≡ a(τ0) a(τ ) ¯lat(τ ) ¯lat(τ0) . (17) 5 The standard relation between the redshift and the scale factor, i.e. 1 + z = a(τ0)/a(τ ) is recovered only in that frame in which the reference wavelength lat is constant in time and space. In ST theories this is the case of the Jordan frame, whereas, in terms of the scale factor of any other frame one has 1 + z = a(τ0)/a(τ ) exp( ¯f (τ ) − ¯f (τ0)), f being the function connecting the frame under consideration with the Jordan one, according to eq. (2). It should be stressed that the reason for the Jordan frame to be singled out from all the possible ones, is a matter of practical utility, namely, the choice of lat in the definition of eq. (17), but has nothing to do with it having a better physical status than the others. In principle, a physicist could decide to measure the wavelength of cosmological photons in Planck units. In this case, his definition of redshift would have the standard relation to the scale factor of the Einstein frame, not the Jordan one. Since the comoving coordinate volume is frame-invariant, the total entropy per comoving coordinate volume is a frame-invariant quantity. In the perfect fluid approximation it is given by the standard expression where r is the comoving radius, ρ and p the energy density and the pressure, and T is the temperature. Recalling the definition of the energy-momentum tensor for matter, S = (ra)3(ρ + p) T , (18) with SM the matter action, we get the transformation laws, T µ ν = e4f T µ ν . Tµν = − 2 √−g δSM δgµν , (19) (20) It should be noted that the above relation is valid as long as the matter action SM does not depend on derivatives of the metric, as is the case in ST theories. In the case of barotropic fluids with background energy ρ = −T 0 0 and pressure p δi j = −T i j , we have From the frame-invariance of the comoving entropy and using eqs. (14) and (21), we get that the dimensionless combination T lR is frame invariant, and that, choosing again lR = lat, the temperature-redshift relation is ρ = e4 ¯f ρ , p = e4 ¯f p . (21) Notice that T ∼ 1/a, in any frame. T (τ )¯lat(τ ) T (τ0)¯lat(τ0) = 1 + z(τ ) . B. Distances (22) The basic quantity entering the definition of the different distance indicators used in cosmology is the comoving distance traveled by a light ray emitted at redshift z, r(z). This is obtained using the (frame-invariant) condition for photon geodesics, i.e. ds2 = 0, that is, using eqs. (12) and (17), r(z) =Z r 0 dr =Z a0 a da a =Z z 0 dz 1 + z a a − ¯lat ¯lat!−1 , (23) where the overdot indicates a derivative with respect to the conformal time τ . The redshift-dependence of the frame-invariant combination a/a − ¯lat/¯lat appearing at the denominator can be computed in any frame. In ST gravity, the ¯lat/¯lat term vanishes in the Jordan Frame and one needs to solve the Friedmann equations for the scale factor in that frame. On the other hand, in the Einstein frame, one needs also the time-dependence of the field ¯f relating the two frames, since ¯lat/¯lat = ¯f in this frame. For practical purposes it may be convenient to work in the Einstein frame, since its equations are simpler. The angular distance of an object of proper diameter D at coordinate r, which emitted light at time τ (and redshift z(τ )) is given by dA ≡ D δ ¯l0 R ¯lR(z) = a0 r(z)(1 + z)−1 , (24) R ≡ ¯lR(τ0) and a0 ≡ a(τ0). In the definition where δ is the observed angular diameter today, and we have defined ¯l0 above, we had to keep track of the possibility that the length of the standard rod lR evolves in time in the frame under consideration. Since a0/¯l0 R, z, and r(z) are all frame invariant, so is also the measure of dA expressed in terms of the present value of the unit length ¯l0 R, that is, the ratio dA/¯l0 R. Analogously, the luminosity distance can be defined in a frame-invariant way as 6 d2 L l0 R 2 ≡ L ¯lR(z)2 4π F l0 R 4 , (25) where L is the luminosity (energy per unit time) of the object at redshift z and F is the energy flux measured today. One can verify that the angular and luminosity distances defined above satisfy the standard relation dL(z) = (1 + z)2dA(z) . (26) Finally, the number counts of objects (galaxies, clusters, . . . ) as a function of redshift measure the frame-invariant observable dN dz = nc(z) r(z)2 a a − ¯lat ¯lat!−1 dz 1 + z dΩ , (27) where nc(z) is the comoving number density. V. FRAME-INVARIANT PERTURBATIONS Now we include first order perturbations of the metric and of the function f , eq. (13). We will work in Newtonian gauge, leaving to the Appendix the extension to the synchronous gauge and to a generic gauge. The line element in a generic frame is given by with Ψ and Φ two scalar functions of space-time. Considering also the fluctuation of lR, lR = ¯lR + δlR, we can write down the frame-invariant line element as ds2 = a2(τ )(cid:2)−(1 + 2 Ψ)dτ 2 + (1 − 2 Φ)δijdxidxj(cid:3) , (28) (29) ds2/l2 R = a2(τ )/¯l2 R(cid:20)−(cid:18)1 + 2 Ψ − 2 δlR lR (cid:19) dτ 2 +(cid:18)1 − 2 Φ − 2 δlR lR (cid:19) δijdxidxj(cid:21) , which still has the form of an invariant line element in a Newtonian gauge, with frame-invariant scale factor and potentials ¯a ≡ a/¯lR , ¯Ψ ≡ Ψ − δlR lR , ¯Φ ≡ Φ + δlR lR . (30) All the physical observables, up to first order, must depend on the above quantities. In the previous section we have seen already how it works at zeroth order. In the following we will extend the program to first-order. A. Frame invariant geodesics It is convenient to work with comoving coordinates, since they are frame-invariant. Then, besides space-time coordinates x0 = τ and xi, we will also consider the conjugate momenta. For a particle of mass m they are given by where ds ≡ √−ds2. As we have seen in sect. III, the lagrangian mass of a particle is not a frame-invariant quantity, m = lR/lR = ef m, independent of whether the particle is a scalar, spinor or a vector. Pµ = m gµν dxν ds , (31) 7 Taking into account the frame dependence of the mass, one can verify that the canonical momenta (31) with low indices are frame-invariant. The geodesic equation for a particle with space-time dependent mass is Using the metric (28) we arrive at the equation for the frame-invariant momentum Pi, P 0 dP µ dτ + Γµ λσ P λ P σ = −m ∂σ m gσµ . dPi dτ − P0 ∂i(Ψ + log m) = 0 , (32) (33) which is manifestly frame-invariant since ∂i(Ψ + log m) = ∂i( ¯Ψ + log ¯m), where ¯m = lR m. In some applications, such as the Boltzmann equation entering the computation of the CMB spectra, see sect. VI, it is convenient to eliminate the perturbations from the definition of the momenta by going to new variables qi and ǫ [17], which can be defined in a frame-invariant way as Pi = (1 − ¯Φ) qi , P0 = −(1 + ¯Ψ) ǫ . Writing qi = q ni, with ninjδij = 1, one can verify the relation ǫ = [q2 + ¯a2 ¯m2]1/2. The geodesic equation in these variables have the standard form which is valid also in the massless case ǫ = q. Photon trajectories are given by the frame-invariant equation ds2 = 0. As a consequence, the expressions for the deflection angles due to weak lensing depend on the frame-invariant combination Ψ + Φ = ¯Ψ + ¯Φ [18]. q = q ∂ ∂τ ¯Φ − ǫ ni ∂ ∂xi ¯Ψ , (35) B. The energy-momentum tensor One can define a frame-invariant distribution function F (xi, Pj, τ ), giving the number of particles in a (frame- invariant) differential volume in phase space, (34) (36) (37) (38) From F one can define the comoving number density, F (xi, Pj , τ ) dx1dx2dx3dP1dP2dP3 = dN . and the energy-momentum tensor, nc(xi, τ ) = gsZ d3P (2π)3 F (xi, Pj, τ ) , Tµν = gsZ d3P (2π)3 (−g)−1/2 PµPν P 0 F (xi, Pj, τ ) , where d3P = dP1dP2dP3 and gs counts the spin degrees of freedom. One can verify that the above definition fulfills the transformation rule of eq. (20). The distribution function for bosons (-) and fermions (+) in equilibrium is the standard one [17], F 0(ǫ) =(cid:16)eǫ/T0 ± 1(cid:17)−1 . (39) Using the variables qi, ǫ defined in eq. (34), the components of the energy momentum tensor are explicitly given, at first order, by ¯T 0 0 = l4 R T 0 ¯T 0 i = l4 R T 0 d3q δ ¯ρ 0 = −gs¯a−4Z i = gs¯a−4Z j = gs¯a−4Z ¯ρ (cid:19) , (2π)3 ǫf = − ¯ρ(cid:18)1 + d3q (2π)3 qnif = ¯ρ(1 + w) vi , d3q ninjf = ¯ρ(cid:20)w + c2 (2π)3 q2 ǫ ¯ρ (cid:21) δi where f (xi, q, nj, τ ) = F (xi, Pj, τ ), vi ≡ dxi/dτ , w is the equation of state, and c2 ¯T i j = l4 R T i δ ¯ρ s j + Σi j , s = ∂ ¯P /∂ ¯ρ. (40) VI. THE BOLTZMANN EQUATION 8 From what we have discussed in the previous session, it is now clear that, using the frame-invariant coordinates xi, Pj, and τ , it is possible to study departures from thermal equilibrium in a frame-invariant way. The tool is, as usual, the Boltzmann equation. The evolution of the phase space density of a particle ψ, Fψ(xi j , τ ) is given by ψ, P ψ ∂F ψ ∂τ + dxi ψ dτ ∂F ψ ∂xi ψ dP ψ j dτ + ∂F ψ ∂P ψ j =(cid:20) dF ψ dτ (cid:21)C . (41) The frame-invariance of the LHS is trivially checked. One can cast it in a more useful form by using dxi and eq. (33). ψ/dτ = P i ψ/P 0 ψ The collisional term for a generic process ψ + a + b + ··· ↔ i + j + ··· reads 0 Z dΠadΠb ··· dΠidΠj ··· 1 2P ψ (xi ψ, P ψ j , τ ) = (cid:20) dFψ dτ (cid:21)C ×(2π)4δ4(P ψ + P a + P b ··· − P i − P j ···) ×(cid:2)M2 −M2 ψ+a+b+···→i+j+···FψFaFb ··· (1 ± Fi)(1 ± Fj)··· i+j+···→ψ+a+b+···FiFj ··· (1 ± Fψ)(1 ± Fa)(1 ± Fb)···(cid:3) , where the ′′+′′ applies to bosons and the ′′−′′ to fermions, and dΠ is the frame-invariant quantity dΠ ≡ l2 R = l2 R d4P (2π)3 (−g)−1/2 δ(P 2 + m2)Θ(P 0) d3P (2π)3 (−g)−1/2 ¯l2 R a−2 (2π)3 2P 0 = d3q 2ǫ , (42) (43) with d4P = dP0 d3P . The delta-function in eq. (42) depends on momenta with low indices. A. Freeze out As a first example, we consider the case of a heavy particle decaying into two lighter ones, which are assumed to equilibrate rapidly. Following the standard procedure (see for instance ref. [19]) the (conformal) time dependence of the comoving number density is given by nψ c = gsZ d3P (2π)3 M2 =Z d3P Ψ 2P ψ 0 dτ (cid:21)C (2π)3 (cid:20) dFψ dΠadΠb(2π)4δ(4)(P ψ − P a − P b)(Fψ − F 0 a F 0 b ) , where we have approximated 1 ± F ≃ 1.Using energy conservation we can write, as usual, a F 0 F 0 b = F 0 ψ , and then nψ c = −(nψ c − nψ c 0 )Z dΠadΠb(2π)4δ(4)(P ψ − P a − P b) M2 2mψa , (44) (45) (46) where for the non-relativistic particle Ψ we have used P ψ 0 ≃ −mψa. Recognizing the integral as the decay rate per unit conformal time, Γ a, where Γ is the decay probability per unit physical time, and turning to the variable x = mψ/T , we get the frame-invariant equation d nψ c d log x = − Γ a H(1 + m′ ψ/mψ) (nψ c − nψ c 0 ) , (47) where primes denote derivatives with respect to log a and we have used the relation T /T = −H. From the above equation one can see that the usual rule of thumb for a particle interaction to be efficient in the expanding Universe, that is Γ a >∼ H, now generalizes to the frame-invariant relation Γ a >∼ H(1 + m′ ψ/mψ) . (48) 9 We stress again that the frame independence of the above equation is a consequence of the frame-independence of the product Γa, and then, ultimately, of that of the matrix element M2. Analogously, in the case of a 2 ↔ 2 scattering process one gets the result 0!2 nψ Γ a H(1 + m′ d nψ c d log x = − c nψ c 1 nψ c 0 ψ/mψ)  , − 1  (49) where Γ = nψ c 0 a−3hσvi, with hσvi the thermally averaged cross section. From these examples one can appreciate the utility of the frame-invariant Boltzmann equations. In practice, it turns out that rates, cross sections, and all the particle physics related quantities are more conveniently computed in a frame different from that in which the Einstein equations are simpler. For instance, in scalar-tensor theories, particle physics is conveniently computed in the Jordan frame, whereas gravitational equations are simpler in the Einstein frame. Since frame-invariant combinations -- such as Γa -- appear in the equations above, one can first compute rates and cross sections in the more convenient frame and then translate them in the other one using the relations of eq. (11). Using the 0-0 component of the energy-momentum tensor, eq. (40) one can define (space and direction-dependent) temperature fluctuations for a gas of photons (ǫ = q, gs = 2) as B. Sachs-Wolfe effect From the collisionless Boltzmann equation for the function f , Θ(xi, nj, τ ) ≡ ∆T T (xi, nj, τ ) = 1 4π2 ¯ρ¯a4 Z dqq3f − 1 . ∂f ∂τ + xi ∂f ∂xi + q ∂f ∂q + nj ∂f ∂nj = 0 , using the geodesic equation (35) and the relation ni = qi/q = xi(1 − ¯Φ − ¯Ψ)ǫ/q, one gets (Θ + ¯Ψ) = ¯Ψ + ¯Φ , d dτ (50) (51) (52) where use has been made of the fact that potentials and δlR do not depend on the angle explicitly, and ( ) ≡ ∂ /∂τ . If the potential are static, the quantity Θ + ¯Ψ is conserved, which is the frame-invariant expression for the Sachs-Wolfe effect [20]. C. Phase space evolution for CMB photons The evolution of the CMB photon distribution function is described by the Boltzmann equations discussed, for instance, in [17, 21, 22]. To reduce the number of variables, one integrates out the q dependence and expands the angular dependence in Legendre polynomials, Pl. Going to Fourier space, one defines F (~k, n, τ ) ≡ R f 1(~k, ~q, τ )q3dq R f 0(q)q3dq ≡ (−i)l(2l + 1)Fl(~k, τ )Pl(k · n) , Xl=0 ≡ ∞ G(~k, n, τ ) ≡ R Q1(~k, ~q, τ )q3dq R Q0(q)q3dq ≡ (−i)l(2l + 1)Gl(~k, τ )Pl(k · n) , Xl=0 ≡ ∞ 10 (53) (54) where ~k = kk is the wavevector and n the direction of the photons 3-momentum ~q. f 0 is the zeroth order (equilibrium) distribution function and f 1 the first order deviation from it, while Q0 and Q1 are the zeroth and first order Stokes parameter, respectively. The Boltzmann equations for F and G take the form ∂F ∂τ ∂τ (cid:19)C + ikµF − 4( ¯Φ − ikµ ¯Ψ) =(cid:18) ∂F ∂τ (cid:19)C + ikµG =(cid:18) ∂G ∂G ∂τ , , with µ ≡ k · n. Again, the LHS are manifestly frame-invariant. The collisional terms are given by [17] (cid:18) ∂F ∂τ (cid:19)C = a−2ne ∂τ (cid:19)C (cid:18) ∂G c σT (cid:20)−F + F0 + 4n · ~ve − = a−2ne c σT (cid:20)−G + 1 2 1 2 (F2 + G0 + G2) P2(cid:21) , (F2 + G0 + G2) (1 − P2)(cid:21) , (55) (56) c are respectively the proper velocity and comoving density of the electrons, and σT the Thomson c σT = ¯a−2ne c ¯σT is frame-invariant, and so is the proper velocity, the frame-invariance of the collisional where ~ve and ne cross section. Since a−2ne terms is also manifest. VII. EQUATIONS OF MOTION FOR SCALAR-TENSOR THEORIES It is convenient to define the frame-invariant metric hµν ≡ l−2 P l gµν , (57) where the unit length lP l will be later identified with the Planck length. If gµν is a FRW metric in Newtonian gauge, so is hµν, with scale factor ¯a and potentials ¯Ψ, ¯Φ as defined in eq. (30) with lR = lP l. Notice that hµν and ¯a2 have dimension of (mass)2. Scalar-tensor theories can be defined in terms of frame-independent quantities by the action S = SG[hµν , ϕ] + SM [hµνe−2b[ϕ], ¯φ , ¯ψ , . . . ; ¯λn] , (58) where the frame-independent fields ¯φ , ¯ψ , . . ., and coupling constants ¯λn's, appearing in the matter action SM are given by the combinations in eq. (10) with lR = lP l. The gravity action is given by SG = κZ d4x√ −h(cid:2)R(h) − 2 hµν∂µϕ ∂ν ϕ − 4 ¯V (ϕ)(cid:3) . (59) The only feature differentiating the action in eq. (58) from that of standard GR is the function b[ϕ(x)]. In the b = 0 limit, the scalar-tensor theory reduces to GR with an extra scalar field, ϕ, which in this limit can be seen as an extra matter component minimally coupled to gravity. In this case, a constant lP l can be univocally taken as the most convenient choice to measure all the dimensional quantities in the theory. In this units, both the Planck mass and particle masses, as well as atomic wavelengths, are constant. On the other hand, when b 6= 0, the scalar field ϕ(x) is non-minimally coupled to gravity and one has a genuine scalar-tensor theory. In the literature, these theories are usually discussed in two frames, the Einstein and the Jordan ones. In our language, choosing a frame corresponds to fixing the function lP l(x) appropriately. 11 The first possible choice is to take a constant lP l, which corresponds to the Einstein frame. The gravity action takes the usual Einstein-Hilbert form SG = κ l−2 P l Z d4x√−g [R(g) − 2 gµν∂µϕ ∂ν ϕ − 4V (ϕ)] , (60) ¯V . The combination in front of the integral fixes the Einstein-frame Planck mass, κ l−2 with V = l−2 ∗ /2 = P l (16πG∗)−2. In other words, in this frame, dimensional units are set by the Planck scale. The matter action is obtained from the one of quantum field theory by substituting the Minkowsky metric ηµν with gµνe−2b. Since in this frame the matter energy-momentum tensor is not conserved (see eq. (66)), particle physics quantities, like masses and wavelengths are not constant. P l = M 2 The other choice corresponds to the Jordan frame, which is obtained by making the Planck length space-time dependent such as to reabsorb b[ϕ(x)] in SM . This is accomplished if one choses lP l = lP l e−b, where lP l is the previously defined Planck length in the Einstein frame. With this choice the matter action takes the standard form of quantum field theory (with ηµν → gµν ), whereas the gravity action is SG = where V = l−2 P l ¯V and M 2 ∗ 2 Z d4xp−g e2bhR(g) − 2 gµν∂µϕ ∂νϕ (1 − 3 α2) − 4 V (ϕ)i , db dϕ . α ≡ (61) (62) Notice that, in this frame, the role of the Planck mass is played by the space-time dependent quantity M∗eb. Since b[ϕ(x)] disappears from the matter action, the energy-momentum tensor is now covariantly conserved. Of course, any other choice for lP l is possible in principle and leads to the same physical consequences. However, for practical purposes, only the Einstein and Jordan frames are employed. The discussion of the previous sections shows how one can exploit the good aspects of the two. One can compute cross sections, decay rates, etc., in the Jordan frame, where masses and couplings are constant and the usual rules of quantum field theory apply straightforwardly. Then, one constructs frame-invariant combinations out of these, such as Γa, and insert them into the frame-independent Boltzmann equations like eqs. (47, 49). On the other hand, the gravity part, like the combination H(1 + m′/m) can be computed in the Einstein frame. Equivalently, both the particle physics part and the gravity part can be computed frame-invariantly from the beginning, using the particle physics parameters defined in eq. (10) and solving the equations of motion obtained from the action in eq. (58), that we are going to write down explicitly up to first order. Before doing that, we give the expression of the redshift (17) in terms of the frame-invariant scale factor ¯a = a/lP l, that is where ¯b(τ ) ≡ b[ ¯ϕ(τ )]. 1 + z = ¯a(τ0) ¯a(τ ) ¯b(τ )−¯b(τ0) , e A. Background equations The background equations for the scale factor ¯a are The energy-momentum tensor ¯T µ 2 3(cid:18) 1 2 ¯a ¯a ϕ2 + ¯a2 ¯V(cid:19) = ¯a(cid:19)2 (cid:18) ¯a − 3(cid:0) ϕ2 − 4¯a2 ¯V(cid:1) = − ν is not conserved, ν; µ = −α ϕν ¯T µ ¯T µ µ , 12κ 1 1 + 1 6κ ¯ρ¯a2 , ¯ρ¯a2(1 − 3w) . which, at zeroth-order implies ¯ρ + 3 ¯ρ (1 + w) ¯a/¯a = −α ϕ ¯ρ (1 − 3w) . (63) (64) (65) (66) (67) One can verify that, with the Jordan frame choice, i.e. lP l ∼ e−b, the covariant conservation of the energy-momentum tensor is recovered. The equation of motion for the field ϕ is 12 which, at zeroth-order, gives hµν ¯Dνϕµ − δ ¯V δϕ = α 4κ ¯T µ µ , ϕ + 2 ¯a ¯a ϕ + ¯a2 ∂ ¯V ∂ϕ = α 4k ¯a2 ¯ρ (1 − 3w) . B. First-order equations Here we give the set of first order equations: δϕ = − 1 4κ δ ¯ρ¯a2 , k2 ¯Φ + 3 k2(cid:18) ¯Φ + ¯Φ + ¯a ¯a ¯ρ¯a2(1 + w)θ , ¯a ¯a ¯Ψ(cid:19) − ( ϕδ ϕ − ϕ2 ¯Ψ) + 2¯a2 δ ¯V δϕ ¯a ¯a ¯a 1 4κ ¯a(cid:18) ¯Φ + ¯Ψ(cid:19) − ϕk2δϕ = ¯a(cid:19)2# ¯Ψ − ¯a −(cid:18) ¯a + ¯Ψ ϕ2 − ϕδ ϕ + ( ¯Ψ + 2 ¯Φ) +"2 ¯a 1 k2( ¯Ψ − ¯Φ) + 3 δ ¯V 1 δϕ 4κ δϕ = c2 sδ ¯ρ¯a2 , k2( ¯Φ − ¯Ψ) = 3 4κ ¯ρ¯a2(1 + w)σ , where θ ≡ ikjvj and ¯ρ(1 + w)σ ≡ −( ki kj − δij /3)Σi j, with ki = ki/k. At first order, the continuity equation for the matter energy-momentum tensor, eq. (66), yields ∂α ∂ϕ δϕ(cid:19) − ∂w ∂ ¯ρ ¯δ = −(1 + w)(θ − 3 ¯Φ) − (1 − 3w)(cid:18)αδ ϕ + ϕ ¯ρ¯δ(cid:18) ¯a − 3 θ = −(1 − 3w)(cid:18) ¯a + k2( ¯ψ − σ) − ¯a − α ϕ(cid:19) , ¯a − α ϕ(cid:19) θ − (1 − 3w) (1 + w) k2αδϕ , 1 + w w ∂ ¯ρ ¯ρ(cid:17) θ + (cid:16)w + ∂w 1 + w k2¯δ where ¯δ ≡ δ ¯ρ/ ¯ρ. The equation of motion for the scalar field fluctuation δϕ, from eq. (68) is δ ϕ + 2 ¯a ¯a δ ϕ − ∇2δϕ − ¯ψ ϕ − 3 ϕ ¯Φ + 2 ¯ψ¯a2 ∂ ¯V ∂ϕ − ∂ϕ2 δϕ = 4k (cid:20)¯ρ(1 − 3w) δϕ +(cid:18)1 − 3w − 3 ¯ρ¯a2(1 − 3w) ¯ψ + ¯ρ(cid:19) αδ ¯ρ(cid:21) . ∂w ∂ ¯ρ ∂α ∂ϕ α 2k ¯a2 + ¯a2 ∂2 ¯V (68) (69) (70) (71) (72) (73) (74) (75) (76) A particular ST theory is identified by two functions: b(ϕ) and the effective potential ¯V (ϕ), see eqs. (58) and (59). Therefore, the physical deviations from GR should be parameterized in terms of these two functions alone, VIII. CONCLUSION 13 irrespectively of the frame one choses to solve the equations of motion [2, 3]. Actually, as we have shown, fixing a frame is not necessary, provided one carefully expresses all the observables in terms of frame-invariant quantities. From a practical point of view, the formulation of ST gravity presented in this paper allows a straightforward modification of the available codes based on Boltzmann equations for the study of Nucleosynthesis, CMB, or the calculation of the Dark Matter relic abundances in the context of GR. It is enough to redefine the redshift as in eq. (63) and add the scalar field ϕ to the GR equations of motion, as in sects. VII A and VII B. Then, the code will work in the standard way, the only difference being given by the two extra inputs b and ¯V . The implementation of this procedure to the publicly available CMBFAST [23] code is under way. We will show how the results obtained in the text in the Newtonian gauge can be extended to the synchronous gauge and to a generic gauge. IX. APPENDIX The synchronous gauge is defined by A. The synchronous gauge ds2 = a2(cid:2)−dτ 2 + (δij + hij) dxidxj(cid:3) . After a frame transformation ds2 → ds2 = e−2f ds2 the metric of eq. (77) transforms in ds2 = a2e−2fB(cid:2)−(1 − 2δf )dτ 2 + (δij + hij − 2δf δij) dxidxj(cid:3) , (77) (78) Unlike for the Newtonian gauge, a frame transformation doesn't preserve the synchronous gauge. However it is possible to give a frame-independent description of all the physical phenomena also when the metric perturbations are described in the basis (hij, δf ). In fact, as we will see, the geodesic motion depends only from the frame-independent combination ¯hij = hij − 2(δlR/lR)δij . To show this in a simple way let's define a synchronous gauge as in the following In this way at least the spatial components of the metric are frame-independent. dh2 = ¯a2(cid:2)−dτ 2 +(cid:0)δij + ¯hij(cid:1) dxidxj(cid:3) . As for the Newtonian gauge we now relate the 4-momentum Pµ to the frame-independent variables qi and ǫ P0 = −ǫ , Pi = (cid:18)δij + 1 2 ¯hij(cid:19) qj . With this definitions the relation ǫ =pq2 + ¯m¯a2 is preserved. Using eqs. (80) and the metric of eq. (79) the geodesics equation at first order yields q = − 1 2 qninj ¯hij . Also if the metric of eq. (79) is not frame-invariant, eq. (81) is not affected by a frame transformation. Let's consider now the metric B. The case of a generic gauge (79) (80) (81) ds2 = a2{−(1 + 2ψ)dτ 2 + 2∂iBdτ dxi + [(1 − 2Φ)δij + DijE] dxidxj} (82) where Dij =(cid:0)∂i∂j − 1 After a frame transformation ds2 → ds2 = e−2f ds2 eq. (82) transforms in 3 δij∇2(cid:1). ds2 = a2e−2fB{−(1 + 2ψ − 2δf )dτ 2 + 2∂iBdτ dxi + + [(1 − 2Φ − 2δf )δij + DijE] dxidxj} 14 (83) The transformation properties of the metric in eq. (82) suggest to define the following frame-invariant line element where the frame-invariant quantities ¯a, ¯ψ and ¯Φ are given in eq. (30). We also wrote ¯B = B and ¯E = E to underline dh2 = ¯a2{−(1 + 2 ¯ψ)dτ 2 + 2∂i ¯Bdτ dxi +(cid:2)(1 − 2 ¯Φ)δij + Dij ¯E(cid:3) dxidxj} , (84) that such a quantities do not transform under frame transformations. The 4-momentum is now related to the frame invariant variables qi and ǫ by the following relations P0 = −(cid:2)qi∂i ¯B + ǫ(1 + ¯ψ)(cid:3) , Dij ¯E(cid:21) qj , Pi = (cid:20)(1 − ¯Φ)δij + 1 2 with ǫ =pq2 + ¯m¯a2. Using now eqs. (85) and the metric of eq. (84) at first order the geodesic equation yields q = q ¯Φ − ǫni∂i ¯ψ + 2ǫni(cid:18) ¯a ¯a ∂i ¯B + ∂i ¯B(cid:19) − qninj(cid:18)∂i∂j ¯B + 1 2 Dij ¯E(cid:19) (85) (86) If in eq. (86) we impose ¯B = ¯E = 0 we recover eq. (33). Choosing instead ¯B = ¯ψ = 0 and −2 ¯Φδij + Dij ¯E = ¯hij eq. (86) reduces to eq. (81). R. Catena acknowledges a Research Grant funded by the VIPAC Institute. Acknowledgments [1] P. Jordan, Schwerkaft und Weltall (Vieweg, Braunschweig, 1955); Nature (London), 164, 637 (1956); M. Fierz, Helv. Phys. Acta 29, 128 (1956); C. Brans and R.H. Dicke, Phys. Rev. 124, 925 (1961). [2] C.Will, Theory and experiments in gravitational physics, Cambridge University ress, 1990), pp. 313. [3] T. Damour, gr-qc/9606079, lectures given at Les Houches 1992, SUSY95 and Corf´u 1995. [4] B. Bertotti, L. Iess and P. Tortora, Nature 425, 374 (2003). [5] K. Choi, hep-ph/9912218; M. Pietroni, Phys. Rev. D 72, 043535 (2005). [6] T. Damour and K. Nordtvedt, Phys. Rev. Lett. 70, 2217 (1993); Phys. Rev. D 48, 3436 (1993);T. Damour and A. M. Polyakov, Nucl. Phys. B 423, 532 (1994). [7] N. Bartolo and M. Pietroni, Phys. Rev. D 61, 023518 (2000). [8] R. Catena, N. Fornengo, A. Masiero, M. Pietroni and F. Rosati, Phys. Rev. D 70 (2004) 063519; R. Catena, M. Pietroni and L. Scarabello, Phys. Rev. D 70, 103526 (2004). [9] G. Esposito-Farese and D. Polarski, Phys. Rev. D 63, 063504 (2001). [10] A. Coc, K. A. Olive, J. P. Uzan and E. Vangioni, [arXiv:astro-ph/0601299]. [11] C. Schimd, J. P. Uzan and A. Riazuelo, Phys. Rev. D 71, 083512 (2005). [12] J. Martin, C. Schimd and J. P. Uzan, Phys. Rev. Lett. 96, 061303 (2006). [13] F. Perrotta, S. Matarrese, M. Pietroni and C. Schimd, Phys. Rev. D 69, 084004 (2004). [14] S. Matarrese, M. Pietroni and C. Schimd, JCAP 0308, 005 (2003). [15] R. H. Dicke, Phys. Rev. 125, 2163 (1962). [16] C. Armendariz-Picon, Phys. Rev. D 66, 064008 (2002); E. E. Flanagan, Class. Quant. Grav. 21, 3817 (2004). [17] C. P. Ma and E. Bertschinger, Astrophys. J. 455, 7 (1995). [18] N. Kaiser Astrophys.J. 498 (1998) 26; V. Acquaviva, C. Baccigalupi and F. Perrotta, Phys. Rev. D 70, 023515 (2004). [19] E. W. Kolb and M. S. Turner, "The Early Universe", 1990, Redwood City, USA, Addison-Wesley, 547 pp. [20] R. K. Sachs and A. M. Wolfe, Astrophys. J. 147, 73 (1967). [21] J. R. Bond and G. Efstathiou, Astrophys. J. 285, L45 (1984); Mon. Not. Roy. Astron. Soc. 226, 655 (1987). [22] A. Kosowsky, Annals Phys. 246, 49 (1996). [23] U. Seljak and M. Zaldarriaga, Astrophys. J. 469, 437 (1996). [24] R. Catena, M. Pietroni and L. Scarabello, J. Phys. A 40, 6883 (2007). 15
astro-ph/0208331
1
0208
2002-08-17T19:35:20
NGC 3576 IRS 1 in the Mid Infrared
[ "astro-ph" ]
We present the results of high-resolution mid-infrared observations of the source NGC 3576 IRS 1. Near diffraction-limited images were taken at the Gemini South Observatory through OSCIR's filters N, 7.9, 9.8, 12.5 and IHW18. The source IRS 1 was resolved into 3 sources for the first time at mid-infrared wavelengths. For each source we constructed the SED from 1.25 to 18 microns, as well the color temperature and the spatial distribution of the dust in the region. The optical depth of the silicate absorption feature at 9.8 microns is presented also.
astro-ph
astro-ph
**TITLE** ASP Conference Series, Vol. **VOLUME***, **YEAR OF PUBLICATION** **NAMES OF EDITORS** NGC 3576 IRS 1 in the Mid Infrared C´assio Leandro D. R. Barbosa IAG-USP, R.do Matao 1226, 05508-900, Sao Paulo, Brazil [email protected] Augusto Damineli IAG-USP, R.do Matao 1226, 05508-900, Sao Paulo, Brazil [email protected] Robert D. Blum Cerro Tololo Interamerican Observatory, Casilla 603, La Serena, Chile [email protected] Peter S. Conti JILA, University of Colorado, Campus Box 440, Boulder, CO, 80309 [email protected] Abstract. We present the results of high-resolution mid-infrared ob- servations of the source NGC3576 IRS 1. Near diffraction-limited images were taken at the Gemini South Observatory1 through OSCIR's2 filters N (10.8 µm), 7.9, 9.8, 12.5 and IHW18 (18.2 µm). The source IRS1 was resolved into 3 sources for the first time at mid-infrared wavelengths. For each source we constructed the SED from 1.25 to 18 µm, as well the color temperature and the spatial distribution of the dust in the region. The optical depth of the silicate absorption feature at 9.8 µm is presented also. 1. Introduction The formation mechanism of massive stars is essentially unknown. This is mostly due to observational difficulties in finding and establishing an evolutionary se- 1Based on observations obtained at the Gemini Observatory, which is operated by the Associ- ation 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 Particle Physics and Astronomy Research Council (United Kingdom), the National Re- search Council (Canada), CONICYT (Chile), the Australian Research Council (Australia), CNPq (Brazil) and CONICET (Argentina). 2This paper is based on observations obtained with the mid-infrared camera OSCIR, developed by the University of Florida with support from the National Aeronautics and Space Admin- istration, and operated jointly by Gemini and the University of Florida Infrared Astrophysics Group. 1 2 C´assio Barbosa et al. quence for young stellar objects (YSOs). It is believed that during the accretion phase, YSOs remain heavily enshrouded in dusty cocoons, behind hundreds of magnitudes of extinction at visual wavelengths (even at near-infrared wave- lengths for the earliest phases). A very interesting group of YSOs has been identified in the Galactic gi- ant HII regions (GHII) M17 (Hanson, Horwarth & Conti 1997), W43 (Blum, Damineli & Conti 1999), W42 (Blum, Conti & Damineli 2000), W31 (Blum, Damineli & Conti 2001) and NGC 3576 (Figueredo et al. 2002). These are lu- minous YSOs with excess emission in the K band and have color indexes H-K>2. J, H and K spectra of these objects typically show a featureless continuum. In some cases the CO 2.3 µm bandhead is seen in emission or absorption, and in others FeII and/or H2 are seen in emission. Figueredo et al. (2002) identified two massive YSO candidates in the posi- tion of IRS 1 in NGC 3576. This source was discovered by Lacy, Beck & Geballe (1982) and it was recently observed at 10 µm by Walsh et al. (2001), but none of their images have enough spatial resolution to resolve the source. 2. Observations Observations were obtained at the Gemini South Observatory 8-m telescope on December 4th, 2001. OSCIR is based on a Rockwell 128×128 pixel Si:As BIB detector, with a 0.089 arcsec/pixel plate scale at Gemini. The total field of view of the camera is 11×11 arcsec and the spatial resolution (FWHM of the PSF star) is ∼0.5 arcsec. The images were taken through the wide N-band filter (10.8µm) and IHW18 (18.2 µm) and the 7.9, 9.8 and 12.5 µm narrow filters. Flux calibration was performed by taking the flux densities of the mid-infrared standard star α CMa observed during the night as part of the baseline calibration program. The uncertainty of these procedures is estimated to be ∼10%, which is good enough for our purposes. All images presented have on-source exposure time of 46 seconds, except the N-band image which has 40 seconds. Background and sky subtraction was done via the standard chop and nod technique. 3. Results The field of NGC 3576 IRS 1 is shown in the Figure 1, for the N and IHW18 bands. IRS 1 has been resolved into 3 sources embedded in extended emis- sion. All images were smoothed by convolving each image by the corresponding normalized PSF star fitted from the standard star using a Fourier transform. Figure 2 shows the spectral energy distribution (SED) of each source. The JHK fluxes are from Figueredo et al. (2002), the L-band flux is from Moneti (1992). The NIR-to-MIR SEDs are very similar to the results found for a group of less massive AeBe stars by Hillenbrand et al. (1992). The stars of this group (called group II) have SEDs with infrared excess and are supposed to be young stars with intermediate masses (M∼< 10 M⊙, average spectral type A5) still accreting. The infrared excess is attributed to a dense dusty accretion disk surrounding the young stars. NGC 3576 in the Mid Infrared 3 Left:The N-band image of IRS-1. The contours are at 0.01 Figure 1. 0.025, 0.05, 0.1, 0.2, ..., 0.5 Jy/ ✷". Right: The 18µm image of IRS-1. The contours are at 0.044, 0.088, 0.13, 0.17, 0.2, 0.27, 0.35, 0.4 Jy/ ✷". North is up, East is left for both images. Each image is 11×11 arcsec. NIR/MIR NGC 3576 IRS-1 SED J H K L N #48 #50 #60 -7.5 -8 -8.5 -9 -9.5 -10 -10.5 -11 14.5 14.25 14 13.75 logν (Hz) 13.5 13.25 ) y J ( x u F g o l l 2 1 0 -1 -2 NGC 3576 IRS-1 SED #50 #60 #48 f-f 8 10 12 Wavelength (µm) 14 16 18 20 ) 2 m c / s / g r e ( ν F ν g o l Left:NIR-to-MIR SED for each source identified in the IRS Figure 2. 1 position. Right: The SED constructed from the fluxes measured at mid infrared.The dashed line is the projected free-free emission of the gas from De Pree, Nysewander & Goss (1999), assuming Sν ∝ ν −1. 4 C´assio Barbosa et al. Figure 3. Left: The map of the silicate absorption at IRS-1 position. The darker is the region, the strongest is the absorption. The contour at the source #60 position indicates where the absorption turns into emission. Right: The dust color temperature map of IRS-1. The con- tour levels are represent temperatures of 160, 180, 200, ..., 280 K. Both images are 5.8×5.8 arcsec. Figure 3 shows the spatial distribution of the dust, as well as, its color temperature map. The map at the left was obtained dividing the calibrated image taken at 9.8 µm by the 7.9 µm image. From this map, we have calculated the optical depth of the silicate absorption feature and we found τ9.8=3.7. The right panel is the dust color temperature map which was calculated from the ratio of the 7.9/18 µm images. From this map we can see the dust tempera- ture associated with each source: T∼280K (#50), T∼210K (#48) and T<160K (#60). 4. Summary We presented high-resolution mid-infrared images of NGC 3576 IRS 1, that has been resolved into 3 sources for the first time at mid-infrared wavelengths. The SEDs of each source were constructed from 1.25 to 18 µm with data taken from literature. Each SED shows increasing fluxes toward to longer wavelengths. Hil- lenbrand et al. (1992) found similar SEDs for a group of less massive young stars in process of accreting mass via an accretion disk embedded in a dusty cocoon. This similarity suggests the same interpretation for our results. Finally we presented maps of dust distribution, temperature and optical depth as well. We thank Chris De Pree for kindly providing the FITS file of the 3.4 cm map of NGC 3576. CLB and AD acknowledge the financial support from PROAP, PRONEX and NGC 3576 in the Mid Infrared 5 FAPESP. PC appreciates continuing support from the National Science Foundation. References Blum, R. D., Damineli, A. & Conti, P. S. 1999, AJ, 117, 1392 Blum, R. D., Conti, P. S. & Damineli, A. 2000, AJ, 119, 1860 Blum, R. D., Damineli, A. & Conti, P. S. 2001, AJ, 121, 3149 De Pree, C. G., Nysewander, Melissa C. & Goss, W. N. 1999, AJ, 117, 2902 Figueredo E., Blum, R. D., Damineli, A. & Conti, P. S. 2002, AJ, accepted Hanson, M., M., Horwarth, I. D. & Conti, P. S. 1997, ApJ, 489, 698 Hillenbrand, L. A., Strom, S. E., Vrba, F. J., Keene, J. 1992 ApJ, 397, 613 Lacy, J., Beck, S. & Geballe, T. 1982, ApJ, 255, 510 Moneti, A. 1992, A&A, 259, 627 Walsh, A. J., Bertoldi, F., Burton, M. G. & Nikola, T. 2001, MNRAS, 326, 36
0809.3265
1
0809
2008-09-18T22:17:46
HI and CO in the circumstellar environment of the oxygen-rich AGB star RX Lep
[ "astro-ph" ]
Circumstellar shells around AGB stars are built over long periods of time that may reach several million years. They may therefore be extended over large sizes (~1 pc, possibly more), and different complementary tracers are needed to describe their global properties. In the present work, we combined 21-cm HI and CO rotational line data obtained on an oxygen-rich semi-regular variable, RX Lep, to describe the global properties of its circumstellar environment. With the SEST, we detected the CO(2-1) rotational line from RX Lep. The line profile is parabolic and implies an expansion velocity of ~4.2 km/s and a mass-loss rate ~1.7 10^-7 Msun/yr (d = 137 pc). The HI line at 21 cm was detected with the Nancay Radiotelescope on the star position and at several offset positions. The linear shell size is relatively small, ~0.1 pc, but we detect a trail extending southward to ~0.5 pc. The line profiles are approximately Gaussian with an FWHM ~3.8 km/s and interpreted with a model developed for the detached shell around the carbon-rich AGB star Y CVn. Our HI spectra are well-reproduced by assuming a constant outflow (Mloss = 1.65 10^-7 Msun/yr) of ~4 10^4 years duration, which has been slowed down by the external medium. The spatial offset of the HI source is consistent with the northward direction of the proper motion, lending support to the presence of a trail resulting from the motion of the source through the ISM, as already suggested for Mira, RS Cnc, and other sources detected in HI. The source was also observed in SiO (3 mm) and OH (18 cm), but not detected. The properties of the external parts of circumstellar shells around AGB stars should be dominated by the interaction between stellar outflows and external matter for oxygen-rich, as well as for carbon-rich, sources, and the 21-cm HI line provides a very useful tracer of these regions.
astro-ph
astro-ph
Astronomy&Astrophysicsmanuscript no. 0089 November 1, 2018 c(cid:13) ESO 2018 H i and CO in the circumstellar environment of the oxygen-rich AGB star RX Lep Y. Libert1, T. Le Bertre1, E. G´erard2, and J.M. Winters3 8 0 0 2 p e S 8 1 ] h p - o r t s a [ 1 v 5 6 2 3 . 9 0 8 0 : v i X r a 1 LERMA, UMR 8112, Observatoire de Paris, 61 Av. de l'Observatoire, 75014 Paris, France e-mail: [email protected] 2 GEPI, UMR 8111, Observatoire de Paris, 5 Place J. Janssen, 92195 Meudon Cedex, France 3 IRAM, 300 rue de la Piscine, 38406 St. Martin d'H`eres, France Received: 29/04/2008; accepted: 25/08/2008 ABSTRACT Context.Circumstellar shells around AGB stars are built over long periods of time that may reach several million years. They may therefore be extended over large sizes (∼ 1 pc, possibly more), and different complementary tracers are needed to describe their global properties. Aims.We set up a program to explore the properties of matter in the external parts of circumstel- lar shells around AGB stars and to relate them to those of the central sources (inner shells and stellar atmospheres). Methods.In the present work, we combined 21-cm H i and CO rotational line data obtained on an oxygen-rich semi-regular variable, RX Lep, to describe the global properties of its circumstellar environment. Results. With the SEST, we detected the CO(2-1) rotational line from RX Lep. The line profile is parabolic and implies an expansion velocity of ∼ 4.2 km s−1 and a mass-loss rate ∼ 1.7 × 10−7 M⊙ yr−1 (d = 137 pc). The H i line at 21 cm was detected with the Nanc¸ay Radiotelescope on the star position and at several offset positions. The linear shell size is rel- atively small, ∼ 0.1 pc, but we detect a trail extending southward to ∼ 0.5 pc. The line profiles are approximately Gaussian with an FWHM ∼ 3.8 km s−1 and interpreted with a model developed for the detached shell around the carbon-rich AGB star Y CVn. Our H i spectra are well-reproduced by assuming a constant outflow ( M = 1.65 × 10−7 M⊙ yr−1) of ∼ 4 × 104 years duration, which has been slowed down by the external medium. The spatial offset of the H i source is consistent with the northward direction of the proper motion measured by Hipparcos, lending support to the presence of a trail resulting from the motion of the source through the ISM, as already suggested for Mira, RS Cnc, and other sources detected in H i . The source was also observed in SiO (3 mm) and OH (18 cm), but not detected. Conclusions. A detached shell, similar to the one around Y CVn, was discovered in H i around RX Lep. We also found evidence of an extension in the direction opposite to the star proper motion. The properties of the external parts of circumstellar shells around AGB stars should be dominated by the interaction between stellar outflows and external matter for oxygen-rich, as well as for carbon-rich, sources, and the 21-cm H i line provides a very useful tracer of these regions. Key words. stars: individual: RX Lep -- stars: mass-loss -- stars: AGB and post-AGB -- stars: winds, outflows -- radio lines: stars -- circumstellar matter 2 Libert et al.: H i & CO in RX Lep. 1. Introduction Evolved stars on the asymptotic giant branch (AGB) are often surrounded by circumstellar shells. The material in these shells is flowing outwards with velocities from a few km s−1 up to 40 km s−1 (Nyman et al. 1992). The observed mass-loss rates range from ∼ 10−8 to a few 10−4 M⊙ yr−1 (e.g. Knapp & Morris 1985; Olofsson et al. 2002), the lower limit being probably set by detectability. In this phase of the stellar life, the evolution is dominated by mass loss rather than nuclear processes (Olofsson 1999). The history of mass loss over the full AGB is complex and the details of this process are currently not well known (e.g., Lafon & Berruyer 1991; Habing 1996; Leao et al. 2006). A general picture, however, has arisen from both theoretical and observational findings that - on average - the mass-loss rate increases towards the end of the AGB phase, leading in some cases to the formation of a planetary nebula (e.g., Renzini 1981; Hrivnak & Bieging 2005). The validity of this simple picture may depend on the parameters of the star, e.g., on its initial mass. Schroder et al. (1999) combine mass-loss rates derived from consistent wind models with stellar evolution calculations and find that the mass-loss rate should increase along the AGB for stars with initial masses greater than 1.3 M⊙. Stars with lower initial mass would experience a single short-lived (∼ 1000 yr) episode of high mass loss only, which would leave behind a very narrow detached shell as observed in the case of, e.g., TT Cyg (Olofsson et al. 2000). On the other hand, the mass-loss phenomenon appears to be highly variable on even shorter time scales as indicated e.g. by concentric arcs observed in scattered light around the prototype carbon Mira IRC +10216 (Mauron & Huggins 2000) and around some proto-planetary nebulae (e.g., Hrivnak et al. 2001). The time scale of these mass-loss variations would be around a few 102 yr. The physical mechanism responsible for these variations still needs to be identified, although different possibilities have already been proposed: e.g., interaction between gas and dust within stellar outflows (Simis et al. 2001), or solar-like magnetic cycle (Soker 2002). In contrast to these later phases of AGB mass loss at rather high rates (∼ 10−5M⊙ yr−1), information about the mass-loss process on the early AGB, is even scarcer. To unravel the processes involved in the mass-loss phenomenon, we have to find suitable trac- ers. One of the most studied among these tracers is the CO molecule, because so far it has been considered to provide the best estimate of the mass-loss rate for AGB stars (Ramstedt et al. 2008). Not only can it be used to estimate this mass-loss rate, but it also yields important parameters of the AGB wind (e.g.: expansion velocity, central star velocity, etc.). Since CO is photodissociated by UV radiation from the interstellar radiation field (ISRF), it can only probe the inner parts (r ≤ 10−3 - 10−1 pc, Mamon et al. 1988) of circumstellar shells (CSs). Therefore, the CO emission is only related to "recent" (i.e. a few 103 - 104 years) mass-loss episodes. On the other hand, H i is in general protected from photoionization by the surrounding inter- stellar medium (ISM). As a result, H i can be used to probe the external parts of circumstellar shells and can give indications on the mass-loss on longer timescales (a few 105 years: Libert et al. 2007). Hence, CO and H i complement each other nicely to describe the history of the mass-loss rate of an AGB star. The drawbacks of H i circumstellar observations are that hydrogen is ubiquitous in the Galaxy and that the genuine stellar H i must be separated from the ambient H i. Ideal cases would be bright H i sources, with relatively high velocity with respect to the local standard of rest (LSR) and Libert et al.: H i & CO in RX Lep. 3 reasonably far above the Galactic plane. For the other sources, the interstellar H i should be studied with care. In the present paper, we analyze this confusion with a new approach that consists in a 3D-mapping using H i spectra. The mass-loss phenomenon is different from one AGB star to another and may vary highly with time. Nevertheless, observing in H i provides a global view of the CS behavior and, on timescales of about 105 years, small variations in the mass-loss rate may be flattened out. Thus, we have developed a model of the circumstellar gas, based on a scenario already proposed by Young et al. (1993), in which CSs are the result of a constant outflow eventually slowed down by the sur- rounding medium. This deceleration produces a snowplough effect around the source, resulting in a detached shell of compressed matter originating from the star and the external medium (Lamers & Cassinelli 1999, Chap. 12). A schematic view of this model can be pictured as follows: a wind is flowing outward from the star, in free expansion with a constant velocity (Vexp) and a constant rate. It encounters a shock at a radius r1 (termination shock), due to the slowing down by the surrounding matter. Between r1 and rf (contact discontinuity), the stellar matter is compressed. Between rf and until a second shock at r2 (bow shock), the interstellar matter has been swept up by the wind of the AGB star. Finally, beyond r2, the external matter is considered to be at rest. Recently, we successfully applied this model to a carbon-rich star: Y CVn (Libert et al. 2007). In H i at 21 cm, this star exhibits a composite profile, made of a broad, rectangular component and a narrow, Gaussian-shaped one. In our description, the broad component is the signature of the freely expanding wind, whereas the narrow component is produced by the H i compressed in the snowplough between r1 and r2. Our model provides a simple explanation for some of the so-called "detached dust shells" observed in the far infrared (Izumiura et al. 1996). If this approach is correct, then it should also apply to detached shells around oxygen-rich AGB stars. In this paper we present H i and CO data that we obtained on an oxygen-rich AGB star, RX Lep, and interpret them with the model that we developed for Y CVn. The H i interstellar confusion in the direction of RX Lep is moderate and we illustrate, in that case, our new approach to extract a genuine H i spectrum. In this simplified description we assume spherical symmetry. However, recent H i, far-infrared and UV data (G´erard & Le Bertre 2006; Matthews & Reid 2007; Ueta et al. 2006; Martin et al. 2007) have shown that the AGB star motion with respect to the ISM may lead to a distortion, and eventually a disruption, of the circumstellar environment. Previous, and more recent, numerical modelings (Villaver et al. 2003; Wareing et al. 2007) are in line with this interpretation. The cir- cumstellar environment of RX Lep might provide a new illustration of this phenomenon, and we will discuss this possibility. 2. RX Lep RX Lep has been classified as an irregular variable, Lb star (General Catalogue of Variable Stars, GCVS 3rd ed., Kukarkin et al. 1971). A photometric monitoring over 8 years shows variations of about ± 1 magnitude in the V band (Cristian et al. 1995). The periodogram analysis gives a main period in the range 80-100 days and a possible secondary period around 60 days. Recently, the star has been re-assigned to the type SRb (GCVS 4.2, Samus et al. 2004), because it may exhibit a periodic variability of a few tenths of a magnitude. 4 Libert et al.: H i & CO in RX Lep. The Hipparcos parallax (7.30±0.71 mas) places the star at 137+15 −12 pc from the Sun and at ∼ 65 pc away from the Galactic plane (bII=-27.51◦). The proper motion, also given by Hipparcos, is 31.76±0.58 mas yr−1 in right ascension (RA) and 56.93±0.50 mas yr−1 in declination (Dec.). At 137 pc, it translates into a motion in the plane of the sky of 44 km s−1 (corrected for solar motion, as determined by Dehnen & Binney 1998) in the northeast direction (PA ∼ 31◦). Fouqu´e et al. (1992) have obtained near-infrared photometry data and, using the bolometric correction of Le Bertre et al. (2001), we derived a luminosity of 4500 L⊙. This luminosity con- firms that RX Lep is on the AGB. The effective temperature is ∼ 3300 K (Dumm & Schild 1998). This means that hydrogen is expected to already be mostly in atomic form in the atmosphere and throughout the CS (Glassgold & Huggins 1983). Technetium lines (99Tc) were searched for in the 4200-4300 Å region and not detected (Lebzelter & Hron 1999), confirming an older result from Little et al. (1987). This tends to indicate that RX Lep has not gone through a thermal pulse and that it is still in the early phase of the AGB (E-AGB). This is in good agreement with the results of Mennessier et al. (2001) who, using astrometric and kinematic data, place RX Lep among E-AGB stars that belong to the Galactic disk population with initial masses in the range 2.5-4 M⊙. Paschenko et al. (1971) did not detect the source in the OH satellite line at 1612 MHz. As this is the only OH observation reported in the literature, we observed RX Lep again at 18 cm on Jan. 12, 2006 and July 14, 2006 with the NRT. No emission was detected at a level of 0.015 Jy in any of the 4 OH lines (1612, 1665, 1667, and 1720 MHz). RX Lep might be associated with an IRAS extended source (X0509-119, IRAS Science Team 1988) at 60 (diameter ∼ 1.1′) and 100 µm (diameter ∼ 6.0′). However, X0509-119 is centered at about 2.5′ east from RX Lep, and that association might only be fortuitous. We present a re-analysis of the IRAS results farther down (Sect. 6). Kerschbaum & Olofsson (1999) report a CO (1-0) and CO (2-1) detection, but their radial ve- locity is doubtful: vhel ∼ 29 km s−1 (as compared to vhel ∼ 46 km s−1 cited in the General Catalogue of Stellar Radial Velocities, GCRV, Wilson 1953). Our new results (Sect. 3) now suggest there has likely been a confusion between the heliocentric and LSR reference frames. 3. Molecular line observations RX Lep was part of a CO observing program dedicated to the Valinhos 'b' class stars (Epchtein et al. 1987). This class of sources is defined by a weak near-IR excess as compared with the IRAS fluxes (0.2 < K-L' < 0.7 and 0.8 < L'-[12] < 2). The central stars are generally identified with late-M giants surrounded by tenuous circumstellar shells. Those stars were suspected by Winters et al. (2000) to show preferentially low expansion velocity winds. Most of the data from this pro- gram have been published in Winters et al. (2003). Subsequently, RX Lep's CO (2-1) emission at 230 GHz and SiO (v=1, J=2-1) maser transition at 86 GHz have been searched using the 15-m Swedish-ESO Submillimetre Telescope, SEST (Booth et al. 1989) on January 30, 2003. At 1.3 mm, the FWHM of the SEST beam is 23′′. We used the position-switch mode with a beam throw of 11.5′. The spectra were recorded on the high-resolution spectrometer (HRS) giving a resolution of 80 kHz, for a channel separation of 43 kHz and a bandwidth of 86 MHz. The CO (2-1) transition was clearly detected (Fig. 1). The resulting profile was fitted with a parabolic curve, and we derived an LSR velocity of Vlsr = 28.9 ± 0.1 km s−1, an expansion Libert et al.: H i & CO in RX Lep. 5 Fig. 1. CO (2-1) emission of RX Lep. The solid line represents a parabolic fit (Sect. 3). velocity of Vexp = 4.2 ± 0.1 km s−1, and an amplitude of Tmb = 0.45 K ± 0.13 K. By using the same method as Winters et al. (2003, Sect. 4.3), we estimated both RX Lep mass-loss rate and CO photo-dissociation radius using the results of the line fitting. We find M ∼ 1.7 × 10 −7 M⊙ yr−1 and rCO ∼ 0.8 × 10−2 pc (≡ 12.5′′). On the other hand, the SiO maser was not detected at a level of 0.2 Jy. Our CO measurement of the LSR radial velocity (28.9 km s−1) is consistent with the heliocen- tric velocity quoted in the GCRV. Combining this result with the Hipparcos determination of the velocity in the plane of the sky (44 km s−1), we get a 3-D space velocity of 53 km s−1. 4. H i observations RX Lep has been observed during a total of 141 hours between February 2005 and February 2008 with the Nanc¸ay Radiotelescope (NRT). The NRT is a meridian telescope with a rectangular aper- ture of effective dimensions 160×30 m. At 21 cm and at the declination of RX Lep, the FWHM of the beam is 4′ in RA and 22′ in declination. We used the position-switch technique with two off-positions in the east-west direction, every 2′, and up to 24′ from the source. Thus, the total time spent on-source was 47 hours. To fully describe the environment of RX Lep, we sampled our map every half beam in RA and in Dec. (hereafter, the position-switch spectra will be referred to as C ± n × 4′ EW, n being the number of beams for the off positions). At 21 cm, the spectra have a bandwidth of 165 km s−1 and a channel width of 0.08 km s−1. For convenient analysis, we smoothed the data with a Hanning filter so that the spectral resolution was 0.16 km s−1. The data are processed with the CLASS software, part of the GILDAS1 package developed at IRAM (Pety 2005). The different steps of the data processing can be described as follows. First of all, we determine the spatial extent of the source by comparing the C ± n × 4′ EW spectra. When the maximum inten- sity of the peak is reached (n = nmax), the source does not contribute to the flux of the offset spectra anymore. For example, according to Fig. 2 (upper panel), RX Lep does not extend farther than 6′ in the E-W direction. Once the maximum extent is estimated, the average of the C ± n × 4′ EW spec- tra with n > nmax gives the central spectrum (Fig. 2, lower panel). Simple arithmetic then allows to extract the spectra at the offset positions, using the central spectrum. We present a new visualization of the H i spectra to better separate the genuine stellar H i from the contamination due to interstellar hydrogen. The operation can be described as a stacking of the 1 http://www.iram.fr/IRAMFR/GILDAS 6 Libert et al.: H i & CO in RX Lep. Fig. 2. Upper panel: Spectra obtained in position-switch mode with the NRT. The positions are expressed in number of beams (4′). The maximum intensity is reached at C±1.5 × 4′ EW. For clar- ity, the individual spectra have been displayed with vertical offsets of 0.4 Jy. Lower panel: average spectrum computed with C ± n × 4′ EW, n > 1.5. Dotted line: baseline subtracted f-switch spectrum. The horizontal line shows the width of the CO signal. Fig. 3. Left panel: 3D velocity-position representation of the H i flux density; east is to the left. The arrow points to the expected position of the source. Right panel: the same data set represented in 2D; west is to the right. The dashed circle surrounds the expected position of RX Lep. Libert et al.: H i & CO in RX Lep. 7 Fig. 4. H i map of RX Lep. The steps are 2′ in RA and 11′ in Dec. The positions are indicated with respect to the stellar position. The abscissae and ordinates are LSR radial velocities (km s−1) and flux densities (Jy) as indicated on the lower left corner. NRT spectra, processed as above, in the east-west direction, for a given declination (Fig. 3). In this view, velocity is given as a function of right ascension, and intensity is represented using a colored scale. This aims at visualizing, and thus separating, the H i emission coming from the source and that from the Galaxy. Indeed, on the resulting image, the stellar H i should be nearly centered in RA and close to the LSR velocity given by CO observations. While this process emphasizes the difficulties coming from the contamination due to the Galactic hydrogen emission, it also allows an evaluation of the possible problems when processing the data and a design of the best strategy for extracting the intrinsic source emission. According to Fig. 3, RX Lep is definitely a suitable candidate for H i observation, as it is clearly separated from the interstellar emission spectrally and spatially, although the confusion increases for velocities lower than 29 km s−1. Indeed, the image shows 2 potential sources of contamination: one around 4′W from the source and at ∼ 20 km s−1, the other increasing (negatively) at 16′E and around 24 km s−1. This information is crucial to safely extracting the intrinsic emission of RX Lep. Thus, to build the map of the source, we fitted polynomial baselines (in some cases of degree up to 3 when the confusion reaches its highest level) to a portion of the spectrum between 21 and 54 km s−1. The resulting map of RX Lep is shown in Fig. 4. We independently confirmed these results by observing the source using the frequency-switch mode (Fig. 2, lower panel). We spent 5 hours on source and detected it at the same velocity and with the same flux density as shown on the map for the central position. From our observations, we can readily derive some important properties of the CS. The map of RX Lep in Fig. 4 reveals that the H i line profile, at the central position, shows a quasi- 8 Libert et al.: H i & CO in RX Lep. Gaussian shape of central velocity 28.84±0.03 km s−1, FWHM 3.8±0.1 km s−1, and flux density 0.22±0.03 Jy. The shape of this line differs from that of the parabolic CO line. It clearly indicates a slowing down of the wind within the outer parts of the CS (Le Bertre & G´erard 2004). Moreover, assuming that the broadening of the H i emission line is dominated by the thermal Doppler effect, the FWHM of the spectrum allows us to estimate an upper limit to the average temperature in the shell (Libert et al. 2007, Eq. 1). It gives us Tmean < 312 K. Evidence of a composite line profile such as that of Y CVn (Sect. 1) is difficult to see, given the fairly low intensity of the signal and the narrow width of the expected pedestal (2 × Vexp). Nevertheless, from the central spectrum, we can set an upper limit to the amplitude of a possible pedestal. We estimate this limit at 20 mJy by assuming it has the same width as the CO profile (Fig. 5). The map of RX Lep (Fig. 4) shows that the H i brightness distribution of the envelope is offset from the stellar position both in RA and Dec. There is a slight westward RA offset < 1′ (since the flux density at 2′ west is larger than at 2′ east but smaller than at the center). There is a southward Dec. offset close to 5.5′ (since the flux density at 11′ south is nearly equal to the central flux). It is useful to give quantitative estimates, not only of the offsets but also of the spatial extents for the model calculations that will be discussed in Sect. 5. If one assumes that the H i brightness distribution is Gaussian in RA (and Dec.) and offset, the convolution by a Gaussian beam also produces a Gaussian distribution and one can retrieve from the data both the offset and half power width (HPW) in RA (and Dec.). The RA offset and HPW are respectively −0.4′ (± 0.2′) and 2.3′ (± 0.5′). The Dec. offset and HPW are respectively −4.4′ (± 0.6′) and 15′ (± 3′). Thus the H i envelope is elongated southwards and offset from the stellar position by 4.4′ at PA 185◦ (i.e. also nearly southward). This suggests an HI envelope trailing south. The integrated flux throughout the map gives 1.22 Jy×km s−1, which translates into a hydrogen mass of ∼ 5.42 × 10−3 M⊙ (assuming no hydrogen in H2; cf. Sect. 2). Adopting a mean molecular weight of 1.3, it translates into a total mass of the gas in the shell of ∼ 7.05 × 10−3 M⊙. If we consider the mass-loss rate to be constant and adopt the estimate given by CO, then the age of the CS is ∼ 42700 years, about one order of magnitude less than the age we estimated for Y CVn. Fig. 5. Average of the spectra C ± n × 4′ EW and 11′S ± n × 4′ EW with n > 1. The horizontal line represents the width of the CO line. Libert et al.: H i & CO in RX Lep. 9 5. Model The high-quality H i spectral profiles that we have obtained in the direction of RX Lep are similar to those of Y CVn (Libert et al. 2007). This type of profile is indicative of a slowing-down of stellar outflows in the external parts of CSs (Le Bertre & G´erard 2004). In the following we apply the model that we developed for the carbon-rich star Y CVn in order to evaluate the physical conditions within the shell of RX Lep. Of course, as this model assumes sphericity, it cannot reproduce the more complex geometry suggested by the map presented in the previous section. In the east-west direction the map is fairly symmetric, so the model could apply. However, there is also a clear north-south extension that would require a 2-D model, as well as a spatial resolution better than 22′ (the NRT beam). A 1D-hydrodynamic code provides the density distribution within the detached shell based on the hypothesis of a slowing down of the stellar gas by the surrounding local material. The mass-loss rate is constant, and the gas expanding outward from the atmosphere is in free expansion with a constant velocity Vexp. Then, the outflow encounters a shock (r1). Its velocity decreases by a factor of about 4 and the matter keeps on decelerating until it reaches the external medium (at rf). The external matter that has been swept up and compressed by the expansion of the stellar envelope lies outside rf. Finally, beyond r2, the gas is at rest. The expansion velocity and the LSR velocity of the source are based on the results from our CO observations. But RX Lep has not been studied much, so we lack some spatial information such as estimates of r1 and r2 that could have been obtained, for example, with dust continuum observations. Nevertheless, the NRT map indicates that the object is fairly small in the east/west direction (∼ 2.3′ i.e. ∼ 9 × 10−2 pc at 137 pc). One of the results of our model is that detached shells are flagged by a composite H i spectrum. The first component (Comp. 1) is narrow, with a quasi-Gaussian shape and it arises from the matter slowed down by the local medium. The second component (Comp. 2) is broad with a rectangular shape, as it probes the inner part of the shell where the gas is in free expansion. In Sect. 4, we set an upper limit of Comp. 2 of ∼ 20 mJy. For a constant mass-loss rate, we can derive a relation (Eq. 1) to estimate r1: r1 ≈ 2.17 × 10−9 × d V2 exp FComp. 2 M (1) where r1 is expressed in arcmin, d is the distance in pc, Vexp is in km s−1, FComp.2 is the intensity of the pedestal in Jy, and M is in M⊙ yr−1. With Vexp = 4.2 km s−1 and M = 1.65 × 10−7 M⊙ yr−1 (Sect. 3), we estimate an upper limit for r1 of 0.64′. We set r2 at 1.15′, in agreement with the H i observations in the east-west direction (HPW/2, Sect. 4). As our model assumes spherical symmetry, we performed a fitting on a symmetrized map, i.e. a map in which the offset positions have been averaged (Fig. 6, upper panel). In the model, the total flux is forced to be equal to that measured in the map (i.e. 1.22 Jy × km s−1). We set the central velocity at 28.8 km s−1. The results are summarized in Table 1. In this simulation, the temperature and the velocity are constant inside r1 (resp. 20 K and 4.2 km s−1). The shock at r1 decreases the velocity (increases the density) by a factor of 3.9 and the temperature rises to 530 K (Figs. 7 & 8). Then, inside the region of compressed matter (between r1 and rf), the temperature decreases to ∼ 175 K. The physical conditions between rf and r2 are in fact not constrained either by our model 10 Libert et al.: H i & CO in RX Lep. Fig. 6. H i observations vs model (dashed line): the upper panel shows the model discussed in Sect. 5 and compared to a symmetrized H i map of RX Lep. The lower panel presents the same model shifted by 4.4′S and 0.4′W and compared to the H i map of RX Lep (as in Fig. 4). or by the data at 21 cm, and in Table 1 they are only extrapolated (for more details, see Libert et al. 2007). The model assumes that the H i emission is optically thin (τ ≪ 1). This can be verified using the output column density profile (Fig. 8, right panel) and the expression τ = 5.50 × 10−19 NH T ∆V (Eq. 12, Libert et al., 2007) where NH is in cm−2 and ∆V, the line width, in km s−1. With T > 175 K and ∆V ∼ 3.8 km s−1, the optical depth stays below 0.5 at all impact parameters > 0.1′ from the central star. In general, the model provides a satisfactory fit to the symmetrized H i spectra that we have obtained on RX Lep. However, it predicts a flux above the observations on the central position and below at 22′ in declination. This can be understood as a consequence of the 4.4′ offset to the south noted in Sect. 4. By moving the model 4.4′ south and 0.4′ west with respect to the central star, we can improve the fit to the observed data (Fig. 6, lower panel). This gives support to the offset values that we have determined by Gaussian-fitting in Sect. 4. Yet, the spectrum on the position at 22′ south is not well reproduced, suggesting that the source is more extended along the north-south direction than along the east-west one, as suspected in Sect. 4. In the past (Y CVn, Libert et al. 2007), we already attempted to better fit the data by a shift in position to take into account the deformation of the envelope by the ISM. However, this approach is Table 1. Model parameters (d = 137 pc), with notations the same as in Libert et al. (2007) Libert et al.: H i & CO in RX Lep. 11 M (in hydrogen) 1 ), T+ 1 µ t1 tDS r1 r f r2 T0(≡ T− T f (= T2) v0(≡ v− v f v2 n− 1 , n+ 1 n− f , n+ f n2 Mr<r1 (in hydrogen) MDT,CS (in hydrogen) MDT,EX (in hydrogen) 1 ), v+ 1 1.27 × 10−7 M⊙ yr−1 1.3 5 927 years 36 800 years 2.55 × 10−2 pc (0.64 ′) 3.67 × 10−2 pc (0.92 ′) 4.58 × 10−2 pc (1.15 ′) 20 K, 528 K 175 K 4.2 km s−1, 1.07 km s−1 0.16 km s−1 1.2 km s−1 148 H cm−3, 578 H cm−3 2.1 × 103 H cm−3, 2.5 H cm−3 1.3 H cm−3 0.75 10−3 M⊙ 4.67 10−3 M⊙ 0.010 10−3 M⊙ Fig. 7. Upper panel: velocity profile. The dashed line represents the isothermal sound velocity. Lower panel: temperature profile adopted for the model. artificial because we used a spherical model that is not centered on the star position. It is only meant to illustrate the need for an H i mapping of this interesting source with a better spatial resolution and the need to develop a true non-spherical modeling of detached shells. 12 Libert et al.: H i & CO in RX Lep. Fig. 8. Left panel: atomic hydrogen density profile. Center panel: atomic hydrogen mass-flow pro- file. Right panel: atomic hydrogen column density calculated by the model. The vertical dotted lines show the radii, r1, rf and r2, used in the model. 6. Discussion RX Lep shows evidence of a circumstellar envelope of ∼ 0.01 M⊙ that may be the result of its stellar wind decelerated by the external medium. This star is an oxygen-rich, semi-regular variable on the E-AGB (no evidence of Tc, cf. Sect. 2). It is in the same evolutionary stage as EP Aqr and X Her, which have also been detected in H i and for which the emission at 21 cm shows evidence of significant circumstellar envelopes (Le Bertre & G´erard 2004; Gardan et al. 2006). We note that these 3 stars share the same variability properties and have about the same luminosity (∼ 4500 L⊙) and the same effective temperature (∼ 3200 K). It suggests that mass loss can already occur efficiently for this type of star on the E-AGB. Our model strongly relies on the mass-loss rate estimated from CO observations. It is note- worthy that this estimate is consistent with Reimers' relation (Reimers 1978). Indeed, by adopting M ∼ 3 M⊙, L ∼ 4500 L⊙ and Teff ∼ 3300 K (Sect. 2), this relation gives M ∼ 2 × 10−7 M⊙ yr−1. However, the luminosity was probably lower in the past, as was the mass-loss rate. This suggests that the age (42 700 years) is underestimated. The model and our observations together put constraints on the physical conditions within the CS between the termination shock (r1) and the interface (rf). Directly from the observations, the mean temperature should be . 300 K. Based on the assumption of an adiabatic shock at r1, it implies an increase in temperature to ∼ 500 K. Thus, the gas must be cooled down in the CS. Estimating the cooling rate is difficult at such low temperatures. Nevertheless, the H i line-profiles put constraints on the behavior of the temperature because it is coupled to the kinematics (Libert et al. 2007). The temperature profile shown in Fig. 7 (lower panel) yields the best fit to the shape of the H i spectra. Between rf and r2, our model has only been extrapolated. This region is probably dominated by interstellar material flowing at ∼ 50 km s−1 through the bow shock. The material should be denser than assumed in our model; indeed, this region is fed by the interstellar medium that has been swept up through the relative motion of RX Lep circumstellar shell, at ∼ 50 km s−1, rather than by the expansion of the shell during the same period of 4 104 years. Also it is expected to be ionized, and therefore might not contribute significantly to the H i emission that we detected. The H i data indicate that RX Lep's CS is offset about 4′ to the south and 0.5′ to the west. This agrees within 25◦ with the direction of the proper motion given by Hipparcos (PA ∼ 31◦). In addi- tion, the model hints that the shell is not completely spherical, and that RX Lep is slightly elongated Libert et al.: H i & CO in RX Lep. 13 Fig. 9. Reprocessed IRAS images (IRIS) at 60 µm (upper panels) and 100 µm (lower panels). To enhance the suspected extended emission to the south, we present a non-saturated version (left) and a saturated one (right) for both wavelengths. The field is ∼ 65′ × 39′ and the green reticles mark the position of RX Lep (north is to the top and east to the left). mostly in the N/S direction. This suggests that the elongated shape observed in H i is connected to the motion of RX Lep through the local ISM. Villaver et al (2003) have made numerical simula- tions of the evolution of a low-mass star moving supersonically through the ISM and find that, due to the ram-pressure stripping, most of the mass ejected during the AGB phase is left downstream. The left panel in their Fig. 1 shows that CSs are progressively distorted and become elongated in the direction of the motion with respect to their surrounding ISM. The 25◦ difference between the space motion of the star and the elongation of the shell could then be due to the intrinsic velocity of the ISM local to RX Lep relative to the LSR. Such intrinsic motions are currently found in the local solar neighborhood (Redfield & Linsky, 2008, Fig. 16 and references therein). A significant fraction of the velocity of the local ISM is a reflection of the solar motion; nevertheless, the velocity of the Sun with respect to the LSR (13.4 km s−1, according to Dehnen & Binney, 1998) is 25◦ away from the direction of the velocity of the average local ISM with respect to the Sun (26.7 km s−1). It is also worth noting that for RX Lep there is no significant difference between the H i and CO central velocities, as if the interaction only occurs in the plane of the sky. We have examined the IRAS maps that have been reprocessed recently by Miville-Deschenes & Lagache (2005; IRIS: Improved Reprocessing of the IRAS Survey). The 60 µm and 100 µm images (Fig. 9) suggest a small extended source (φ ∼ 6′ - 8′). The source at 100 µm might be shifted by ∼ 2′ to the south. There is also an extension (∼ 12′) to the south; however, it might be an artifact due to the satellite scanning in the north-south direction. In these images, we cannot confirm the X0509-119 offset with respect to RX Lep (cf Sect. 2). New data with a better spatial resolution, e.g. from the Far Infrared All-Sky Survey of Akari, may help to clarify this situation. In their H i survey of evolved stars, G´erard & Le Bertre (2006) found that the line-profiles are Gaussian-shaped and often offset with respect to the stellar velocity by ∼ 1-3 km s−1 towards 0 km s−1 LSR. Several H i sources were also noted to be spatially offset from the central star. They suggest that these effects could be related to a non-isotropic interaction with the local ISM. Matthews & Reid (2007) have imaged the H i emission around RS Cnc with the VLA. They find that it is elongated with a peak on the stellar position and a filament extending ∼ 6′ to the northwest, 14 Libert et al.: H i & CO in RX Lep. in a direction opposite to that given by the proper motion. Recently, Matthews et al. (2008) have imaged the H i emission of Mira with the VLA. As for RS Cnc, they find a "head-tail" morphology oriented along the star proper motion and consistent, on large scales, with the far-ultraviolet emis- sion discovered by GALEX (Martin et al. 2007). Furthermore, the high spectral resolution H i data obtained with the NRT along the 2-degree GALEX trail reveal a deceleration of the gas caused by interaction with the local ISM. Finally, using Spitzer MIPS data obtained on R Hya at 70 µm, Ueta et al. (2006) discovered a bow-shock structure ahead of the star in the direction of its motion. The excess emission that delineates this bow shock is seen at 70 µm, but not at 160 µm; it may partly come from the [O i] cooling line at 63 µm. Although we have presently no direct evidence in H i of a bow-shock, both structures, bow-shock and H i trail, should be causally related (Wareing et al. 2006). In fact, as the velocity of these sources with respect to the ISM is often high (see e.g. Nyman et al. 1992, or Mennessier et al. 2001), the interstellar material is probably ionized through the bow shock, so that we may never detect directly such a bow-shock structure in H i at 21 cm. Better tracers would likely be line emission in the UV/optical/IR ranges (Hα, [Fe ii], [O i], etc.). We therefore have a convergent set of results that shows that AGB stars are associated with large-scale circumstellar shells distorted by the motion of these evolved objects through the ISM (Villaver et al. 2003). We suggest that RX Lep is one more source in such a case. That the source is elongated in the same direction as its offset and nearly opposite to the direction of motion, argues in favor of a head-tail morphology. 7. Conclusions We detected CO(2-1) and H i line emissions from the semi-regular oxygen-rich E-AGB star, RX Lep. These emissions indicate a stellar outflow at a velocity ∼ 4.2 km s−1 and a rate ∼ 1.7 × 10−7 M⊙ yr−1, with a duration of 4 × 104 years. The H i source has a size of ∼ 2′ (≈ 0.08 pc) in the east-west direction and possibly 15′ (≈ 0.6 pc) in the north-south direction. The modeling of the H i line profiles obtained at different positions suggests that the outflow is slowed down by the interaction with the ambient ISM, and that the external part of RX Lep circumstellar shell is made of compressed material, at ∼ 200 K, as in the well-known detached shell around Y CVn. The elongated shape of the RX Lep H i source is compatible with the direction of its proper motion, as in the cases of Mira and RS Cnc, which have already been studied at high angular resolution with the VLA. Acknowledgements. The Nanc¸ay Radio Observatory is the Unit´e scientifique de Nanc¸ay of the Observatoire de Paris, associated as Unit´e de Service et de Recherche (USR) No. B704 to the French Centre National de la Recherche Scientifique (CNRS). The Nanc¸ay Observatory also gratefully acknowledges the financial support of the Conseil R´egional de la R´egion Centre in France. This research made use of the SIMBAD database, operated at the CDS, Strasbourg, France, and of the NASA's Astrophysics Data System. We thank the referee, Dr T. Ueta, and Dr L. Matthews for helpful suggestions. References Booth, R.S., Delgado, G., Hagstrom, M., et al., 1989, A&A, 216, 315 Cristian, V.-C., Donahue, R. A., Soon, W. H., Baliunas, S. L., & Henry, G. W., 1995, PASP, 107, 411 Dehnen, W., & Binney, J.J., 1998, MNRAS, 298, 387 Dumm, T., & Schild, H., 1998, New Astron., 3, 137 Libert et al.: H i & CO in RX Lep. 15 Epchtein, N., Le Bertre, T., L´epine, J.R.D., et al., 1987, A&AS, 71, 39 Fouqu´e, P., Le Bertre, T., Epchtein, N., Guglielmo, F., & Kerschbaum, F., 1992, A&AS, 93, 151 Gardan, E., G´erard, E., & Le Bertre, T., 2006, MNRAS, 365, 245 G´erard, E., & Le Bertre, T., 2006, AJ, 132, 2566 Glassgold, A. E., & Huggins, P. J., 1983, MNRAS, 203, 517 Habing, H., 1996, A&AR, 7, 97 Hrivnak, B.J., & Bieging, J.H., 2005, ApJ, 624, 331 Hrivnak, B.J., Kwok, S., & Su, K.Y.L., 2001, AJ, 121, 2775 IRAS Science Team, 1988, IRAS Catalogs and Atlases, NASA RP-1190, vol. 7 Izumiura, H., Hashimoto, O., Kawara, K., Yamamura, I., & Waters, L. B. F. M., 1996, A&A, 315, L221 Kerschbaum, F., & Olofsson, H., 1999, A&AS, 138, 299 Knapp, G. R., & Morris, M., 1985, ApJ, 292, 640 Kukarkin, B. V., Kholopov, P. N., Pskovsky, Y. P., et al., 1971, General Catalogue of Variable Stars, 3rd ed. Lafon, J.-P.J., & Berruyer, N., 1991, A&AR, 2, 249 Lamers, H.J.G.L.M., & Cassinelli, J.P., 1999, "Introduction to Stellar Winds", Cambridge University Press, Cambridge Leao, I.C., de Laverny, P., M´ekarnia, D., De Medeiros, J.R., & Vandame, B., 2006, A&A, 455, 187 Le Bertre, T., & G´erard, E., 2004, A&A, 419, 549 Le Bertre, T., Matsuura, M., Winters, J. M., et al., 2001, A&A, 376, 997 Lebzelter, Th., & Hron, J., 1999, A&A, 351, 533 Libert, Y., G´erard, E., & Le Bertre, T., 2007, MNRAS, 380, 1161 Little, S. J., Little-Marenin, I. R., & Bauer, W. H., 1987, AJ, 94, 981 Mamon, G.A., Glassgold, A.E., & Huggins, P.J., 1988, ApJ, 328, 797 Martin, D.C., Seibert, M., Neill, J.D., et al., 2007, Nature, 448, 780 Matthews, L.D., Libert, Y., G´erard, E., Le Bertre, T., & Reid, M.J., 2008, ApJ, 684, 603 Matthews, L.D., & Reid, M.J., 2007, AJ, 133, 2291 Mauron, N., & Huggins, P. J., 2000, A&A, 359, 707 Mennessier, M. O., Mowlavi, N., Alvarez, R., & Luri, X., 2001, A&A, 374, 968 Miville-Deschenes, M.-A., & Lagache, G., 2005, ApJS, 157, 302 Nyman, L.-Å., Booth, R. S., Carlstrom, U., et al., 1992, A&AS, 93, 121 Olofsson, H., 1999, Asymptotic Giant Branch Stars, IAU Symposium ♯191, ed. T. Le Bertre, A. L`ebre, & C. Waelkens, p. 3 Olofsson, H., Bergman, P., Lucas, R., et al., 2000, A&A, 353, 583 Olofsson, H., Gonz´alez Delgado, D., Kerschbaum, F., & Schoier, F.L., 2002, A&A, 391, 1053 Paschenko, M., Slysh, V., Strukov, I., et al., 1971, A&A, 11, 482 Pety, J., 2005, in SF2A-2005: Semaine de l'Astrophysique Franc¸aise, ed. F. Casoli, T. Contini, J. M. Hameury, & L. Pagani, 721 Ramstedt, S., Schoıer, F.L., Olofsson, H., & Lundgren, A.A., 2008, A&A, 487, 645 Redfield, S., & Linsky, J.L., 2008, ApJ, 673, 283 Reimers, D., 1978, A&A, 67, 161 Renzini, A., 1981, in Physical processes in Red giants, Iben Jr., I. & Renzini, A. (Eds.), D. Reidel, p.431 Samus, N. N., Durlevich, O. V., et al., 2004, Combined General Catalog of Variable Stars (ed. 4.2; Moscow: Sternberg Astron. Inst.) Schroder, K.-P., Winters, J.M., & Sedlmayr, E., 1999, A&A, 349, 898 Simis, Y.J.W., Icke, V., & Dominik, C., 2001, A&A 371, 205 Soker, N., 2002, ApJ, 570, 369 Ueta, T., Speck, A. K., Stencel, R. E., et al., 2006, ApJ, 648, L39 Villaver, E., Garc´ıa-Segura, G., & Manchado, A., 2003, ApJ, 585, L49 Wareing, C.J., Zijlstra, A.A., & O'Brien, T.J., 2007, MNRAS, 382, 1233 Wareing, C.J., Zijlstra, A.A., Speck, A.K., et al., 2006, MNRAS, 372, L63 Wilson, R. E., 1953, General Catalogue of Stellar Radial Velocities, Carnegie Inst. Washington D. C. Publ. 601 Winters, J. M., Le Bertre, T., Jeong, K. S., Helling, Ch., & Sedlmayr, E., 2000, A&A, 361, 641 Winters, J. M., Le Bertre, T., Jeong, K. S., Nyman, L.-Å., & Epchtein, N., 2003, A&A, 409, 715 Young, K., Phillips, T. G., & Knapp, G. R., 1993, ApJ, 409, 725
astro-ph/0701430
2
0701
2007-02-23T08:34:06
"Late prompt" emission in Gamma Ray Bursts?
[ "astro-ph" ]
The flat decay phase in the first 1e2-1e4 seconds of the X-ray light curve of Gamma Ray Bursts (GRBs) has not yet found a convincing explanation. The fact that the optical and X-ray lightcurves are often different, with breaks at different times, makes contrived any explanation based on the same origin for both the X-ray and optical fluxes. We here assume that the central engine can be active for a long time, producing shells of decreasing bulk Lorentz factors Gamma. We also assume that the internal dissipation of these late shells produces a continuous and smooth emission (power-law in time), usually dominant in X-rays and sometimes in the optical. When Gamma of the late shells is larger than 1/theta_j, where theta_j is the jet opening angle, we see only a portion of the emitting surface. Eventually, Gamma becomes smaller than 1/theta_j, and the entire emitting surface is visible. Thus there is a break in the light curve when Gamma=1/theta_j, which we associate to the time at which the plateau ends. After the steeply decaying phase which follows the early prompt, we see the sum of two emission components: the "late-prompt" emission (due to late internal dissipation), and the "real afterglow" emission (due to external shocks). A variety of different optical and X-ray light curves are then possible, explaining why the X-ray and the optical light curves often do not track each other (but sometimes do), and often they do not have simultaneous breaks.
astro-ph
astro-ph
Draft version July 4, 2019 Preprint typeset using LATEX style emulateapj 7 0 0 2 b e F 3 2 2 v 0 3 4 1 0 7 0 / h p - o r t s a : v i X r a "LATE PROMPT" EMISSION IN GAMMA RAY BURSTS? G. Ghisellini, G. Ghirlanda, L. Nava1, C. Firmani2 INAF -- Osservatorio Astronomico di Brera, via Bianchi 46, I -- 23807 Merate, Italy Draft version July 4, 2019 ABSTRACT The flat decay phase in the first 102 -- 104 seconds of the X -- ray light curve of Gamma Ray Bursts (GRBs) has not yet found a convincing explanation. The fact that the optical and X -- ray lightcurves are often different, with breaks at different times, makes contrived any explanation based on the same origin for both the X -- ray and optical fluxes. We here assume that the central engine can be active for a long time, producing shells of decreasing bulk Lorentz factors Γ. We also assume that the internal dissipation of these late shells produces a continuous and smooth emission (power -- law in time), usually dominant in X -- rays and sometimes in the optical. When Γ of the late shells is larger than 1/θj, where θj is the jet opening angle, we see only a portion of the emitting surface. Eventually, Γ becomes smaller than 1/θj, and the entire emitting surface is visible. Thus there is a break in the light curve when Γ = 1/θj, which we associate to the time at which the plateau ends. After the steeply decaying phase which follows the early prompt, we see the sum of two emission components: the "late -- prompt" emission (due to late internal dissipation), and the "real afterglow" emission (due to external shocks). A variety of different optical and X -- ray light curves are then possible, explaining why the X -- ray and the optical light curves often do not track each other (but sometimes do), and often they do not have simultaneous breaks. Subject headings: gamma rays: bursts -- X-rays: general -- radiation mechanisms: general 1. INTRODUCTION One of the puzzling results of the Swift satellite (Gehrels et al., 2004) is the discovery that the [0.3 -- 10 keV] X -- ray light curve of Gamma Ray Bursts (GRBs) is much more complex than thought in the pre -- Swift era. After a steep decline of the flux [F (t) ∝ t−α1 , with α1 ∼ 3 -- 5; Taglia- ferri et al. 2005], which is most commonly interpreted as off axis radiation of a switching -- off fireball (see e.g. Ku- mar & Panaitescu 2000), the flux decay becomes shallow [F (t) ∝ t−α2, with α2 ∼ 0.2 -- 0.8], up to a break time of 103 -- 104 s (Willingale et al. 2007, hereafter W07), after which the flux decays "normally" [F (t) ∝ t−α3 with α3 ∼ in a similar way as ob- 1 -- 1.5; Nousek et al. 2005], i.e. served in the pre -- Swift era. In addition, several bursts show flares superposed to this power law evolution (Bur- rows et al. 2005), leading Fan & Wei (2005) and Lazzati & Perna (2006) to suggest a long lasting central engine. Unpredicted beforehand, the complex structure of the X -- ray light curve, characterized by a steep -- flat -- steep behav- ior, has been interpreted in several ways (for reviews, see e.g. Panaitescu 2007; Granot 2007; Zhang 2007) none of which seems conclusive. The three main possibilities al- ready proposed (but there are more, see the review by Zhang 2007), are: i) energization of the forward shock by the arrival of shells being produced late (with large Γs), or just after the prompt phase (with small Γs); ii) chang- ing microphysical parameters, assuming that the efficiency of the forward shock to produce radiation increases with time: iii) off -- axis jets, whose prompt and early afterglow radiation is not fully beamed towards the observer. Note that the spectral slope does not change across the tem- poral break from the shallow decay phase to the normal decay phase, ruling out the crossing of a spectral break across the band. This favors instead an hydrodynamical or geometrical nature of the break. All these ideas do not obviously distinguish between X -- ray and optical radiation, which should have the same origin. As a consequence, the light curves in both bands should be similar, contrary to what observed in several cases (see e.g. Panaitescu et al. 2006; Panaitescu 2007). These difficulties recently led Uhm & Beloborodov (2007) and Grenet, Daigne & Mochkovitch (2007) to con- sider the possibility that the X -- ray plateau emission is not due to the forward, but to the reverse shock running into ejecta of relatively small Lorentz factors. This however requires an appropriate Γ -- distribution of the ejecta, and also the suppression of the X -- ray flux produced by the forward shock. Here we make the alternative proposal that the plateau phase of the X -- ray emission (and sometimes even of the optical) is due to a "late -- prompt" mechanism: after the early prompt (the prompt which we are used to) there may be a tail of activity of the central engine, producing for a long time (i.e. days) shells of progressively lower power and bulk Lorentz factor. The dissipation process during this and the early phases can occur at similar radii. The to- tal energetics involved in this late activity phase is smaller than (and at most comparable to) the energetics of the early phase, but diluted on a much longer time. The rea- son for the shallow decay phase, and for the break ending it, is that the Γ -- factor is decreasing, allowing to see an increasing portion of the emitting surface, until all of it is visible. 2. LATE PROMPT EMISSION? 1Univ. dell'Insubria, V. Valleggio, 11, I -- 22100, Como, Italy 2Instituto de Astronom´ıa, U.N.A.M., A.P. 70-264, 04510, M´exico, D.F., M´exico 1 2 Assume that the central engine, after having emitted most of the power in the usual duration of what we call "prompt" emission, continues to create shells of much smaller power, but for a much longer time. For simplic- ity, let us call "early prompt" and "late prompt" the two phases of activity. By contrast, we call "real afterglow" the emission produced in the forward shock created by the interaction of the shells with the circumburst medium. The early prompt emission is due to internal dissipa- tion of shells of large Γ -- factors (changing erratically) and energy, due to e.g. internal shocks (Rees & Meszaros, 1994) or interactions with the funnel of the progenitor star (Thompson 2006; Thompson Meszaros & Rees 2006), or some form of magnetic reconnection (e.g. Spruit, Daigne & Drenkhahn 2001). We suggest that the late prompt emission can be due to the same dissipation processes, but by shells created at late times with smaller Γ and much lower power. The ra- diation can then be produced at distances relatively close to the central engine (even less than 1013 -- 1014 cm), in a different region where the shells, produced during the early prompt, interacts with the circumburst medium produc- ing the real afterglow. Note that a smaller Γ implies less Doppler time contraction, and therefore a less pronounced variability during the late prompt. Furthermore, if Γ is decreasing with time, a new effect appears. In fact, when Γ > 1/θj, the emission surface seen by the observer is of the order of (R/Γ)2 (here R is the distance from the black hole where dissipation takes place, and θj is the jet open- ing angle), and becomes (θjR)2 when Γ ≤ 1/θj. In the same way as in the afterglow case, we should then see a steepening of the light curve when the central engine pro- duces shells with Γ ∼ 1/θj. This should occur at the time ta, in a similar way as the jet break time tj for the after- glow. According to this scenario, there is a link between ta and tj: both are times at which Γ = 1/θj, but they refer to two different processes. The plateau phase of the X -- ray emission is characterized by a power law decay L(t) ∝ t−α2, followed by a steeper decay L(t) ∝ t−α3 . The end of the plateau phase occurs at ta, and the transition is smooth. If j ′ is the bolometric emissivity in the comoving frame, we have Γ(cid:19)2 L(t < ta) ∼ π(cid:18) R L(t ≥ ta) ∼ π (θjR)2 ∆R′Γ2j ′(t) ∝ t−α3 , Γ ≤ 1/θj(1) ∆R′Γ2j ′(t) ∝ t−α2 , Γ > 1/θj Assuming that j ′(t) is with constant slope before and after ta, we have t−α3 ∝ Γ2t−α2. Therefore we find Γ ∝ t−∆α/2, with ∆α ≡ α3 − α2. This behavior should be appropri- ate after the time t∗ characterizing the start of the late prompt phase of emission. Finally we have: Γ = Γ∗(cid:18) t t∗(cid:19)−∆α/2 (2) Since hα2i = 0.6 ± 0.3 and hα3i = 1.25 ± 0.25 (Panaitescu 2006), ∆α/2 is of the order of 0.33 ± 0.2. This means that the bulk Lorentz factor, at the beginning of the late 3We use the notation Qx = 10xQ, in cgs units. prompt phase (t∗), is of the order of3 Γ∗ = θ−1 j (cid:18) ta t∗(cid:19)∆α/2 ∼ 46 θj,−1 (cid:18) ta,4 t∗.2(cid:19)1/3 (3) which is smaller than what is usually assumed for the early prompt emission. The barion loading of the late shells can be estimated assuming a given efficiency η for the dissipation process leading to the radiation produced in the plateau phase. After ta, when we see the entire emitting surface, the jet kinetic power per unit solid angle Lkin = Γ M c2 ∼ 2LX/(θ2 j η) and then M = 2L(t > ta) j Γc2 ∝ t−(α2+α3)/2 ηθ2 (4) M ∝ t−1, with a large dis- which approximately gives persion (the dependence is the same also for the plateau phase). The total mass in the late ejecta is relatively small, again at most comparable with what can be estimated for the ejecta of the early prompt. This is because, although the Γ -- factors are smaller, the total energetics of the late prompt shells is smaller than the one of the early prompt. This agrees with the findings of W07 that the total radi- ated energy of the late shells (which is called X -- ray af- terglow in that paper) is on average a factor ∼10 smaller than the early prompt. Therefore, if η of the early and late shells is similar, we do not expect a big effect from the possible refreshed shocks. We also expect that the late shells reach the front forward shock at very different times. This is due both because the late shells are pro- duced at later and later times, and also because the later the shell is produced, the smaller its bulk Lorentz factor, and the longer the time needed to reach the shock front which is decelerating by the interaction with the circum- burst medium. The refreshing effect is long lasting, but diluted. For the same reasons, we expect that the reverse shock is also long lasting, but diluted, and therefore not contributing much to the total flux. To see this, assume for illustration a circumburst density with a wind -- like profile. Calling ti the time (after the trigger) at which a late shell is created, and td the deceleration time, we have that at the observed time t the front shock and the i-th late shell are at the distances Rs and Ri, respectively, given by: Rs = ctΓ2 = ctΓ2 0(cid:18) t td(cid:19)−1/2 = cΓ2 0(ttd)1/2 (5) Ri = c(t − ti)Γ2 i (6) where Γ0 is the initial Lorentz factor of the early shells. Late shells are assumed not to decelerate until they catch up the front shock. Equating the two above radii we have that the ith shell reaches the front shock at the time tc: tc ∼ Γ4 0 Γ4 i td; if tc ≫ ti (7) For td = 100 s, tc ranges from 1600 s (Γ0/Γi = 2) to tc = 106 (Γ0/Γi = 10). Since the decrease of the bulk Lorentz factor is associated with a decrease of the kinetic power, we have some effects only for the first refreshed shocks, at times of the order of thousands of seconds, but not later. 2.1. The real afterglow In the pre -- Swift era, it was generally believed that the real afterglow should contain at least a comparable amount of energy (in the emitted radiation) of the prompt, that it should start a few tens -- hundreds of seconds after the trigger, and that the energy band containing most of the emission is the X -- ray band (since its energy spectrum F (ν) ∝ ν−1 indicates that the peak in νF (ν) is within or close the X -- ray band). The prediction of the start time of the afterglow seems well confirmed in two cases: GRB 060418 and GRB 060607, as shown in Molinari et al. (2007), thanks to ground based near IR observation by the REM telescope. Quite remarkably, the X -- ray light curve in these two cases does not track the near IR, confirming that, although the afterglow theory can correctly explain what seen in the optical, we need another component to explain the X -- ray i) the real afterglow X -- ray flux. We also conclude that: component is much weaker than thought before: ii) the X -- ray band is likely not the band where most of the af- terglow energy is, and iii) the total energetics of the real afterglow is much smaller than thought before. A weak afterglow can result if the microphysical param- eters ǫe, ǫB, at least for the first afterglow phases, are much smaller than commonly thought. Alternatively, the fireball can have a small kinetic energy, as a result of a very efficient prompt phase, that was able to convert a large fraction of the fireball energy into radiation. Further- more, the radiation produced by the real afterglow should be mostly in the IR -- optical, not in the X -- ray band. There is a spectral transition between the varying and generally hard slope of the hard X -- ray emission and a more stable and generally softer slope of the later X -- ray emis- sion. In W07, the distributions of the spectral index βx of the prompt and the plateau phase are broad, but a slight narrowing around a value βx ∼ 1 for the plateau is visible. We propose that this is not due to the prompt/afterglow transition, but it is instead associated to the transition between the early prompt phase, characterized by large Γ -- factors changing erratically, and the late prompt phase characterized by smaller Γ decreasing monotonically. Note also that the pre -- Swift observations of the X -- ray "afterglow" should be re -- interpreted: in many cases, what observed even days after the trigger time should be late prompt, not real afterglow, emission. 2.2. Flares The flares occurring mainly in X -- ray (Burrows et al. 2007) but sometimes also in the optical light curve have been associated to internal dissipation (internal shocks) either by late shells, or by shells produced within the first early prompt phase, but moving with a small Lorentz fac- tor. In our framework, the most likely possibility is that a flare is produced by a late shell, moving with a somewhat larger Lorentz factor than the shells created just earlier. Thus there will be a chain of interactions between this (faster -- than -- average) shell and the slower previous ones, and this mechanism can be efficient in converting the ki- netic energy of the shell into radiation. Due to the differ- ent time Doppler contraction, late flares should also last longer than early ones. Alternatively, the flares could flag periods of enhanced activity of the central engine, able for 3 some time (i.e. tens or hundreds of seconds) to produce shells (or a continuous flow) of higher energy. 3. DISCUSSION The scenario here proposed allows to explain some GRB properties which are puzzling and mysterious. In general, the X -- ray flux can receive contributions from the steep part of the early prompt phase, the late prompt emission, and from the decelerating early shells which are producing the real afterglow emission. The late prompt emission, in the optical -- UV bands, may be reduced if the main radiation process is multiple Comptonization of UV/soft X -- ray seed photons, or by self -- absorption if it is synchrotron radiation. The real afterglow emission, in- stead, should produce synchrotron (and self -- Compton) ra- diation both in the X -- ray and in the optical bands. Fo- cusing to the plateau and later phases, we may have a complex behavior: 1. the X -- ray flux is dominated by late prompt emission, while the optical is dominated by the real afterglow. In this case the light curves in the two bands are independent, and show no simultaneous break. In particular, the jet break time is possibly seen in the optical, but not in the X -- rays. 2. Both the X -- ray and the optical fluxes are domi- nated by late prompt emission. In this case the two light curves are similar, they may have simultane- ous breaks (at the time ta) and the jet break time (due to real afterglow emission) can be masked. The dominance of the late prompt emission may however end after some time, beyond which the real afterglow can become visible. 3. Both the X -- ray and the optical fluxes are dominated by real afterglow emission. This is the case foreseen before Swift. The light -- curves should have an achro- matic jet -- break, should track one another, and they should not show the break at ta. The first case is sketched in Fig. 1. While the other two cases can also occur, the case of a late prompt emission dominating in the optical but not in the X -- rays seems contrived. In terms of total energetics, the scenario proposed here may be the least demanding for explaining what observed. In fact, in the refreshed shock scenario, the plateau phase is flat because the shock running into the circumburst medium is energized by the arrival of shells with kinetic en- ergies which largely overtake the energy of the first shells, which have contributed to the early prompt emission. Al- ternatively, in the increasing ǫe, ǫB scenario, the radia- tion produced during the plateau phase is a very tiny frac- tion of the carried kinetic energy. Instead, interpreting the plateau phase as late prompt emission, we need that the extra energy created by the central engine in the late phase is less than (or at most comparable to) the total energetics of the shells responsible for the early prompt emission. If the radiative efficiency of the early prompt is large, we can also explain the weakness of the real afterglow emission, since the kinetic energy of the fireball, remaining after the early prompt phase, may be relatively small. 4 Our scenario is not based on a detailed model on how the central engine works. The following ideas should then be considered as speculations to be studied in future work. After the black hole formation following the core collapse of the progenitor star, the equatorial core material which failed to form the black hole in the first place can form a very dense accreting torus, which can sustain a strong magnetic field, which in turn extracts the rotational energy of the black hole. This accretion phase could correspond to the early prompt phase of the burst. After this phase, some fall -- back material may also be accreted. This phase of "late accretion" can last for a longer time, with a density of the accreting matter smaller than in the early phases. If so, the magnetic field that this matter can sustain is weaker than before, with a corresponding smaller power extracted from the black hole spin. This may well corre- spond to production of shells of smaller Γ -- factors. These shells can dissipate part of their energy with the same mechanism of the early ones. Occasionally, the central en- gine produces a faster than average shell, originating the late flares often observed in the Swift/XRT light curves. Our suggestion may be not the unique solution of the puzzle concerning the unpredicted behavior of the X -- ray and the optical light curves. Indeed, Uhm & Beloborodov (2007) and Grenet, Daigne & Mochkovitch (2007) recently proposed that the X -- ray flux may be dominated by the re- verse shock emission in slow shells. These and our propos- als share the common view that the X -- ray flux can be due to a component different from what produces the optical. The difference is that in our model the late prompt and real afterglow emissions are completely decoupled, while in the reverse shock scenario the two emission processes are linked. Furthermore, in the reverse shock scenario, one has to assume a somewhat ad hoc time profile of the Γ -- factor to explain the flat -- steep X -- ray transition, (Uhm and Beloborodov 2007), which is not a simple (unbroken) power law as in our case. Finally, there are some features that our model can pre- dict. Observationally, we should have the three cases men- tioned above: both the optical and the X -- rays are late prompt emission; both are real afterglow emission; X -- rays and optical are "decoupled", with the X -- ray due to late prompt and the optical due to real afterglow emission, re- spectively. One obvious way to check these possibilities is through the construction of the simultaneous spectral energy distribution (SED), which can confirm or not if the X -- ray and the IR -- optical fluxes belong to the same com- ponent. The unknown extinction due to the host galaxy material may complicate this test, but having enough pho- tometric data, especially in the infrared, may result in a good determination of the extinction, and thus a good esti- mate of the extrapolation of the IR -- optical spectrum into the X -- ray range. The SED so obtained may clearly show that the IR -- optical and X -- ray emission belong (or not) to two different components. Another test concerns the total kinetic energy of the fireball after its radiative phase, using the radio data, as done e.g. for GRB 970508 by Frail, Waxman & Kulka- rni (2000). Should the derived energetics be smaller than what required by the refreshed shock scenario, one could exclude this possibility, and instead favor our scenario. In cases in which the late prompt emission ends, the underlying real afterglow emission can be revealed. In the light curve, this should appear as a steep -- flat transition at late times (not to be confused with the usual steep -- flat -- steep X -- ray decay). This can also be confirmed by the corresponding SEDs. We thank the anonymous referee for helpful comments, F. Tavecchio for discussion and a 2005 PRIN -- INAF grant for funding. REFERENCES Burrows, D.N., Romano, P., Falcone, A., et al., 2005, Science, 309, Burrows, D.N., Falcone, A., Chincarini, G., et al., 2007, preprint (astro -- ph/0701046) Campana, S., Mangano, V., Blustin, A.J., et al., Nature, 2006, 442, 1833 1008 Fan, Y.Z. & Wei, D.M., 2005, MNRAS, 364, L42 Frail, D.A., Waxman, E. & Kulkarni, S.R., 2000, ApJ, 537, 191 Gehrels, N., Chincarini, G., Giommi, P., et al., 2004, ApJ, 611, 1005 Granot, J., 2007, Il Nuovo Cimento, in press (astro -- ph/0612516) Grenet, F., Daigne, F. & Mochkovitch, R., 2007, subm. to MNRAS (astro -- ph/0701204) Kumar, P. & Panaitescu, A., 2000, ApJ, 647, 1213 Lazzati, D. & Perna, R., 2007, MNRAS, 375, L46 Liang, E. & Zhang, B., 2005, ApJ, 633, L611 Molinari, E., Vergani, S., Malesani, D., et al., 2007, subm to Science (astro -- ph/0612607) Nousek, J.A., Kouveliotou, C., Grupe, D., et al., 2005, ApJ, 642, 389 Panaitescu, A., Meszaros, P., Burrows, D., Nousek, N., O'Brien, P. & Willingale R., 2006, MNRAS, 369, 2059 Panaitescu, A., 2007, Il Nuovo Cimento, in press (astro -- ph/0607396) Rees, M.J. & Meszaros, P., 1994, ApJ, 430, L93 Spruit, H.C., Daigne, F. & Drenkhahn, G., 2001, A&A, 369, 694 Tagliaferri, G., Goad, M., Chincarini, G., et al., 2005, Nature, 436, 985 Thompson, C., 2006, ApJ, 651, 333 Thompson, C., Meszaros, P. & Rees, M.J., 2007, subm to ApJ, (astro -- ph/0608282) Uhm, L.Z. & Beloborodov, A.M., 2007, subm to ApJ, (astr -- ph/0701205) Willingale, R., O'Brien, P.T., Osborne, J.P., et al., 2007, subm to ApJ (W07) (astro -- ph/0612031) Zhang, B., (astro -- ph/0611774) 2007, Advances in Space Research, in press 5 Fig. 1. -- Schematic illustration of the different components contributing to the X -- ray and optical light curves, as labelled. Scales are arbitrary. The case illustrated here is only one (likely the most common) possible case (see text), when the X -- ray flux is dominated by late prompt emission (solid line, the dotted line corresponds to an extrapolation at very late times), while the optical flux is dominated by the real afterglow (dashed). ΓLP and ΓF S indicate the Γ of the late shells and the forward shocks, respectively.
astro-ph/0507456
2
0507
2005-08-04T21:46:48
The Type Ia Supernova Rate
[ "astro-ph" ]
We explore the idea that the Type Ia supernovae (SNe Ia) rate consists of two components: a prompt piece that is proportional to the star formation rate (SFR) and an extended piece that is proportional to the total stellar mass. We fit the parameters of this model to the local observations of Mannucci and collaborators and then study its impact on three important problems. On cosmic scales, the model reproduces the observed SNe Ia rate density below z=1, and predicts that it will track the measured SFR density at higher redshift, reaching a value of 1-3.5 X 10^-4 per yr per Mpc^3 at z=2. In galaxy clusters, a large prompt contribution helps explain the iron content of the intracluster medium. Within the Galaxy, the model reproduces the observed stellar [O/Fe] abundance ratios if we allow a short (approximately 0.7 Gyr) delay in the prompt component. Ongoing medium-redshift SN surveys will yield more accurate parameters for our model
astro-ph
astro-ph
Preprint typeset using LATEX style emulateapj v. 6/22/04 THE TYPE IA SUPERNOVA RATE Evan Scannapieco & Lars Bildsten Kavli Institute for Theoretical Physics, Kohn Hall, University of California, Santa Barbara, CA 93106; [email protected], [email protected] ABSTRACT We explore the idea that the Type Ia supernovae (SN Ia) rate consists of two components: a prompt piece that is proportional to the star formation rate (SFR) and an extended piece that is proportional to the total stellar mass. We fit the parameters of this model to the local observations by Mannucci and collaborators and then study its impact on three important problems. On cosmic scales, the model reproduces the observed SN Ia rate density below z = 1, and predicts that it will track the measured SFR density at higher redshift, reaching a value of 1-3.5 × 10−4 yr−1 Mpc−3 at z = 2. In galaxy clusters, a large prompt contribution helps explain the Fe content of the intracluster medium. Within the Galaxy, the model reproduces the observed stellar [O/Fe] abundance ratios if we allow a short (≈ 0.7 Gyr) delay in the prompt component. Ongoing medium-redshift SN surveys will yield more accurate parameters for our model. Subject headings: supernovae: general -- galaxies:evolution -- galaxies: clusters: general -- stars: abun- dances 1. INTRODUCTION 2004). Type Ia supernovae (SNe Ia) play a pivotal role in as- trophysics. On cosmological scales they serve as unparal- leled distance indicators, providing direct evidence that the low-redshift universe is accelerating (Riess et al. 1998; Perlmutter et al. 1999). On galactic scales, they act as the primary source of iron, producing ≈ 0.7 M⊙ per event (Tsujimoto et al. 1995), roughly an order of magnitude more than in core-collapse SNe (Hamuy 2005). On stellar scales, they represent an excellent example of explosive nuclear burning, resulting in radioactively-powered light curves (e.g., Nomoto et al. 1984; Woosley 1990). Nevertheless, many mysteries remain. While the peak magnitude and the decay time of SN Ia light curves are tightly correlated (Pskovskii 1977; Phillips 1993; Hamuy et al. 1995), the origin of this relation is poorly under- stood (see discussion in Pinto & Eastman 2000; Mazzali et al 2001). Similarly, while there is a consensus that SNe Ia originate from thermonuclear ignition and burn- ing of a C/O white dwarf in a binary system, it is un- certain whether they are triggered by accretion from a hydrogen-rich companion or from a merger with another white dwarf (see Branch et al. 1995). This uncertainty in the progenitor makes it difficult to predict the Ia rate in galaxies of varying masses, ages, and star formation rates. Consequently, a wide range of models of this rate have been developed (e.g., Matteucci & Recchi 2001; Greggio 2005). These are commonly parameterized by a delay function, whose convolution with the star formation rate (SFR) yields the SN Ia rate. In principle, this is a com- pletely general approach, as the SN Ia rate must depend on the mass and age of the underlying stars. In prac- tice, most of this generality is lost to the assumption of a single "delay time" (e.g., Madau et al. 1998; Dahl´en & Fransson 1999; Gal-Yam & Maoz 2004; Strolger et al. 2004). These fits are used to draw conclusions about SN Ia progenitors, the most recent example of which is the claim that there must be a 2-4 Gyr delay in all SNe Ia relative to the burst of star formation (Strolger et al. However, this approach neglects the possibility that multiple evolutionary paths lead to SNe Ia. Indeed, there is direct observational evidence that this is the case. The brightest events (such as 1991T) only occur in ac- tively star-forming galaxies, while substantially under- luminous events (such as 1991bg) are most prevalent in E/S0 galaxies (Hamuy et al. 1996; Howell 2001; van den Bergh et al. 2005). This is an important clue that SNe Ia have at least two evolutionary channels with different characteristic times: one "prompt," basically tracking the current SFR, and another so delayed that it simply scales with the stellar mass (much as is seen in accreting binaries with low-mass companions, such as Cataclysmic Variables [Townsley and Bildsten 2005] or low-mass X- ray binaries in E/S0s [Gilfanov 2004]). We show here, that in addition to explaining the SN Ia rates seen in nearby galaxies as described by Mannucci et al. (2005, hereafter M05), such a simple two-component model also resolves a wide range of outstanding issues. We begin in §2 by presenting the model, fitting the con- stants to observations of SNe Ia in nearby galaxies, and stating a few of the immediate repercussions. In §3 we study the implications of this model in three important contexts: the iron content of galaxy clusters, the evolu- tion of the average cosmic SN Ia rate density, and the [O/Fe] abundance ratios of Galactic stars. In §4 we con- trast this approach with other models, and we conclude in §5. 2. THE TWO-COMPONENT MODEL Following M05, we assume that there are two avenues for SNe Ia. Specifically, we adopt a model in which the SN Ia rate is the sum of two components: a term propor- tional to the total stellar mass, M⋆(t) (regardless of its age) and a term proportional to the instantaneous SFR, M⋆(t), SN RIa(t) (100 yr)−1 = A(cid:20) M⋆(t) 1010M⊙(cid:21) + B" M⋆(t) 1010M⊙ Gyr−1# , (1) 2 where A and B are dimensionless constants that we fix with observations (see also M05, eq. [2]). The first of these terms is dominant in old stellar populations (and contains underluminous 1991bg-like SNe), while the sec- ond term is most important in starbursts (and contains the brightest 1991T-like SNe). To measure A, we use the recent observations from M05, who utilized detailed K-band data to update the analysis presented by Cappel- laro et al. (1999). As there were 21 Type Ia SNe observed in E/S0 galaxies in this sample, and no instances of core- collapse supernovae, we consider the SFR term to be neg- ligible in this population. This gives A = 4.4+1.6 −1.4 × 10−2. M05 showed that the SN Ia rate was 0.35 ± 0.08 of the core collapse rate in young stellar populations. Since they arise from massive, short-lived stars, the core- collapse SN rate should directly trace the SFR, and thus can be used to determine B. Presently, the pri- mary uncertainty in this measurement comes from re- lating the core-collapse rate to the SFR. We take two approaches to determining B, fully aware that ongoing SN surveys will soon reduce these uncertainties. First we use the z ≤ 1.0 core-collapse SN rate density as mea- sured by Dahl´en et al. (2004), comparing it against the SFR density as measured by Giavalisco et al. (2004), and considering only the statistical error bars. This gives SN Rcc/ M⋆ = (7.5 ± 2.5) × 10−3 M⊙ −1 and corresponds to a B value for SNe Ia of 2.6 ± 1.1, which we adopt throughout this Letter. An alternative approach is to use the blue (B − K ≤ 2.6) population observed by M05, in which the measured SN Ia rate is 0.86+0.45 −0.35 per 100 yr per 1010 M⊙ of stars. The colors of these starbursting galaxies are consistent with a 0.7 Gyr population (M05), as computed from the population synthesis models of Bruzual & Charlot (2003). Within the errors, this gives B = 1.2+0.7 −0.6, con- sistent with our first estimate. Our values of A and B directly yield the relative contri- butions of these two components. For example, 10 Gyrs after a starburst, only 20% of all SNe Ia will have come from the extended piece. Hence, in our model, most of the SNe Ia over any galaxy's lifetime come from the prompt contribution as originally suggested by Oemler & Tinsley (1979). Our model also allows for a comparison between SN types, as illustrated in Figure 1. In our model, by construction, the rate of core-collapse SNe is approxi- mately 3 times the SN Ia rate in starbursting galaxies, while only SNe Ia are found in galaxies without star formation. For a galaxy with a total stellar mass of 1010 M⊙ the transition between these two regimes oc- M⋆ ≈ 0.1 M⊙ yr−1, which corresponds to a Scalo curs at parameter b ≡ M⋆(t)/D M⋆(t)E = M⋆(t)t/M⋆(t) ≈ 0.08 or an age of 7 Gyrs if we assume a star formation rate ∝ e−t/2Gyr as in M05. Finally, since 0.74 M⊙ of Fe is expelled in a typical SN Ia (Tsujimoto et al. 1995) while 0.062 M⊙ is expelled in a typical core-collapse SN (Hamuy 2003), in any given starburst the overall SN Ia contribution to Fe produc- tion as compared with core-collapse SNe in our model is approximately 3 to 1. 3. IMPLICATIONS AND PREDICTIONS Fig. 1. -- The supernova rate in a galaxy with a final stellar mass of 1010M⊙. The solid line gives our model predictions for the Type-Ia SN rate (bracketed by 1 sigma errors), which is made up of the prompt (dotted) and extended (dot-dashed) components. The dashed line gives the core-collapse SN rate. In all cases we assume a star formation rate ∝ e−t/2Gyr. Choosing a different characteristic star-formation decay time would rescale the time axis, while leaving the Scalo b values unchanged. We now apply this simple model to three important issues. In this section and below we adopt a Hubble con- stant of 70 km s−1 Mpc−1 and total matter and vacuum energy densities of Ωm = 0.3 and ΩΛ = 0.7 in units of the critical density (e.g., Spergel et al. 2003). First we consider the intracluster medium (ICM) in galaxy clus- ters, which is measured to have an Fe content ≈ 0.3Z⊙ (Baumgartner et al. 2003). Although clusters are dom- inated by elliptical galaxies, models that combine the observed SN Ia rate in ellipticals with the total cluster stellar mass result in Fe estimates that are roughly an order of magnitude too small (e.g., Renzini et al. 1993; Renzini 2004). While Maoz & Gal-Yam (2004) were able to provide a single-component resolution to this prob- lem, they were not able to reconcile this fit with SN Ia measurements in field galaxies. In fact, even ICM models that appeal to Fe production by pair-instability SNe from very massive primordial stars (e.g., Lowenstein 2001) fall far short of the observed metallicity (Scannapieco et al. 2003). Our two-component model, however, addresses this is- sue from a different perspective. The dominant source of Fe is not the late-time SN Ia contribution, as observed in ellipticals, but rather the prompt contribution, which took place at high redshifts. In the upper panel of Figure 2, we plot the ICM metallicity as a function of redshift, assuming a star formation redshift of 3, and an ICM to stellar mass ratio of MICM/M⋆ = 10 ± 3, as appropri- ate for 7 keV clusters (Lin et al. 2003). Again, we take the core-collapse SN rate to directly trace the SFR. This results in [Fe/H] values broadly consistent with obser- vations and an order of magnitude higher than previous estimates (Renzini 2004). Our model also naturally pre- dicts the recently observed lack of [Fe/H] evolution with redshift (Tozzi et al. 2003), a feature that does not ap- 3 Fig. 2. -- Top: [Fe/H] of the intracluster medium in kT ≥ 5 keV galaxy clusters as a function of redshift. The solid line shows the results of our two-component model, with the one-sigma errors defining the shaded region. The low redshift point (open triangle) is an average from the Baumgartner et al. (2003) sample, while the higher redshift points (solid circles) are from Tozzi et al. (2003). Bottom: Type Ia SNR density as a function of redshift (solid) which is the sum of the prompt (dotted) and extended (dot-dashed) com- ponents. Again the solid line corresponds to our two-component model, with the one-sigma errors given by the shaded region. The lower redshift measurements (open triangles) are taken (in order of increasing redshift) from Cappellaro et al. (1999), Hardin et al. (2000), Blanc et al. (2004), Reiss (2000), Pain et al. (1996), Tonry et al. (2003), and Pain et al. (2002). The higher redshift measure- ments (solid circles) are from Dahl´en et al. (2004). Fig. 3. -- Top: Evolution of the Fe content in a closed-box system as a function of the age of the stellar population, assuming an SFR ∝ e−t/2Gyr. Here the (thin) upper solid line corresponds to our fiducial two-component model (bracketed by the 1-sigma errors), while the dashed line gives the iron provided by core-collapse SNe. Finally, the (thick) lower sold line is the result of our delayed two- component model. Bottom: Variation of the [O/Fe] abundance ratio as a function of metallicity, in a simple closed box model. The (thin) lower line gives the results of our fiducial model, while the upper (thick) line corresponds to the two-component model with a 0.7 Gyr delay in the prompt component. The data points are taken from observations of Galactic disk and halo stars compiled in McWilliam (1997), following the symbol convention used in his Figure 3. pear in models dominated by late-time SNe Ia. Next we turn to the cosmic SN Ia rate density, or the number of Type Ia SNe per year per comoving Mpc3. We adopt a cosmic star formation rate density of log10[SFR/( M⊙ yr−1 Mpc−3)] = −2.2 + 3.9 log10(1 + z) − 3.0[log10(1 + z)]2, which is a simple fit to the most recent measurements (Giavalisco et al. 2004; Bouwens et al. 2004). The resulting SN Ia rate density is compared with observations in the lower panel of Figure 2, provid- ing an excellent match, except for the highest-redshift point from Dahl´en et al. (2004). This is because at z ≈ 1 our model is dominated by the prompt piece, and no cor- responding dip is seen in the SFR density at this redshift. Furthermore, requiring agreement with this point is the source of the 2 − 4 Gyr delay-time derived by Strolger et al. (2004). Thus a strong prediction of our model is that future observations will revise the z ≈ 1.5 measurement upward. In fact, more detailed analyses of the SN Ia rate density at z > 1 represent the single best way to falsify (or confirm) our approach. Finally we turn to the measured abundance ratios of Galactic stars, which probe the relationship between Type Ia and core-collapse SNe at short times. In particu- lar we compare iron with oxygen, an alpha-element that is synthesized primarily in core-collapse SNe. We take 0.14 M⊙ of oxygen per SN Ia (Tsujimoto et al. 1995), 1.2 M⊙ oxygen per core-collapse SN (a Salpeter inital mass function average computed in Scannapieco et al. 2003), and a closed-box model. Following M05 we adopt M⋆ = Mgas exp[−t/2Gyr]/2Gyr, a star formation rate which results in the [Fe/H] and [O/Fe] values shown in Figure 3. As the overall Ia Fe contribution as compared with core-collapse SNe in our model is nearly 3 to 1, [O/Fe] should drop by a factor of 3, as observed. The value of [Fe/H] where this drop occurs, however, depends on the star formation model, and requires a short delay in the prompt component. Assuming that both the core- collapse and the prompt Type Ia contributions exactly trace the SFR would fix the [O/Fe] values to a constant, in conflict with measurements of metal-poor halo stars. On the other hand, delaying the prompt component by 0.7 Gyr allows core-collapse SNe to briefly dominate the initial gas enrichment, but has no impact on the SN Ia distribution on the timescales probed by other measure- ments. The timescale of this delay is proportional to the assumed SFR decay time, and a model with an SFR de- cay time of 1.0 Gyr and a prompt SN Ia delay of 0.35 Gyr would give equivalent [O/Fe] values. In any case, this de- lay is so short on cosmic times that we will continue to refer to this component as "prompt." 4. COMPARISON WITH OTHER WORK 4 We now compare our approach with other single- In particular we consider three pos- component fits. sible SN Ia delay functions: an exponential model in which Φ(∆t) = C exp(−∆t/τ )/τ (Tutukov & Yun- gelson 1994; Madau et al. 1998; Gal-Yam & Maoz 2004), and two Gaussian models in which Φ(∆t) = C(2πσ2)−1/2 exp[(∆t − τ )2/(2σ2)] and σ is either "nar- row" (σ = 0.2τ ) or "wide" (σ = 0.5τ ) (Dahl´en & Frans- son 1999; Strolger et al. 2004). Each of these functions has two free parameters: a delay time τ and a normal- ization C; which we fit to the number of SNe Ia in the youngest (B − K < 2.6) and most evolved (B − K > 4.1) galaxies measured by M05. As in that study, we model both galaxy types with M⋆ ∝ exp(−t/2Gyr) with an age of 0.75 ± 0.25 Gyrs in the B − K < 2.6 population and 10.5 ± 1.5 Gyrs in the B − K > 4.1 population, as is consistent with their observed colors and core-collapse SNe rates. Note that these SFRs are averages over entire populations, rather than histories of individual galaxies. Within the one-σ errors, we find τ values of 0.5 − 1.6, 0.6 − 1.0, and 0.5 − 1.3 Gyr, for the exponential, narrow Gaussian, and wide Gaussian models, respectively. Applying these fits to the full range of galaxy popula- tions measured by M05 indicates that the proper number of SNe Ia in old galaxies and starbursting galaxies is ob- tained only at the expense of a large number of SNe Ia in galaxies of intermediate age. For example in a t = 5 Gyr population, all three models predict ≥ 2 SNe Ia per 100 yr per 1010 M⊙, while the measured value is 0.19+0.08 −0.07 (M05). Furthermore the ratio of Type Ia to core-collapse SNe at 5 Gyr is ≥ 4, while the measured ratios are ≤ 1/3. On the other hand, our two-component model, shown as the solid line in Figure 1, falls within the range of ob- served values at all ages. 5. CONCLUSIONS Our two-component model is motivated by the ob- served dichotomy between the environments of the brightest (1991T-like) and the faintest (1991bg-like) SNe Ia. Yet in some sense it is simply an application of the delay function formalism, in which the SN Ia rate is de- scribed as a convolution of an unknown function with the overall star formation history. Other models were limited by the assumption of a single "delay time" and had difficulties in reconciling the Fe content in clusters with the ratio of core-collapse to Type Ia SNe as a func- tion of galaxy age (e.g., Maoz & Gal-Yam 2004). Our model solves this problem because it is dominated by a prompt component, but allows significant numbers of SNe Ia to occur at late times. In addition, it produces the observed cosmic SN Ia rate to z ≤ 1, while also fitting observations of E/S0s. The strongest test of this model is the measurement of the SN Ia rate at z > 1, which we predict to be in the range 1 − 3.5 × 10−4 yr−1 Mpc−3 at z = 2. M⋆/D M⋆E & 0.1. An example Of course, star formation does occur in some ellipti- cal galaxies, and as shown in Figure 1, our simple model predicts that prompt SNe Ia will be the dominant com- ponent in such objects if of such a case is the slow-declining Type Ia SN 1998es, which occurred in the early-type galaxy NGC632. While this galaxy is fairly red (B-K > 3), spectral observations uncover significant star formation (Gallagher et al. 2005). Conversely, underluminous SNe Ia should occasionally be found in star forming galaxies, such as the rapidly- declining SN 1999by. While the host galaxy of this SN Ia is an Sb galaxy, imaging shows that it took place in the old population of halo stars (Gallagher et al. 2005). SNe Ia play a pivotal role in astrophysics, and thus our two-component model has many implications. It allows for an updated assessment of Clayton & Silk's (1969) hy- pothesis that the γ-rays from radioactive decays explain the extragalactic MeV background (see Watanabe et al. 1999; Ruiz-Lapuente et al. 2001; Ahn et al. 2005). It highlights the usefulness of measurements that constrain Type Ia evolution at short time scales, such as studies of the distribution of SNe Ia relative to spiral arms (Maza & van den Bergh 1976; Bartunov et al. 1994; McMil- lan & Ciardullo 1996; Petrosian et al. 2005) and the Fe content of high-redshift quasars (Barth et al. 2003; Di- etrich et al. 2003). It stresses the importance of early SNe Ia in ICM enrichment and exposes the limitations of one-component fits. More generally, our model implies that over a Hubble time, ≈ 80% of the SNe Ia from any galaxy will occur within a Gyr of the initial starburst. The remaining 20% occur in a delayed fashion, clearly extending to times > ∼ 10 Gyrs. This alludes to multiple progenitor scenar- ios: one that occurs "promptly," within a Gyr, and an- other that can occur a Hubble time after star formation. Perhaps this is not surprising given the openly-debated range of possibilities (e.g., Branch et al. 1995; Greggio 2005) and the evidence for dominance (or absence) of some extreme SNe Ia in certain galaxy types. Deeper physical insights into how the age or metallicity of the accreting white dwarf might naturally cause this large range of diversity awaits theoretical work. We thank Avishay Gal-Yam and the anonymous ref- eree for comments. This work was supported by the NSF under grants PHY99-07949 and AST02-05956. REFERENCES Ahn, K., Komatsu, E. & Hoflich, P. 2005, PRD, in press Branch, D., Livio, M., Yungelson, L. R., Boffi, F. R., & Baron, E. (astro-ph/0506126) Barth, A. J., Martini, P., Nelson, C. H., & Ho, L. C 2003, ApJ, 593, L95 Bartunov, O. S., Tsvetkov, D. Yu., Filimonova, I. V. 1994, PASP, 106, 1276 Baumgartner, W. H., Loewenstein, M., Honer, D. J., & Mushotzky, R. F. 2005, ApJ, 620, 680 Blanc, G. et al. 2004, A&A, 423, 881 Bouwens, R. et al. 2004, ApJ, 616, L79 1995, PASP, 107, 1019 Bruzual, G. & Charlot. S. 2003, MNRAS, 344, 1000 Cappellaro, E., Evans, R., & Turatto, M. 1999, A&A, 351, 459 Clayton, D. D. & Silk, J. 1969, ApJ, 158, L43 Dahl´en, T., & Fransson C. 1999, A&A, 351, 459 Dahl´en, T. et al. 2004, ApJ, 613, 189 Dietrich, M., Hamann, F., Appenzeller, I., & Vestergaard, M. 2003, ApJ, 596, 817 Gal-Yam, A., & Maoz, D. 2004, MNRAS, 347, 942 Gallagher, J. S., Garnavich, P. M., Berlind, P., Challis, P., Jha, S., & Kirshner, R. P. 2005, ApJ, in press Giavalisco, M. 2004, ApJ, 600, L103 Gilfanov, M. 2004, MNRAS, 349, 146 Greggio, L 2005, A&A submitted, (astro-ph/0504376) Hamuy, M., Phillips, M. M., Maza, J., Suntzeff, N. B., Schommerr, R. A., & Aviles, R. 1995, AJ, 109, 1 Hamuy, M., Phillips, M. M., Suntzeff, N. B., Schommerr, R. A. Maza, J., & Aviles, R. 1996, AJ, 112, 23910 Hamuy, M. 2005 in Core Collapse of Massive Stars, ed. C. L. Fryer (Kluwer: Dordrecht), in press (astro-ph/0301006) Hardin, D., et al. 2000, A&A, 362, 419 Howell, D. A. 2001, ApJ, 554, L193 Lin, Y.-T., Mohr, J. J., & Stanford, S. A. 2003, ApJ, 591, 749 Lowenstein, M. 2001, ApJ, 557, 573 Mannucci, F. et al. 2005, A& A, 433, 807 (M05) Madau, P., della Valle, M., & Nino, P. 1998, MNRAS, 297, L17 Maoz, D. & Gal-Yam, A. 2004, MNRAS, 347, 951 Matteucci, F., & Recchi, S. 2001, ApJ, 558, 351 Maza, J., & van den Bergh, S. 1976, ApJ, 204, 519 Mazzali, P. A., Nomoto, K., Cappellaro, E., Nakamura, T., Umeda, H., & Iwamoto, K. 2001, ApJ, 547, 988 McMillan, R. J., & Ciardullo, R. 1996 ApJ, 473, 707 McWilliam, A 1997, 35, 503 Nomoto, K., Thielemann, F.-K., Yokoi, K. 1984, ApJ, 286, 664 Oemler, A. & Tinsley, B. M. 1979, AJ, 84, 985 Pain, R., et al. 1996, ApJ, 473, 356 Pain, R, et al. 2002, ApJ, 577, 120 Perlmutter, S. et al. 1999, ApJ, 517, 565 Petrosian, A. et al. 2005, 129, 1369 Phillips, M. M. 1993, ApJ, 413, L105 5 Pinto, P. A., & Eastman, R.G. 2000, ApJ, 530, 744 Pskovskii, Y. P. 1977, Soviet Astron., 21, 675 Reiss, D. J. 2000, Ph.D. thesis, Univ. Washington Renzini, A, Ciotti, L., D'Ercole, A., & Pellegrini, S. 1993, ApJ, 419, 52 Renzini, A. 2004, in in Carnegie Obs. Astrophys. Ser., 3, Clusters of Galaxies: Probes of Cosmological Structure and Galaxy Evolution, ed. J. S. Mulcaey, A. Dressler, & A. Oemler (Cambridge: Cambridge Univ. Press) Riess, A. G. et al. 1998, AJ, 116, 1009 Ruiz-Lapuente, P., Casse, M. & Vangioni-Flam, E. 2001, ApJ, 549, 483 Scannapieco, E., Schneider, R., & Ferrara, A. 2003, ApJ, 589, 35 Spergel, D. N. et al. 2003, ApJS, 14, 175 Strolger, L.-G. et al. 2004, ApJ, 613, 200 Tonry, J. L., et al. 2003, ApJ, 594, 1 Townsley, D. M., & Bildsten, L. 2005, ApJ, 628, 395 Tozzi, P., Rosati, P., Ettori, S., Borgaini, S., Mainieri, V., & Norman, C. 2003, ApJ, 593. 705 Tsujimoto, T., Nomoto, K., Yoshii, Y., Hashimoto, M., Yanagida, S., & Thielemann, F.-K. 1995, MNRAS, 277, 945 Tutukov, A. V. & Yungelson, L. R. 2004, MNRAS, 268, 871 van den Bergh, S., Li, W., & Filippenko, A. V. 2005, PASP, in press (astro-ph/0504668) Watanabe, K., Hartmann, D. H., Leising, M. D., & The, L.-S., 1999, ApJ, 516, 285 Woosley, S. E. 1990, in Supernovae, ed. A. G. Petscheck (New York: Springer-Verlag), 182
astro-ph/0211643
1
0211
2002-11-29T07:58:39
The CYDER Survey: First Results
[ "astro-ph" ]
We present the Calan-Yale Deep Extragalactic Research (CYDER) Survey. The broad goals of the survey are the study of stellar populations, the star formation history of the universe and the formation and evolution of galaxies. The fields studied include Chandra deep pointings in order to characterize the X-ray faint populations. Here we present the results on the first fields studied. We find that the redshift distribution is consistent with that found in the Chandra Deep Field North. The distribution of hardness ratios is, however, softer in our sample. We find a high redshift quasar, CXOCY J125304.0-090737 at z=4.179, which suggests that the abundance of low luminosity high redshift quasars may be larger than what would be expected from reasonable extrapolations from the quasar optical luminosity function.
astro-ph
astro-ph
Astron. Nachr./AN 32X (200X) X, XXX -- XXX The CYDER Survey: First Results 2 0 0 2 v o N 9 2 1 v 3 4 6 1 1 2 0 / h p - o r t s a : v i X r a FRANCISCO J. CASTANDER1 COPPI3, THOMAS J. MACCARONE4 , STEPHEN E. ZEPF5, RAFAEL GUZM ´AN6, MAR´IA TERESA RUIZ2 3 , EZEQUIEL TREISTER2 3 , JOS ´E MAZA2, PAOLO S. , 2 , , 1 Institut d'Estudis Espacials de Catalunya/CSIC, Gran Capit`a 2-4, 0834 Barcelona, Spain 2 Departamento de Astronom´ıa, Universidad de Chile, Casilla 36-D, Santiago, Chile 3 Astronomy Department, Yale University, P.O. Box 208101, New Haven, CT06520, USA 4 SISSA, via Beirut 2-4, 34014 Trieste, Italy 5 Department of Physics and Astronomy, Michigan State University, East Lansing, MI 48824, USA 6 Department of Astronomy, University of Florida, P.O. Box 112055, Gainesville, FL 32611, USA Received date will be inserted by the editor; accepted date will be inserted by the editor Abstract. We present the Cal´an-Yale Deep Extragalactic Research (CYDER) Survey. The broad goals of the survey are the study of stellar populations, the star formation history of the universe and the formation and evolution of galaxies. The fields studied include Chandra deep pointings in order to characterize the X-ray faint populations. Here we present the results on the first fields studied. We find that the redshift distribution is consistent with that found in the Chandra Deep Field North. The distribution of hardness ratios is, however, softer in our sample. We find a high redshift quasar, CXOCY J125304.0-090737 at z = 4.179, which suggests that the abundance of low luminosity high redshift quasars may be larger than what would be expected from reasonable extrapolations from the quasar optical luminosity function. Key words: surveys -- quasars: general -- galaxies: active -- X-rays -- galaxies: evolution 1. Introduction: the CYDER Survey The Cal´an-Yale Deep Extragalactic Research (CYDER) Sur- vey is a collaborative effort between the Universidad de Chile and Yale University to study in detail faint stellar and ex- tragalactic populations in survey mode. The broad scientific goals are directed towards the core observing goals of the new generation of large optical and millimeter facilities. The program takes full advantage of these facilities by combining deep optical and near-IR photometric and spectroscopic ob- servations on wide field cameras and spectrographs using a wide variety of 4-m and 8-m class telescopes. The CYDER survey original design goal was to cover 1 square degree down to limiting magnitudes U ∼ 26, B ∼ 26.5, V ∼ 26, R ∼ 25.5, I ∼ 25, z ∼ 24, J ∼ 22 and K ∼ 20 at S/N ∼ 10. So far, the optical coverage is larger than 1 square degree in some filters while there is no area coverage in others. In the near infrared only 1/2 of a square degree has been imaged. Optical spectroscopy is underway in a few selected fields, while near infrared spectroscopy has not started yet. 1.1. Strategy 2. Observations Fields were selected at high galactic latitude to minimize the effects of extinction. They were also spread out in right as- cension to allow flexibility in the allocation of telescope time. Amongst our fields, we selected fields observed by the Chan- dra X-ray Observatory satellite with exposure times longer than 50 ks. Correspondence to: [email protected] The first fields studied were three of the earliest deep Chan- dra pointings to become publicly available. One of these field is in the Northern hemisphere. It is the Chandra pointing to- wards the blazar 1156+295. The X-ray exposure time was 75ks. The other two fields are in the South. They were point- ing to the Hickson compact group HCG62 (exposure time 50ks) and the blazar Q1127-145 (30ks). These two Southern fields will be the ones discussed in this paper. A.N. Author: Title 2.1. X-ray data Both the HCG62 and Q1127-145 fields were observed with the Chandra ACIS-S instrument. We retrieved these images from the archive and analyzed them using standard tech- niques with the CIAO package. In the HCG62 field we de- tect 34 X-ray sources in the s3 ACIS CCD and 16 sources in the s4. In Chandra pointing towards Q1127-145, the s3 CCD was the only CCD read. It was read in subraster mode and therefore only 5 X-ray sources are detected in that field. 2.2. Optical and Infrared Imaging We have observed both X-ray fields with the CTIO 4m MOSAIC-II camera. The total integration time for the HCG62 field is 200 minutes in U, 36min in B, 80min in V and 25 min in I under 1.0-1.5" conditions. In the Q1127-145 field the integration times are 80, 75 and 25 minutes in V, R and I respectively. Images were reduced using standard tech- niques with the IRAF/MSCRED package. In the near infrared we have observed both fields at Las Campanas Observatory with the DuPont 2.5m telescope us- ing the Wide Field InfraRed Camera during 60 minutes in J and 120 minutes in Ks under typical 0.6-0.7" seeing con- ditions. We have reduced the data with the IRAF DIMSUM package following standard procedures. 2.3. Optical Spectroscopy Follow up spectroscopy of these fields was obtained at the ESO Cerro Paranal Observatory with the UT4/Yepun tele- scope using the FORS2/MXU instrument and at the Las Cam- panas Observatory with the LDSS-2 instrument at the Magel- lan Baade telescope. Several masks were designed which in- cluded slits for most, but not all, of the X-ray sources in these fields. Masks were observed for approximately two hours. The instrument configuration used resulted in a spectral res- olution of R ∼ 520 at VLT and R ∼ 350 at Magellan. The spectra were reduced using standard techniques with IRAF and calibrated in wavelength using He-Ar comparison lamps exposures and the night sky lines. 3. First Results In our first two Southern fields we have spectroscopically identified 25 X-ray sources which corresponds to approxi- mately half of the total X-ray sources detected. Table 1 sum- marizes the percentages of the different type of objects in our sample. For comparison we also present the results for the Chandra Deep Field North (CDF-N; Brandt et al 2001, Barger et al 2002) and the Chandra Multiwavelength Project (ChaMP; Green et al 2003). We have grouped the different object types in broad classes and have translated the source types of Barger et al (2002) into these types. The CDF-N, CY- DER and ChaMP surveys reach different X-ray flux limits; the CDF-N survey reaching the faintest, the ChaMP survey, the brightest. Table 1 demonstrates how the source compo- sition changes with flux limit in an X-ray survey, although 1 Fig. 1. I magnitude plotted against full band (0.5-8.0 keV) X- ray flux. Blue crosses represent the CDF-N sources and red circles, the CYDER sources. The black solid line represents the line of constant X-ray-to-optical flux ratio, that is, objects with fI /fx = 1, where fI is the flux in the I optical band and fx the X-ray flux. The two dashed lines correspond to objects with log(fI /fx) = ±1, enclosing the region populated by normal AGN/QSOs the identification incompleteness may hide or enhance cer- tain trends. At bright flux limits, the broad line active galactic nuclei/quasars dominate the extragalactic sources. Their per- centage contribution diminishes as the flux limit lowers due to the appearence of a new population of X-ray fainter narrow emission lines AGN/QSOs and normal galaxies. Table 1. Approximate percentages of different type of sources in the CDF-N (Barger et al 2002), the CYDER and ChaMP (Green et al 2003) surveys CDFN CYDER ChaMP Stars Broad line AGN/QSO Narrow line AGN/QSO Galaxies Total Number of Identified Sources Total Number of Detected Sources 5% 25% 50% 20% 170 370 5% 40% 40% 15% 25 55 10% 65% 15% 10% ∼200 - Figure 1 shows the I band and total X-ray flux of our Chandra sources. For comparison we also plot the CDF-N sources. In our sample we find that approximately 30% of our sources do not show an optical counterpart down to I ∼ 24. In the case of the CDF-N 16% are undetected down to I ∼ 26 and 35% down to I ∼ 24. Overall, the distribution of our sources in the I .vs. X-ray flux plane occupy the same param- eter space as the CDF-N sources cut at our same X-ray flux limit. 2 Astron. Nachr./AN XXX (200X) X Fig. 2. Cumulative redshift distribution of CYDER (solid red) and CDF-N (dashed blue) sources. We compare the CYDER and CDF-N source redshift distributions in Figure 2. A Kolmogorov-Smirnov (KS) test shows that both distribution are compatible with having been drawn from the same parent population and are therefore sta- tistically indistinguishable. Given that the CDF-N sources are typically fainter, we have cut their sample at the effective X- ray flux limit of our sample. If we compare the redshift distri- bution in this case, they remain to be statistically compatible. It is worth noting that in our sample there are 5 QSOs at a redshift z ∼ 1.2 in the same field. This large scale structure feature is noticeable in our redshift distribution and stresses the need to study sufficient sources and fields as to not be biased by such structures. We have also compared the X-ray properties of our sam- ple to the CDF-N sources. We find that our sources are in gen- eral softer than those in the CDF-N (see Treister & Castander 2003). A KS test indicates that the hardness ratio distributions are incompatible with being drawn from the same parent pop- ulation. This results may be expected as the CDF-N sources are typically fainter and fainter sources are in general harder (e.g., Giacconi et al 2001; Tozzi et al 2001; Brandt et al 2001; Stern 2002). However, if we impose our effective flux limit to the CDF-N sample we still find that our sources are signifi- cantly softer than the reduced brighter CDF-N subsample. We also investigate what the optical properties of our sample are. Figure 3 is a colour-colour plot including our X-ray sources counterparts (see the figure caption for an explanation of the symbols). Two sources lie on the stel- lar locus. One is spectroscopically identified as a star, while the other has not been observed spectroscopically. The rest of the sources populate the region of colour-colour space of extragalactic sources. Some of our broad emission line AGN/QSOs are close to the expected location of typical quasars. Others, on the other hand, deviate from their ex- Fig. 3. J − K .vs. V − I colour-colour plot including the CYDER X-ray sources. The small dots represent the colours of the objects in a reduced area of our optical and near in- frared coverage of our Chandra fields. The main sequence stars (band of objects from J − K ∼ 0.3 and V − I ∼ 0.6 to J − K ∼ 0.8 and V − I ∼ 3.5) nicely separate from the rest of the extragalactic sources. The grey symbols indicate the X-ray sources (circles: sources in the HCG62 field; trian- gles: Q1127-145 field). The size of the circles and triangles is proportional to the X-ray flux. The grey intensity of the sym- bols shows the hardness ratio. The softest sources (HR=-1) are represented in dark grey and the hardest sources (HR=1) in light grey. The black line indicates the expected location of a QSO at different redshifts. The small dots in the line in- dicate separations of 0.5 units in redshift. The track starts at z = 0 at J − K ∼ 1.9 and V − I ∼ 1.1, then moves blue- wards in V − I at approximately constant J − K up to z ∼ 2, then moves blue-wards in J − K at approximately constant V − I up to z ∼ 3 − 3.5 and finally quickly moves red-wards in V −I at approximately constant J −K. The grey (coloured in electronic version) tracks represent the expected evolution- ary path of different galaxy types. The dark grey curve most to the right (red) is a 1 Gyr burst of star formation occurring 15 Gyr ago that evolves passively since. The other evolution- ary tracks are computed with star formation prescriptions that broadly reproduce the observed photometric properties of E, S0, Sa, Sb, Sbc, Sc, Sd and irregular galaxies (from right to left run from earlier types to later types). The points at the bluest J − K colour in each track represent those galaxies at z = 0. Small dots on the tracks are then spaced in 0.5 units in redshift. pected position indicating that they are probably reddened. The majority (although not all) of our soft sources lie close to the expected locus of early-type galaxies. The hardest sources populate the regions of mildly active galaxies at redshifts z ∼ 0.5 − 1.0. They may be at such locations because they are indeed this type of galaxies or because they have been reddened to that part of colour-colour space. We compare the photometric and X-ray properties of our sources. We find that the hardest sources are preferentially redder than the rest of the objects. However, we also find soft sources that are red, implying that while blue sources A.N. Author: Title 3 Fig. 4. Optical spectrum of CXOCY J125241.0-091622 ob- tained at UT4/Yepun VLT. are preferentially soft, red sources can be either X-ray hard or soft. Given the reduced number of sources in our sample, this effect could simply be a statistical fluctuation. The spectra of our sources reveal a diverse variety of type of objects. We find typical examples of broad and narrow emission line AGN/QSOs. We find typical old stellar pop- ulation spectral energy distributions. There are also examples of poststarburst galaxies and objects whose spectral classifi- cation is difficult as they have broad emission lines typical of quasars and spectral characteristics of old stellar population with some moderate on-going star formation. Some of our objects show obvious signs of strong absorption. Here we comment on two of our X-ray sources. CXOCY J125241.0-091622 is a quasar at redshift z = 2.282 (Fig- ure 4) with a harder than usual X-ray spectrum Γ ∼ 1.3. In the optical its V − K color is 4.0 which is one magni- tude redder than the expected colour of a prototype quasar at this redshift (Figure 3). Both X-ray and optical data thus indicate that this is an obscured object. Such objects are pre- dicted in models as contributors to the X-ray Background at faint fluxes. CXOCY J125304.0-090737 is an optically faint quasar (MB = −23.69+5 ×log(h65)) with a typical QSO spectrum (figure 5). In X-rays, CXOCY J125304.0-090737 is also X- ray faint (fX = 1.7 ± 0.4 × 10−15 ergs s−1 cm−2 in the [0.5- 2.0] keV band) with a somewhat harder spectrum (Γ ∼ 1.7 or HR ∼ −0.35) than typical low redshift or high redshift op- tically selected quasars. We speculate that a reflection com- ponent can slightly harden the spectrum but by no means is this the only mechanism. CXOCY J125304.0-090737 X-ray- to-optical emission is X-ray strong (αox ∼ −1.35) compared to high redshift optically selected quasars (see Castander et al 2002 for further details). Fig. 5. Optical spectrum of CXOCY J125304.0-090737 ob- tained at UT4/Yepun VLT. For comparison, we also show the error weighted average of the SDSS Early Data Release QSO spectra at z > 4. This composite spectrum have arbitrary zero point and scaling offsets for display purposes and it is shown above the spectrum of CXOCY J125304.0-090737. The most common QSO emission lines are indicated as dotted lines. The emission lines in CXOCY J125304.0-090737 are nar- rower than the typical SDSS spectrum. CXOCY J125304.0-090737 is the only quasar above z = 4 found so far in the CYDER survey. However, the space den- sity implied by its discovery is higher than reasonable extrap- olations of the quasar optical luminosity function to fainter luminosities at high redshifts (Fan et al 2001) . Although this object by itself does not constrain the faint end of the lumi- nosity function at high redshift, it demonstrates the possibili- ties that X-ray surveys have to achieve this goal. 4. Summary We have briefly presented the CYDER survey, which is cur- rently underway. We have focused in the first fields studied that were chosen to coincide with moderately deep Chandra X-ray pointings. We have investigated the nature of the X-ray sources in these fields. We find that the broad X-ray and op- tical properties of our sources are similar to the ones studied in the CDF-N. The only difference is that our sources appear to be softer. We also stress the fact that our X-ray sources are of diverse optical spectral types and have highlighted the discovery of a high redshift X-ray selected quasar. The current X-ray surveys (most of which are presented in this volume) that have flourished with the launch of the new X-ray observatories promise to be key to our understand- ing of the faint X-ray populations. The CYDER survey will be one of the surveys contributing to pin down the evolution of accretion material on to black holes which seems to be closely related to the star formation history of the universe. Acknowledgements. We acknowledge the financial support received from the Fundaci´on Andes that has enabled the collaboration be- tween the Universidad de Chile and Yale University and therefore has made the CYDER survey possible. Astron. Nachr./AN XXX (200X) X 4 References Barger, A. J. et al.: 2002, AJ, 124, 1839 Brandt, W. N. et al.: 2001, AJ, 122, 2810 Castander, F. J. et al.: 2002, AJ, submitted Giacconi, R. et al.: 2001, ApJ, 551, 624 Green, P.J. et al: 2003: AN, this volume Fan, X. et al: 2001, AJ, 121, 31 Stern, D. et al.: 2002, AJ, 123, 2223 Tozzi, P. et al.: 2001, ApJ, 562, 42 Treister, E. & Castander, F. J.: 2003, AN, this volume
astro-ph/0505322
1
0505
2005-05-16T15:59:14
An Artificial Neural Network Approach to the Solution of Molecular Chemical Equilibrium
[ "astro-ph" ]
A novel approach is presented for the solution of instantaneous chemical equilibrium problems. The chemical equilibrium can be considered, due to its intrinsically local character, as a mapping of the three-dimensional parameter space spanned by the temperature, hydrogen density and electron density into many one-dimensional spaces representing the number density of each species. We take advantage of the ability of artificial neural networks to approximate non-linear functions and construct neural networks for the fast and efficient solution of the chemical equilibrium problem in typical stellar atmosphere physical conditions. The neural network approach has the advantage of providing an analytic function, which can be rapidly evaluated. The networks are trained with a learning set (that covers the entire parameter space) until a relative error below 1% is reached. It has been verified that the networks are not overtrained by using an additional verification set. The networks are then applied to a snapshot of realistic three-dimensional convection simulations of the solar atmosphere showing good generalization properties.
astro-ph
astro-ph
Astronomy & Astrophysics manuscript no. 2865 (DOI: will be inserted by hand later) October 24, 2018 5 0 0 2 y a M 6 1 1 v 2 2 3 5 0 5 0 / h p - o r t s a : v i X r a An Artificial Neural Network Approach to the Solution of Molecular Chemical Equilibrium A. Asensio Ramos1 and H. Socas-Navarro2 1 Istituto Nazionale di Astrofisica (INAF) Osservatorio Astrofisico di Arcetri, Largo Enrico Fermi 5, 50125 Florence, Italy 2 High Altitude Observatory, NCAR⋆, 3450 Mitchell Ln, Boulder CO 80307-3000, USA Received ¡date¿ / Accepted ¡date¿ Abstract. A novel approach is presented for the solution of instantaneous chemical equilibrium problems. The chemical equilibrium can be considered, due to its intrinsically local character, as a mapping of the three- dimensional parameter space spanned by the temperature, hydrogen density and electron density into many one-dimensional spaces representing the number density of each species. We take advantage of the ability of arti- ficial neural networks to approximate non-linear functions and construct neural networks for the fast and efficient solution of the chemical equilibrium problem in typical stellar atmosphere physical conditions. The neural network approach has the advantage of providing an analytic function, which can be rapidly evaluated. The networks are trained with a learning set (that covers the entire parameter space) until a relative error below 1% is reached. It has been verified that the networks are not overtrained by using an additional verification set. The networks are then applied to a snapshot of realistic three-dimensional convection simulations of the solar atmosphere showing good generalization properties. Key words. Molecular processes -- Astrochemistry -- Methods: numerical 1. Introduction Molecules are usually found in highly dynamic systems (e.g., the solar atmosphere, winds of AGB stars, ...) and their formation is influenced by the time variation of the physical conditions in the medium. In many situa- tions the dynamical timescales are much slower than the timescales of molecular formation, and the approximation of Instantaneous Chemical Equilibrium (ICE) can be used with extremely good results (Russell 1934; Tsuji 1964, 1973; McCabe et al. 1979, etc.). Under the ICE approxi- mation, molecules and ions are assumed to form instan- taneously and their abundances depend only upon the local temperature and density. Another consequence of this assumption is that the specific reaction mechanisms that create and destroy a given molecule are irrelevant and only the molecule and its constituents are impor- tant. Concerning the solar atmosphere, it has been re- cently shown with the comparison between ICE calcula- tions and detailed chemical evolution calculations that the formation of the strong CO lines around 4.7 µm can be well described using the ICE approximation for heights be- low ∼700 km (Asensio Ramos et al. 2003). On the other hand, chemical evolution effects in regions of the atmo- sphere above this height are of importance in setting the CO abundance. In order to calculate the atomic and molecular number densities using the ICE approximation one needs to solve a non-linear system of equations (see below). Although efficient numerical methods exist for solving this kind of systems, the computing time when the ICE problem has to be solved in a large amount of points (e.g., dense grids, multi-dimensional geometries, iterative inversions, etc) be- comes prohibitive. It is then very important to develop a numerical method for the rapid solution of chemical equi- librium problems. Artificial Neural Networks (ANNs) have proven to be a powerful approach to a broad variety of problems (see, e.g., Bishop 1996). In the solar community, they have been recently applied to the problem of inferring the magnetic field from observations of the polarization profiles of se- lected atomic lines (Carroll & Staude 2001; Socas-Navarro 2003, 2005). In this paper, we make use of the ability of ANNs with one hidden layer to approximate any non- linear continuous function (e.g., Jones 1990; Blum & Li 1991) to solve the ICE problem. ⋆ The National Center for Atmospheric Research (NCAR) is sponsored by the National Science Foundation The structure of the paper is as follows. Section 2 de- scribes the ICE approximation and its standard solution. 2 A. Asensio Ramos and H. Socas-Navarro: Neural Network approach to ICE Section 3 describes our approach to the ICE problem us- ing ANNs, discussing how it can be trained to associate a combination of physical parameters with the number density of each species included in the problem. Section 4 details how the trained network can be applied to solve the ICE problem in realistic convection simulations of the solar atmosphere. In section 5 we show the dependence of the outputs of the ANN on the physical parameters, which can be easily done because of the intrinsic analytic character of the mapping generated by the ANN. Finally, the most relevant conclusions are summarized in Section 6. 2. Instantaneous chemical equilibrium 2.1. Basic equations Consider three elements A, B and C that stick together to form a molecule AaBbCc, in which a, b and c are the stoichiometric coefficients indicating the number of times an element is present in the molecule. Because the explicit reactions which form the molecule are irrelevant under the ICE approximation, one needs only to consider the follow- ing dissociative process for the formation of the molecule AaBbCc: AaBbCc ⇔ aA + bB + cC. (1) This reaction is characterized by its dissociative reaction constant or equilibrium constant. It is given by the ratio between the product of the partial pressures of the indi- vidual elements and the partial pressure of the molecule: We now define the fictitious pressure P (i) as the pressure exerted by element i if all the gas were in the neutral form of the atomic species i. It is customary to specify the partial pressures of all the atomic species in terms of the fictitious pressure of hydrogen P (H). The coefficient relating one and the other is the abundance of each ele- ment with respect to hydrogen, A(i), so P (i) = A(i)P (H). The abundance of each atomic species depends on the metallicity of the star we are considering. For the solar case, we have used the standard solar abundances given by Grevesse (1984). Under ICE conditions, the number of atoms and molecules are obtained by solving the conservation of mass and the chemical equilibrium conditions given by Eq. (2). The conservation of mass establishes that the sum of the partial pressures of all the species containing a given atomic element (taking into account the stoichiometry) equals the fictitious pressure of the given element: P (i) = Pi + Pi+ + Pi− +Xk ωi kPk, (7) where Pi, Pi+ and Pi− are the partial pressure of the neu- tral element i, and the ionized elements i+ and i−, respec- tively, while Pk is the partial pressure of molecule k which has species i in its composition. ωi k is the stoichiometric coefficient which indicates the number of times element i appears in molecule k. The sum is extended over all the molecules in which the element i takes part. The previous equation can be rewritten with the aid of Eqs. (2) and (6): Kp(T ) = BP c P a AP b C PABC , (2) P (i) = Pi +Ki+ Pi Pe− +Ki−PiPe− +Xk ωi k ωj P ωi i P k j k · · · P ωl k l K (k) p (T ) , where Kp(T ) is a function of the temperature and Pi the partial pressure of the species i. For example, the equi- librium constants for the dissociation of CO and H2 are given by K CO p (T ) = PC PO PCO , K H2 p (T ) = P 2 H PH2 . (3) The partial pressure is usually related to the number den- sity ni of a given species by the ideal gas equation: Pi = nikT, (4) where k is the Boltzmann constant and T the local tem- perature. This partial pressure is the pressure of the gas if no other species were present. The sum of all the partial pressures is the total pressure in the medium. Following the same line of reasoning, one can consider the ionization equilibria for the atomic species: A ⇔ A+ + e− A + e− ⇔ A−, with their corresponding equilibrium constants: KA+(T ) = PA+ Pe− PA KA−(T ) = PA− PAPe− . . . . , k + ωj (8) l are the atomic species composing where i, j, k + . . . + ωl molecule k, ωi k = nk is the number of atoms of molecule k and K (k) p (T ) is its equilibrium con- stant. Charged molecules can be easily included by taking into account their dissociation and the ionization equilib- rium, simultaneously. Consider the dissociation equilib- rium of an ionized diatomic molecule. One only needs to take into account that when a charged molecule dissoci- ates, the atomic element which remains ionized is the one with the lowest ionization potential: AB+ ⇔ A+ + B, (9) with D0(A) < D0(B). Therefore, using Eq. (2) and Eq. (6) we can calculate the partial pressure of the ionized molecule: Kp(T ) = PA+ PB PAB+ = KA+(T ) Pe− PAPB PAB+ , (10) (5) (6) We can write an equation like Eq. (8) for each of the Ns atomic species included in the calculation. In our case we have selected the 21 species which are shown in Table 1. Therefore, once the molecular and ionization equilib- rium constants are known, we have a set of Ns non-linear A. Asensio Ramos and H. Socas-Navarro: Neural Network approach to ICE 3 algebraic equations which depend on the local tempera- ture and density. The ICE approximation is intrinsically local since it depends only on the local values of the tem- perature and density. Therefore, to calculate the molecu- lar number densities in a model atmosphere, we have to solve the algebraic non-linear system of equations at each point in the atmosphere. This system is solved using a standard Newton-Raphson iterative method (Press et al. 1986). Once the partial pressures for all the atomic species are known, we can calculate the ensuing atomic number densities by using Eq. (4). The molecular number densi- ties are calculated by solving for their partial pressures in Eq. (2), using the equilibrium constant appropriate for each molecule. 2.2. Equilibrium constants As we discussed above, the problem of obtaining the atomic and molecular number densities is completely de- fined once the temperature and the total density (or pres- sure) is known. However, we also need to know the value of the equilibrium constant for each temperature. It is known from considerations of statistical mechanics that the value of the equilibrium constant can be obtained with the aid of the partition function of the molecule and of the individual atomic components, their respec- tive masses and the dissociation energy D0 (see, e.g., Tejero Ordonez & Cernicharo 1991, for detailed informa- tion on how the equilibrium constants can be calculated). The expression is given by (Tejero Ordonez & Cernicharo 1991, and references therein): Kp(T ) = (kT )N A−1(cid:18) 2πkT h2 (cid:19)3(N A−1)/2(cid:18) ma Amb Bmc mABC (cid:19) C 3/2 Aφb ×(cid:18) φa Bφc φABC (cid:19) e−D0/kT , C (11) where mi is the mass of atomic or molecular species i, N A = a + b + c is the number of atoms present in the molecule, φi is the partition function of species i. h and k stand for the Planck and Boltzmann constants, respec- tively. The partition functions can be obtained by per- forming the summation φ = Xj gje−Ej /kT , (12) where gj is the degeneracy of level j and Ej its en- ergy. Ideally, the sum has to be extended over all the energy levels of the species. However, due to the difficulty of accounting for all the energy levels, the most complete available set of energy levels must be used. For molecules, the partition function has to in- clude the contribution from the electronic, vibrational and rotational levels. It is not practical to compute this every ICE calculations. As a workaround, some authors have tabulated polynomial fits to the partition functions (Russell 1934; Tsuji 1971; Sauval & Tatum 1984; Tejero Ordonez & Cernicharo summation for 1991). In this work, we make use of the results obtained by Tejero Ordonez & Cernicharo (1991). It represents one of the most complete compilations of molecular equilibrium constants for more than 240 diatomic and polyatomic molecules. These equilibrium constants were obtained by calculating the partition function with the most up-to-date molecular constants at the time of the work and their dependence with the temperature are given, as in previous works, as polynomial fits. One of the advantages of selecting this database is its self-consistency. 3. Neural network 3.1. Description of the network The ICE problem consists on obtaining the partial pres- sures of the atomic species (or the atomic species number densities) which are consistent with Eq. (8) once the lo- cal temperature, local hydrogen density and local electron density are given. One could also take a different point of view and see the ICE problem as a mapping between the three-dimensional space given by (T, nH , ne) onto sev- eral one-dimensional spaces, one for each atomic species included in the problem. The functional form of the func- tions f : R3 → R is not known. The Artificial Neural Network (ANN) with one hidden layer is a universal approximant to any non-linear con- tinuous function (e.g., Jones 1990; Blum & Li 1991). The schematic structure of such a network is shown in Fig. 1. We have constructed an ANN with an input layer of three neurons where the temperature, hydrogen density and electron density are introduced. The neurons of the input layer have a linear activation function. The hidden layer consists on Nh neurons with a non-linear activation function σ(x). Each hidden neuron is connected to all the neurons of the input layer by a certain weight (indicated by arrows in the figure). The value obtained at each hidden neuron is a linear combination of the values at the neurons of the input layer multiplied by the weights. These values are then applied the non-linear function σ(x), multiplied by another set of weights and summed to give the final output of the neural network. The output of the network can be written as: j T + wH j n(H) + we j n(e) + uj(cid:3) , N (T, n(H), n(e), m) = Nh Xj vjσ(cid:2)wT (13) where the activation function is usually given by the sig- moid or the hyperbolic-tangent function. In this case, we have selected σ(x) = tanh(x). In the previous expression, m formally represents the whole set of 5 × Nh weights which define the neural network. The use of a neural network approach for the solu- tion of the ICE problem is very appealing for several rea- sons. Firstly, we obtain an analytic function which is in- finitely differentiable and which could be easily used in any subsequent calculation. This way, we can investigate with 4 A. Asensio Ramos and H. Socas-Navarro: Neural Network approach to ICE T n(H) n(e )- gives a good coverage of the parameter space, consider- ing that they are randomly selected. The temperature is varied in the range from 3500 K and 14000 K, which is representative of the physical conditions in typical stel- lar atmospheres. Concerning the hydrogen and electronic abundances, we decided to use a logarithmic scaling. The hydrogen density is varied from 1012 cm−3 to 1018 cm−3 while the electronic density is varied from 108 cm−3 to 1017 cm−3. Again, these values are a reasonable represen- tation of realistic conditions. Once the learning set is selected, the training of the network for each species i reduces to a minimization of the following error function: Pi Fig. 1. Schematic structure of the neural network model used for approximating the ICE results. The input layer and the output layer are linear, while the hidden layer is assumed to be non-linear, with an hyperbolic-tangent activation function. The input layer consists of three neu- rons for the temperature, hydrogen number density and electronic number density. The output neuron gives the partial pressure of element i for the physical conditions in the input layer. great detail the dependence of the atomic and/or molec- ular abundances with temperature, hydrogen density and electron density. Secondly, the powerful approximation ca- pabilities of ANNs make them very suitable for multi- dimensional interpolation problems. Even more striking is the ability of ANNs to extrapolate data outside the range of parameters used in the learning process. A direct con- sequence of these properties is that the number of points that need to be used for the learning process can be very small compared to the case of standard interpolation tech- niques. Furthermore, the number of parameters needed for performing the interpolation is also small. Essentially, one transforms the problem from storing in memory the whole set of N points which relates T , n(H) and n(e) with the partial pressures Pi for every species to storing only the 5 × Nh m weights present in Eq. (13) for each species. This reduction is proportional to the ratio N/Nh, which can be very large for large, as we will see below. Finally, it is also important to note that the neural network can be straightforwardly implemented in parallel architectures. 3.2. Learning process In order to obtain a neural network which is able to ap- proximate the abundance of each species included in the chemical equilibrium calculation, we randomly select Nl sets of temperature, hydrogen density and electronic den- sity. For each one of these combinations, we solve the ICE equations and calculate the abundances of all the atomic species. We have verified that taking Nl = 1000 Ei = Nl Xl=1 (cid:2)P l i − N (Tl, nl(H), nl(e), m)(cid:3)2 , (14) so that the optimum values of the parameters m are ob- tained. P l i is the partial pressure (or number density) of species i obtained for the combination of physical condi- tions l. Since almost every minimization algorithm makes use of the derivatives of the error function with respect to the parameters, it is of interest to notice that they can be obtained analytically: ∂E ∂mk = −2 Nl Xl=1 [Pi − N (Tl, nl(H), nl(e), m)] ∂N ∂mk , (15) where the partial derivatives of the neural network with respect to the parameters are obtained in a straightfor- ward manner by using Eq. (13). In order to obtain the set of parameters m which minimize the error function, we have used a quasi-Newton BFGS method (Fletcher 1987) implemented in the Merlin package (Papageorgiou et al. 1998) which presents a very good convergence rate. It is interesting to note that the usual approach of training the neural network with a learning set obtained from the original problem can be circumvented here. We do not need to rely on the precision given by the Newton- Raphson solution and the neural network can be made as precise as desired. This is a consequence of the fact that the equations describing the ICE approximations are known. One can see that the equations of ICE can be for- mally written as f (T, n(H), n(e), {Pi}) = 0, where {Pi} is the set of Ns partial pressures (or atomic number densi- ties) which satisfy the equations. Therefore, we can build as many neural networks as species included in the ICE problem and train the network by minimizing the follow- ing error function: Nl E = [f (Tl, nl(H), nl(e), {Ni(Tl, nl(H), nl(e), m)})]2 . Xl=1 (16) The solution to the problem is transformed into finding the set of parameters m for all the networks which minimizes the previous error function. The summation is extended over the Nl elements of the learning set. Several issues are noteworthy in this case. On the one hand, the derivatives A. Asensio Ramos and H. Socas-Navarro: Neural Network approach to ICE 5 Fig. 2. Relative error obtained after the training of the neural network for hydrogen, oxygen, carbon and iron. The distribution of relative errors are shown for the learning set and for a verification set, which aids at detecting over- training. We indicate the value of the standard deviation σ of the distribution and the number of points below 1σ, 2σ and 3σ. We also indicate these values in the plots as horizontal lines. of the error function with respect to the parameters m are not as straightforward as those found in Eq. (15) since the function f (T, n(H), n(e), {Pi}) = 0 is non-linear. On the other hand, the minimization process has to be carried out simultaneously for the Ns neural networks, so that the problem of finding the minimum might be more difficult to solve. The dimension of the hypersurface in which we have to find the minimum is 5×Nh×Ns, instead of finding Ns minima in hypersurfaces of dimension 5 × Nh. A critical parameter in the training ability of a neural network is the number of hidden neurons Nh. It is impor- tant to build networks with sufficient number of hidden networks to accurately approximate the non-linear func- tion we are interested in. However, since the number of learning points is not infinite, the number of hidden neu- rons one can use is practically limited by the overtraining phenomenon (in principle, the number of training points should be much larger than the number of hidden neu- rons). If Nh is very large, the network is capable of cor- rectly approximating all the points in the learning set but its interpolation capacity is lost, producing large oscilla- tions between the points of the learning set. This is a di- rect consequence of the increase in the degrees of freedom of the network when the number of hidden neurons is in- creased. One way to avoid this overtraining is to use two different data sets: one for learning purposes and the other for testing. When the training process takes places, the er- ror given by expression (14) is evaluated for the training set and for the test set. When the network is training cor- rectly, both errors are reduced. When the network starts to be overtrained, the error of the test set starts to increase. We have followed this scheme in order to stop the train- ing process before overtraining. In our case, we have used two different values for the number of hidden neurons. We have verified that Nh = 20 gives sufficiently accurate re- sults for 13 of the 21 atomic species included in the ICE calculations, while we were forced to increase this num- ber to Nh = 30 for the rest of species. Table 1 lists the number of hidden neurons for each species. We did not investigate the dependence of the minimum error on the number of hidden neurons so that the increase to Nh = 30 was somewhat arbitrary. The training of the neural networks was stopped when the standard deviation of the relative error between the output of the neural network and the values obtained from the solution of the ICE equations for both the learning and the test sets was 1% or below. Obviously, the relative er- ror was always larger for the test set than for the learning 6 A. Asensio Ramos and H. Socas-Navarro: Neural Network approach to ICE Fig. 3. Histograms of relative errors for the 2.55×105 points of the three-dimensional convection simulation of Asplund et al. (2000). We indicate the position of 1σ, 2σ and 3σ of the distribution. Note that they are similar to those obtained for the verification set during the training of the networks. Table 1. Number of hidden neurons used for each ANN. Species Number of hidden neurons He, Na, Mg, Al, P, K, Ca Ti, Cr, Mn, Fe, Ni, Cu H, C, N, O, F, Si, S, Cl 20 30 set, but the training always reached the limit of 1% and was stopped before any indication of overtraining was ob- served. In Fig. 2 we present the relative error obtained for the 1000 training points for the abundances of H, C, O and Fe. The behavior of the relative error for the rest of species follows the same behavior, except for He, which presents a much smaller dispersion. Since He cannot form molecules, it is always in its atomic form and the non-linear mapping that the neural network has to learn is strongly simplified. We have indicated in each plot the value of the stan- dard deviation and the percentage of points which are con- tained within 1σ, 2σ and 3σ. This is an indication of the precision obtained with the neural network. It is common in all the networks that more than ∼ 95% of the points are inside 3σ, which translates into a relative error of ∼ 3% (since σ is always of the order or below 1%). However, we note that it is common to all the networks that more than 75% of the points are below 1σ. Along with the plots of the results obtained for the learning set, we have in- cluded the results for the verification set. It is interesting to note that, although the standard deviation of the dis- tribution of relative errors is larger than that obtained for the learning set, typically more than ∼ 80% of the points are below this limit. In the case of the network for the carbon abundance, we have a relatively high value for σ, but more than 90% of the points are below this limit. 4. Comparison of results Once the neural networks were trained to the desired pre- cision, as shown in Fig. 2, we have applied them to realistic situations in order to compare their performance with the complete ICE calculations. To this end, we make use of a three-dimensional snapshot of the convection simulation of the solar atmosphere of Asplund et al. (2000). We con- sider this simulation to be an adequate representation of the physical conditions in the solar atmosphere since sev- eral investigations are reporting good agreements between the synthetic line profiles and the observed ones (e.g., Shchukina & Trujillo Bueno 2001; Asplund et al. 2004, 2003). The snapshot box size is 50x50x102, so that the number of points is 2.55×105. We have solved the ICE problem in each point and confronted the results with those obtained from the neural networks. A. Asensio Ramos and H. Socas-Navarro: Neural Network approach to ICE 7 Fig. 4. Number density of carbon and oxygen obtained with the simplified model and with the neural network for two different combinations of hydrogen and electron densities and for different values of the temperature. Note that this kind of investigation is greatly simplified due to the analytical character of the neural network. The normalized error histograms are shown in Fig. 3 for H, C, O and Fe, with vertical dashed lines indicating the position of 1σ, 2σ and 3σ of the distribution. Note that the histograms are decreasing functions of the rela- tive error, so that the number of points with large rela- tive error is much smaller than the number of points with small relative errors. It is verified that the value of the standard deviation of the distribution of relative errors is quite similar to those obtained for the verification set. This reinforces our confidence that the learning set and the verification set provide sufficient coverage of the three- dimensional parameter space (T, n(H), n(e)). The shapes of the histograms turn out to be fairly linear in this log- linear scale, so that we can consider, as a first approxima- tion, that the number of points Np with a certain relative error r is given by Np ∝ r−α. For example, we find that α ≃ 5 for hydrogen, α ≃ 3.7 for carbon and α ≃ 7 for oxy- gen. There is an exception for iron, for which the distri- bution of points with relative errors below ∼ 2.3% seems to be constant, while the number of points with errors larger than ∼ 2.3% is three order of magnitude smaller. The behavior of the histograms for the rest of species is very similar to those shown in Fig. 3 for the selected ones, maintaining the properties obtained for the verification sets. 5. T , n(H) and N (e) dependence Since the neural network approach leads to analytic for- mulae for the solution of ICE problems, we now take ad- vantage of them and compare the results for abundant species with simplified models. Our aim is to investigate the variation with temperature, hydrogen density and electronic density of the number density of carbon and oxygen. We have selected these two species because they are the constituents of the highly abundant CO diatomic molecule. This way, we can safely obtain an approximation to the carbon and oxygen abundances by taking only into account the formation of CO in the ICE equations and ne- glecting the rest of molecular species. Note that this sim- plification is only valid for species like C and O and not for species which form less abundant molecules and/or form several high abundance molecules. Specializing Eq. (8) to oxygen and carbon, we obtain: P (C) = PC + P (O) = PO + PC PO KCO(T ) PC PO KCO(T ) + KC + + KO+ PC Pe− PO Pe− + KC−PC Pe− + KO−POPe− . (17) This system of equations can be solved analytically for PC and PO and transformed into number densities by us- ing the ideal gas equation Eq. (4). Their dependence with temperature for two different combinations of hydrogen and electronic densities are shown in Fig. 4. The output of the neural network is also shown in the plots with solid lines. Note the accurate results obtained with this simpli- fied model in this high abundance case. This behavior can be better understood if we plot separately the contribu- tion to the total pressure of each element of each term in Eq. (17). The results are shown in Fig. 5 for the two sets of physical conditions. The fictitious pressure of carbon and oxygen (essentially the abundance of carbon and/or oxygen times the hydrogen density) is shown by solid line, the partial pressure of the neutral element by dashed line, the partial pressure of the carbon monoxide molecule by dash-dotted line and the partial pressure of the ionized element by long dashes. All the quantities shown in the plots have been obtained using the neural network ap- proach. Although the neural network takes into account the formation of other molecules which contain carbon and oxygen (i.e., it gives the solution to the complete ICE problem), we see that accounting only for CO appears to be enough since the addition of all the contribution closely approximates the value of the fictitious pressure. Note that the CO molecule is efficiently formed when the temperature is below ∼5000 K in both situations. The neutral species are mainly situated at intermediate tem- peratures, typical from photospheric regions, while the 8 A. Asensio Ramos and H. Socas-Navarro: Neural Network approach to ICE Fig. 5. Contribution to the total pressure of carbon and oxygen in two different cases. The curves have been obtained using the neural networks. The addition of all the contribution closely matches the fictitious pressure of carbon and oxygen. This is the reason why the very simplified model for C and O works very well. ionized species tend to dominate for high temperatures, above 6000 − 7000 K. 6. Conclusion We have successfully trained 21 neural networks which approximate the solution to the Instantaneous Chemical Equilibrium problem. We have generated a learning set which is representative of the physical conditions in stel- lar atmospheres and have verified that the coverage of the three-dimensional parameter space is sufficiently good with only 1000 points. We have trained the neural net- works so that the standard deviation of the relative er- ror were below or of the order of 1%. In order to avoid overtraining, we have employed an independent verifica- tion set. The standard deviation of the relative error in this verification set is larger than that for the learning set but also of the order of 1%. We apply the neural net- works obtained to the ICE problem in a three dimensional convection simulation, representative of the physical con- ditions in the solar atmosphere. The histograms of rela- tive errors built with all the points in the simulation show that the neural networks are capable of solving the ICE problem with relative errors similar to those obtained for the verification set. The advantage of the neural network approach is their intrinsic analytical character, the possi- bility of parallelization and the very fast evaluation. The development of a fast approach to ICE is of importance for, among others, recent multi-dimensional simulations of stellar atmospheres which include very dense grids and iterative inversions of observed spectra. Acknowledgements. This research has been partly funded by the Ministerio de Educaci´on y Ciencia through project AYA2004-05792 and by the European Solar Magnetism Network (contract HPRN-CT-2002-00313). References Asensio Ramos, A., Trujillo Bueno, J., Carlsson, M., & Cernicharo, J. 2003, ApJ, 588, L61 Asplund, M., Carlsson, M., & Botnen, A. V. 2003, A&A, 399, 31 Asplund, M., Grevesse, N., Sauval, A. J., Allende Prieto, C., & Kiselman, D. 2004, A&A, 417, 751 Asplund, M., Ludwig, H. G., Nordlund, A., & Stein, R. F. 2000, A&A, 359, 669 Bishop, C. M. 1996, Neural networks for pattern recogni- tion (Oxford University Press) Blum, E. K. & Li, L. K. 1991, Neural Networks, 4, 511 Carroll, T. A. & Staude, J. 2001, A&A, 378, 316 Fletcher, R. 1987, Practical Methods of Optimization, 2nd ed. (New York: Wiley) Grevesse, N. 1984, Phys. Scr, 8, 49 Jones, L. K. 1990, in Proceedings of the IEEE, 78, 1585 McCabe, E. M., Connon Smith, R., & Clegg, R. E. S. 1979, Nature, 281, 263 Papageorgiou, D. G., Demetropoulos, I. N., & Lagaris, I. R. 1998, Comput. Phys. Commun., 109, 227 Press, W. H., Teukolsky, S. A., Vetterling, W. T., & Flannery, B. P. 1986, Numerical Recipes (Cambridge: Cambridge University Press) Russell, H. N. 1934, ApJ, 79, 281 Sauval, A. J. & Tatum, J. B. 1984, ApJS, 56, 193 Shchukina, N. & Trujillo Bueno, J. 2001, ApJ, 550, 970 Socas-Navarro, H. 2003, Neural Networks, 16, 355 Socas-Navarro, H. 2005, ApJ, 620, 517 Tejero Ordonez, J. & Cernicharo, J. 1991, Modelos de Equilibrio Termodin´amico Aplicados a Envolturas Circunestelares de Estrellas Evolucionadas (Madrid: IGN) Tsuji, T. 1964, Ann. Tokio. Astron. Obs., 2nd ser. 9, 1 Tsuji, T. 1971, Ann. Tokio Astron. Obs. 2nd ser., 9, 1 Tsuji, T. 1973, A&A, 23, 411
astro-ph/0205121
1
0205
2002-05-08T15:00:31
The X-ray emission from Nova V382 Velorum: II. The super-soft component observed with BeppoSAX
[ "astro-ph" ]
Nova Velorum 1999 (V382 Vel) was observed by BeppoSAX 6 months after optical maximum and was detected as a bright X-ray supersoft source, with a count rate 3.454+-0.002 cts/s in the LECS. It was the softest and most luminous supersoft source observed with this instrument. The flux in the 0.1-0.7 keV range was not constant during the observation. It dropped by a factor of 2 in less than 1.5 hour and then was faint for at least 15 minutes, without significant spectral changes. The observed spectrum is not well fit with atmospheric models of a hot, hydrogen burning white dwarf. This is due mainly to a supersoft excess in the range 0.1-0.2 keV, but the fit can be significantly improved at higher energy if at least one emission feature is superimposed. We suggest that a ``pseudocontinuum'' was detected, consisting of emission lines in the supersoft X-ray range superimposed on the thermal continuum of a white dwarf atmosphere. As a result, an accurate determination of the effective temperature and gravity of the white dwarf at this post-outburst stage is not possible.
astro-ph
astro-ph
Mon. Not. R. Astron. Soc. 000, 1 -- ?? (1994) Printed 24 October 2018 (MN LATEX style file v1.4) The X-ray emission from Nova V382 Velorum: II. The super-soft component observed with BeppoSAX M. Orio1,2, A. N. Parmar3, J. Greiner,4 H. Ogelman5, S. Starrfield6, E. Trussoni1 1 Istituto Nazionale di Astrofisica (INAF), Osservatorio Astronomico di Torino, Strada Osservatorio, 20, I-10025 Pino Torinese (TO), Italy 2 Department of Astronomy, 475 N. Charter Str., University of Wisconsin, Madison WI 53706, USA 3 Astrophysics Division, Space Science Dept. of ESA, ESTEC, Postbus 299, 2200 AG Noordwijk, The Netherlands 4 Astrophysical Institut, 14882 Potsdam, An der Sternwarte 16, FRG 5 Department of Physics, University of Wisconsin, 1500 University Ave., Madison WI 53706, USA 5 Dept of Physics and Astronomy, P.O. Box 87150, Arizona State University, Tempe AZ 85287-1504, USA Received; Accepted ABSTRACT Nova Velorum 1999 (V382 Vel) was observed by BeppoSAX 6 months after optical maximum and was detected as a bright X-ray supersoft source, with a count rate 3.454±0.002 cts s−1 in the LECS. It was the softest and most luminous supersoft source observed with this instrument. The flux in the 0.1 -- 0.7 keV range was not constant during the observation. It dropped by a factor of 2 in less than 1.5 hour and then was faint for at least 15 minutes, without significant spectral changes. The observed spectrum is not well fit with atmospheric models of a hot, hydrogen burning white dwarf. This is due mainly to a supersoft excess in the range 0.1-0.2 keV, but the fit can be significantly improved at higher energy if at least one emission feature is superimposed. We suggest that a "pseudocontinuum" was detected, consisting of emission lines in the supersoft X-ray range superimposed on the thermal continuum of a white dwarf atmosphere. As a result, an accurate determination of the effective temperature and gravity of the white dwarf at this post-outburst stage is not possible. Key words: Stars: individual: V382 Vel, novae, cataclysmic variables -- Sources as function of wavelength: X-rays: stars 1 INTRODUCTION Nova Velorum 1999 (V382 Vel) was the second brightest nova of the last 50 years (V=2.6) (Seargent & Pearce, 1999). It was a "fast " O -- Ne -- Mg nova, with v≃4000 km s−1, t2= 6 d, t3=10 d (Della Valle et al. 1999, Shore et al. 1999). The nova was immediately declared a Target of Opportunity by the BeppoSAX Mission Scientist. BeppoSAX carries instruments that cover the energy range 0.1-300 keV. The instruments used are the coalined Low-Energy Concentrator Spectrometer (LECS; 0.1 -- 10 keV; Parmar et al. 1997), the Medium-Energy Concentrator Spec- trometer (MECS; 1.8 -- 10 keV; Boella et al. 1997), and the Phoswich Detection System (PDS; 15 -- 300 keV; Frontera et al. 1991). The LECS and MECS, used in this work, consist of grazing incidence telescopes with imaging gas scintillation proportional counters in their focal planes. Novae in outburst have been observed to emit X-rays due to thermal emission of shocked ejecta (see Orio et al. 2001a, and references therein). The outburst is normally due to a radiation pressure driven wind and not to a shock wave, however shocks can be produced in interacting winds, or in- teraction between the ejecta and the circumstellar medium. c(cid:13) 1994 RAS After a few months, luminous "supersoft" X-ray emission (luminosity of the order 1037−38 erg s−1) has also been observed. The previous detections have been attributed to residual hydrogen burning in a shell on the white dwarf rem- nant (e.g. Ogelman et al. 1993, Krautter et al. 1996, Orio & Greiner 1999). We expect to detect in this case an at- mospheric continuum at Teff =20-80 eV and the absorption edges of the white dwarf (or even emission edges if the ef- fective temperature is extremely high). V382 Vel was observed with BeppoSAX, with ASCA and RossiXTE two weeks after the outburst (Orio et al. 1999a, Orio et al. 2001b, Mukai & Ishida 1999, 2001) as a hard X-ray source (with plasma temperature kT≃7 keV). It cooled rapidly in the first two months after outburst to kT≃2.4 keV (Mukai & Ishida 2001). In a recent paper (Orio et al. 2001b; hereafter, Paper I) we attributed the initial X- ray emission to shocks in the nebula. Since the initially very large intrinsic absorption of the ejected nebula was thinning out, the equivalent N(H) decreased rapidly (Mukai & Ishida 2001). Thus, in the second BeppoSAX observation we hoped to detect the super-soft X-ray emission with the LECS and derive useful information on the nature of the white dwarf. 2 M. Orio et al. The peak temperature of the hot white dwarf remnant, and the absorption edges that indicate the underlying chemical composition, are extremely important in order to constrain the physical models. In addition to this, the length of the constant bolometric luminosity phase is an essential parame- ter in order to understand whether the nova retains accreted mass after the outburst and is therefore a candidate type Ia supernova (or, even, a candidate neutron star formed by ac- cretion induced collapse). 2 THE SUPERSOFT X-RAY SOURCE In the second BeppoSAX observation on 1999 November 23, the actual length of the LECS exposure was 12.4 ksec, over about a total 16 hours in 11 time intervals, lasting for t≤1200 sec each. The MECS exposure time was 25.9 ksec during the same 16 hours. The count rate measured with the LECS was extremely high, 3.4540±0.0021 cts s−1 in the 0.1-4.0 keV range, due to the emergence of supersoft X-ray emission (Orio et al. 1999b). This high count rate was not unexpected, and it would be equivalent (assuming for sim- plicity a blackbody at 40 eV and N(H)=2 × 1021 cm−2) to ≈50 cts s−1 with the ROSAT PSPC. A count rate of 75 cts s−1 was measured for V1974 Cyg at maximum with the PSPC (Krautter et al. 1996). In the 0.8-10.0 keV range the count rate was only 0.1030±0.0038, consistent with the MECS count rate 0.0454±0.0015 cts s−1 (more than a fac- tor of 3 lower than in June of 1999). There was no PDS detection with a 2σ upper limit of 0.080 cts s−1 in the 15-50 keV range. We have already discussed the evolution of the hard X-ray emission (Paper I). We found that more than one component was necessary to fit the LECS spectrum, how- ever we also concluded that there was no component with a higher plasma temperature than ≃1 keV, that N(H) was consistent with the interstellar value and that the supersoft portion of the flux was dominant. Remarkably, the supersoft X-ray flux (0.1-0.7 keV) was variable. No significant variability was detected at higher en- ergy. Overall, there was irregular flickering with time scales of minutes, and as Fig. 1 shows, in the 9th observation (af- ter ≈13 hours from the beginning of the exposures) the background-subtracted count rate decreased dramatically, by approximately a factor of 2. This low state lasted during the whole LECS coverage of 15 minutes, spaced about 5000 seconds from two observation in which the average count rate was twice higher. Close to the end of the ≈16 hours of intermittent observations, the count rate decreased again in the few minutes (see bottom panel of Fig. 1). If we missed other episodes of this type, they must have been shorter than the 4000 -- 5000 seconds that elapsed between the obser- vations. One possible explanation for the sudden decrease in count rate is of course that a dense clump intervened along the line of sight. However, the spectrum was definitely super- soft during the whole observation and became slightly softer during the dip. Instead, the additional absorption of a thick clump would absorb the softer portion of the spectrum more and produce an apparently "harder" spectrum. Is this phenomenon linked with orbital variability? Given the observed periodicity at optical wavelengths (Bos et al. 2001) the orbital period of V382 Vel is likely to be 3.5 hours. The time that elapsed between minima in the last observation span was a little over 3 hours. We folded the LECS lightcurve with the optical modulation period and even if ≈70% of the phase was covered, we could not detect any modulation in supersoft X-ray flux. We also note that the semi-amplitude of the optical modulation is only 0.02 mag, while the X-ray count rate has a large variation. The time scale of the phenomenon does not give very significant upper limits on the size of the obscuring region: the upper limit on the size of an obscuring clump in the nebula, assumed to be moving at v=4000 km s−1 for 5000 seconds, is 2 ×1012 cm (a small fraction of the nova shell radius at this epoch). Assuming instead some other type of phenomenon, connected with the possible orbital period, an upper limit on the size is obtained assuming the speed of light for the obscuring source: 1.5 ×1014 cm. This variability is a truly puzzling phenomenon which we do not fully understand. We note that also V1494 Aql (N Aql 1999 no.2), which was observed with Chandra and also detected as a supersoft X-ray source, showed time variability in supersoft X-rays: a flare (with increase of flux by a factor 6) that lasted for about 15 minutes and pulsations every 42 minutes (Starrfield et al. 2001). Even for this nova, the supersoft X-ray variability time scale was very short. 3 SPECTRAL ANALYSIS AND INTERPRETATION In Paper I we made the working hypothesis that the super- soft flux observed in November 1999 was entirely due to the central hot white dwarf remnant. Neglecting the LECS flux below 0.8 keV we simultaneously fitted the LECS and MECS spectra with a mekal model of thermal plasma (included in the software package XSPEC, see Arnaud et al. 1986) with parameters: N(H)≃ 2 × 1021 cm−2, kT≃700 eV, and un- absorbed flux ≃ 10−12 erg cm−2 s−1 (the reduced χ 2 was ≈1.2). In this work we analyse instead the properties of the supersoft X ray flux. The spectral distribution observed for V382 Vel, shown in Fig.2 in the range 0.1-1.1 keV, is strik- ingly different from the one observed with the BeppoSAX LECS for three non-nova supersoft X-ray sources: Cal 87 (Parmar et al. 1997), Cal 88 (Parmar et al. 1987) and RX J0925.7-4758 (Hartmann et al. 1999). The LECS spectra are shown in Fig. 7 of Hartmann et al. (1999) and Fig. 3 of Par- mar et al. (1997). V382 Vel is much more luminous than the other observed sources, it appears to have additional, harder spectral components and above all it is very luminous also in the very soft range, 0.1-0.2 keV. It is the only object of this kind ever studied with the LECS. We tried to fit the whole LECS and MECS spectra in the 0.1-10 keV range adding the mekal model used for the hard flux to a) a blackbody, b) a blackbody with absorption edges (included in XSPEC), and c) more detailed model at- mospheres (see Hartmann & Heise 1997, Hartmann et al. 1999). The latter models were used to fit the non-nova su- persoft X-ray sources and very reasonable results were ob- tained. We expected that the blackbody would not give a perfect fit (since white dwarf atmospheres resemble black- bodies only in first approximation) but we thought that it would provide an approximate estimate of the luminosity, of the range of effective temperatures, and indicate whether absorption edges must be included. Adding absorption edges c(cid:13) 1994 RAS, MNRAS 000, 1 -- ?? The X-ray emission from Nova V382 Velorum:II. The super-soft component observed with BeppoSAX 3 Table 1. Spectral fit parameters obtained applying two NLTE models with log(g)=8.5 and different abundances (25% the cosmic value, NLTE-1, and LMC-like depleted, NLTE-2), plus MEKAL and lines, to the BeppoSAX LECS and MECS spectra of Nova Vel observed in 1999 November. The absorption, N(H), is in units of 1022 atom cm−2. Teff is the WD atmospheric temperature, kT the MEKAL 2, dof the number of plasma temperature (of the ejecta), Fa is the absorbed flux and Fx the unabsorbed flux, χ degrees of freedom. 2/dof is the reduced χ Model N(H) Teff (×105 K) kT (keV) Fa Fx 2/dof χ A: NLTE-1+MEKAL B: NLTE-1+MEKAL+L(0.449 keV) C: NLTE-1+MEKAL+L(0.243 keV+0.449 keV +6.4 keV) D: NLTE-2+MEKAL E: NLTE-2+MEKAL+L(0.453 keV) F: NLTE-2+MEKAL+L(0.491 keV+6.52 keV) G: NLTE/CNO+MEKAL H: NLTE/NeOMg+MEKAL 1.38 2.03 2.07 1.56 1.98 1.98 1.68 1.69 5.6 4.0 4.0 5.3 5.2 5.2 4.8 5.0 0.896 0.801 0.836 0.069 0.826 0.813 0.063 0.068 1.6 × 10−9 6.1× 10−6 1.6 × 10−9 1.2 × 10−6 4.53 1.70 1.67 12.66 1.57 1.62 15.81 10.73 dof 105 102 96 122 117 112 106 105 to the black-body (the most likely seemed C vi at 0.49 eV, but we also tried N vi at 0.55 keV and 0.67 keV, O viii at 0.87 keV) did not improve the fit. We also found that a reasonable fit to the supersoft por- tion of the spectrum could not be obtained with any of the atmospheric models we tested, in Local Thermodynamical Equilibrium (hereafter, LTE), and in NON-LTE (hereafter, NLTE), developed by Heise et al. (1994), Hartmann & Heise (1997) and Hartmann et al. (1999). For the NLTE case, we tested four small grids of models with log(g) between 8 and 9, with four different set of abundances. The first set had cos- mic abundances (developed primarily for galactic sources, hereafter NLTE-1 models), in the second set of abundances the abundances were depleted to 0.25 the solar value (de- veloped for LMC sources, here after called NLTE-2 mod- els), and in the third and fourth group of models the abun- dances were enhanced in C, N and O and in Ne, O and Mg, respectively. These models include line opacities. The two grids with enhanced abundances are still unpublished (Hart- mann 2002, private communication). With either models in LTE and NLTE we could not obtain a good fit by, for in- stance, decreasing log(g) and "tuning" the MEKAL compo- nent accordingly. In Table 1 we give the best fit parameters the NON-LTE model atmospheres with log(g)=8.5 and a MEKAL bremsstrahlung component. The value log(g)=8.5 2, but it is smaller does not give an acceptable value of χ than for other values of log(g). The only reason for testing also the NLTE-2 LMC-type models was that we could not obtain a reasonable fit with the other available models and wanted to experiment with all the available grids. We con- cluded that no model is acceptable: with the best NLTE-1 2=4.5 for 105 de- model, model A in the Table, we obtain χ grees of freedom. The fit to the data, shown in the upper part of Fig. 3, predicts counts at energies higher than 200 eV that differ from those observed by more than 5%, which is the likely uncertainty in the knowledge of the LECS spec- tral response. Apart from what can be interpreted as an iron line at ≈6.5 keV, at low energies (<0.2 keV) the predicted counts exceed those observed by at least 100% (see Fig.2). However, the uncertainty in the low-energy LECS response in this range is estimated not to exceed ≃20% (Parmar et al. 1997). In a Chandra HRC-S+LETG observation of V382 Vel done in 2000 March by one of us, S. Starrfield, the HRC- c(cid:13) 1994 RAS, MNRAS 000, 1 -- ?? S+LETG spectrum, to be described in detail in a forthcom- ing paper (Starrfield et al. 2002) is very different from an- other nova which also appeared as a supersoft X-ray source, V1494 Aql (N Aql 1999 no.2, see Starrfield et al. 2001). The main difference is the lack of conspicuous continuum for V382 Vel in March 2000: it was instead an emission line spectrum, with many high ionization emission lines in the supersoft range. These lines presumably have an origin in the ejected nebula. The BeppoSAX instruments would have detected a blend of such lines as a featureless "pseudocontin- uum". Emission lines in the supersoft range may well have existed even at the earlier epoch of our BeppoSAX obser- vations. These lines may have been due to shock ionization in the ejecta, rather than to photoionization by the central source. Even if the spectral resolution of the BeppoSAX LECS and MECS is not sufficient to detect lines, we tried to de- termine which emission lines may explain the observed X- ray spectrum and what constraints we could derive on the level of the continuum. In model C of Table 1 we added not only the bremsstrahlung continuum at the harder energy (kT=0.84 keV) and a line at ≈6.5 keV (which is necessary but is not in the range where the bulk of the flux is emitted), but also one or more "softer" spectral features in emission. 2=1.6 -- 1.7 As an experiment, we obtained best fits with χ adding such lines to both NLTE-1 and NLTE-2 models with log(g)=8.5 (see Table 1). For the NLTE-1 model, in the best fit the added gaussian feature has to be at 449 keV. It could probably be C vi (perhaps a blend of the C vi triplet at 27- 28 A). We tried to add additional lines and improved the fit only marginally, although another emission line at 243 eV may be present (probably the Fe xv line at 50.5 A), and a narrow iron line at ≈6.4 keV is needed to explain the excess at this energy (Model C). This fit is shown in Fig. 3 (lower panel). With the NLTE-2 model we obtained model F with 2=1.62 adding a line (perhaps N vi) at ≈490 keV, and χ again an iron line for the excess at ≈6.5 keV. The total un- absorbed flux in the lines in these two models is a negligible fraction of the total bolometric flux, less than 1%. However, the flux in the line at ≈450 or 490 keV would be about 30% of the absorbed flux in the BeppoSAX LECS range. In model C, the bolometric luminosity at a distance of 2 kpc (Della Valle et al. 1999) is 6.8 × 1038 erg s−1. This value is 4 M. Orio et al. higher, but not much in excess of the model predictions for a ≈ 1 M⊙ WD emitting at Eddington luminosity. The main problem in determining the white dwarf pa- rameters accurately is the excess at low energy (kT<200 eV), which cannot be fit with one or more narrow lines. Around 150 eV several transition exist due to Fe, Si, Mg, and Ni, that could produce an intricate pattern, must be heavily absorbed, and cannot be resolved with the spec- tral resolution of the BeppoSAX LECS. By the time this nova was observed with the Chandra LETG, it had be- come much less luminous and the spectral structure had def- initely changed. We only remark that a complicated multi- temperature structure most likely existed in the ejecta, and that continuum and emission lines with different origins (white dwarf for the first and nebular for the latter) can explain the complicated spectrum we detected with the Bep- poSAX LECS and MECS. We rule out that our determina- tion of effective temperature and gravity of the white dwarf can be accurate if nebular lines overlap with the white dwarf continuum. It is only clear that at this stage the atmosphere of the central source was still the dominating component of the X-ray flux. 4 CONCLUSIONS The observation of V383 Vel 6 months after the outburst revealed a very luminous supersoft X-ray source, compara- ble in luminosity only with V1974 Cyg at maximum. The details of this observation raise new, unexpected and very interesting, questions. We observed irregular variability on a time scale of minutes in the supersoft flux of V382 Vel. The lack of energy dependence of the variability measured below 0.7 keV seems to rule out the ejection of an obscur- ing clump, yet other considerations seems to rule out orbital variability as well. Moreover, the BeppoSAX spectrum of V382 Vel in November 1999 appears much more complex than the ex- pected thermal continuum of a hot white dwarf plus a resid- ual thermal component from the nebula. We cannot justify this spectrum without invoking emission lines in the super- soft range (which were indeed observed shortly after this observation with Chandra), so we suggest that the observed "supersoft X-ray source" in V382 Vel is characterized by unresolved narrow emission lines superimposed on the at- mospheric continuum. We found that even a contribution of the lines of less than 1% to the total bolometric flux can significantly change the shape of the stellar continuum, and make the task of determining white dwarf temperature and effective gravity impossible with the resolution of the Bep- poSAX instruments. However, we conclude that the bulk of the X-ray flux was still due to the atmospheric continuum and not to lines at this epoch, unlike in the later observation performed with the Chandra LETG (Starrfield et al. 2002). Emission from classical novae in outburst can be quite complex and different from one nova to another. The X-ray spectra of N LMC 1995 (Orio & Greiner 1999) and of the recurrent nova U Sco (Kahabka et al. 1999) could be fit well with model atmospheres, although U Sco required also a nebular component at higher energy. We caution however, that the spectral structure may be as complex as the one s / t n u o C s / s t n u o C 3 2.5 2 1.5 1 0.5 46100 46200 46300 46400 46500 46600 46700 46800 46900 47000 Time (seconds from begining of observations) 5 4.5 4 3.5 3 2.5 2 1.5 1 57650 57700 57750 57800 57850 57900 Time (seconds from begining of observations) Figure 1. The LECS lightcurve in the range 0.1-0.7 keV with time bins of 100 s (top) and the 16 s binned lightcurve during the third before last (second plot from the top) and the last (bottom) of the observations done during the 16 hours from beginning to end of LECS exposure. observed for Cal 83 by Paerels et al. (2001), when observed with higher resolution. The situation was more complex for V1974 Cyg, where a model atmosphere and a hotter, thermal component were necessary to fit the spectrum. The relative importance of the two components seemed to vary in each observation (see Balman et al. 1998) and the interplay between them was rather complicated, also due to the lower energy resolution of the PSPC compared to the LECS (a factor 2.4 less) and limited energy range of the PSPC (which could not cover the harder component well, at least at the beginning). We wonder whether the spectrum of V1974 Cyg also had super- imposed nebular emission lines that made it appear hotter than it was, because a lower effective temperature (≤ 20 eV) of the post-nova white dwarf atmosphere may explain a puzzling fact. An ionization nebula was detected only in c(cid:13) 1994 RAS, MNRAS 000, 1 -- ?? The X-ray emission from Nova V382 Velorum:II. The super-soft component observed with BeppoSAX 5 different velocity colliding into the initial radiation driven wind? The nova theory must become more detailed and re- fined once the X-ray spectrum is known in detail for a sta- tistically meaningful sample of objects. The gratings in the new X-ray observatories are opening new and exciting pos- sibilities for nova studies. ACKNOWLEDGMENTS We thank L. Piro, BeppoSAX Mission Scientist, for support and W. Hartmann for providing the atmospheric models he developed. M. Orio is grateful also to another colleague, C. Markwardt, for computer support and useful advice. Finally, we thank the referee, Koji Mukai. This research has been supported by the Italian Space Agency (ASI) and by the UW College of Letters & Sciences. S. Starrfield gratefully acknowledges NSF and Chandra support to ASU. REFERENCES Arnaud K.A., 1996, In: Astronomical Data Analysis Software Sys- tems V. ASP Conf. Series 101, G. Jacoby, J Barnes editors, 17 Balman S., Krautter J., Ogelman H., 1998, ApJ, 499, 395 Bearda H., Hartmann W., et al., 2002, preprint, astroph0202333 Boella G., et al., 1997, A&AS, 122, 327 Bos M., Retter A., McCormick J., Velthuis F., 2001, IAU Circ. 7610 Casalegno R. Orio M., et al., 2000, A&A, 361, 725 Della Valle M., Pasquini L., Williams R., 1999, IAU Circ. 7193 Frontera F., et al., 1991, Advances in Space Research, 11, 281 Hartmann H.W., Heise J., 1997, A&A, 322, 591 Hartmann H.W., Heise J., Kahabka P., Motch C., Parmar A.N., 1999, A&A 125, L346 Heise J., van Teeseling A., Kahabka P., 1994, A&A, 288, L45 Kahabka P., Hartmann H.W., Parmar A.N., Neguerela I., 1999, A&A, 347, 43 Krautter J., et al., 1996, ApJ, 456, 788 Mukai K., Ishida M., 1999, IAU Circ. 7205 Mukai K., Ishida M., 2001, ApJ, 551, 1024 Ogelman H., Orio M., Krautter J., Starrfield S., 1993, Nature, 361, 331 Orio M., Covington J., Ogelman H., 2001a, A&A, 373, 542 Orio M., Parmar A., Benjamin R., Amati L., Frontera F., Greiner J., Ogelman H., Mineo T., Starrfield S., Trussoni E., 2001b, MNRAS, 326, L13O Orio M., Greiner J. 1999, A&A, 344, L1 Orio M., Parmar A.N., Capalbi M., 1999b, IAU Circ. 7325 Orio M., Torroni V., Ricci R., 1999a, IAU Circ. 7196 Paerels F., et al., 2001, A&A, 365, L308 Parmar A.N., et al., 1997, A&AS, 122, 309 Platais I., et al., 2000, PASP, 112, 224 Seargent D.A.J., Pearce A., 1999, IAU Circ. 7177 Shore S.N., et al., 1999, IAU Circ. 7261 Starrfield S., et al., 2001, AAS, 198, 11.09 Starrfield S., et al., 2002, in preparation Figure 2. Spectra observed in the 0.1-8 keV range with the BeppoSAX LECS and MECS in November 1999 and (above) best fit with model A (MEKAL + NLTE model atmosphere for log(g)=8.5 and cosmic abundances) and (below) with model C, which includes also three superimposed emission lines, at 0.243 keV, 0.449 keV and 6.4 keV). The first fit is not acceptable, and this is due to in a large portion to structured residuals in the range 0.1-0.2 keV. The second fit is overall much better, but the resid- uals in the soft range are still large. The lower panel in each plot shows the residuals, as ratio of observed over predicted counts per energy bin. Hα for N Cyg 1992 (Casalegno et al. 2000) while other ion- ization lines (indicating a higher ionization potential) were not present in the nebula. We speculate therefore that in V1974 Cyg the white dwarf (Krautter et al. 1996, Balman et al. 1998) might have been cooler than it appeared by fit- ting the ROSAT PSPC spectrum with just a two component model. We note that even the spectrum of a non-nova super- soft X-ray source, the Galactic MR Vel, shows non- atmospheric emission lines, attributed to a wind from the source (Bearda et al. 2002). Instead, in classical novae in outburst emission lines in the soft X-ray range could be pro- duced by shock ionization within the nebula. Shocked, X-ray emitting material seems to be present since the beginning of the outburst (Krautter et al. 1996, Orio et al. 2001b). The BeppoSAX LECS and the ROSAT PSPC do not re- solve narrow emission lines. For novae, prominent nebular emission lines in the supersoft X-ray energy range indicate interesting possibilities, specially if they should be observed with Chandra or XMM-Newton in the future. We face new questions. Are these lines at times due to collisional excita- tion, do shocks occurs in the nova wind even many months after the outburst? Should we consider a line driven wind at c(cid:13) 1994 RAS, MNRAS 000, 1 -- ??
0709.1111
1
0709
2007-09-07T17:33:54
On the oscillations in Mercury's obliquity
[ "astro-ph" ]
One major objective of MESSENGER and BepiColombo spatial missions is to accurately measure Mercury's rotation and its obliquity in order to obtain constraints on internal structure of the planet. Which is the obliquity's dynamical behavior deriving from a complete spin-orbit motion of Mercury simultaneously integrated with planetary interactions? We have used our SONYR model integrating the spin-orbit N-body problem applied to the solar System (Sun and planets). For lack of current accurate observations or ephemerides of Mercury's rotation, and therefore for lack of valid initial conditions for a numerical integration, we have built an original method for finding the libration center of the spin-orbit system and, as a consequence, for avoiding arbitrary amplitudes in librations of the spin-orbit motion as well as in Mercury's obliquity. The method has been carried out in two cases: (1) the spin-orbit motion of Mercury in the 2-body problem case (Sun-Mercury) where an uniform precession of the Keplerian orbital plane is kinematically added at a fixed inclination (S2K case), (2) the spin-orbit motion of Mercury in the N-body problem case (Sun and planets) (Sn case). We find that the remaining amplitude of the oscillations in the Sn case is one order of magnitude larger than in the S2K case, namely 4 versus 0.4 arcseconds (peak-to-peak). The mean obliquity is also larger, namely 1.98 versus 1.80 arcminutes, for a difference of 10.8 arcseconds. These theoretical results are in a good agreement with recent radar observations but it is not excluded that it should be possible to push farther the convergence process by drawing nearer still more precisely to the libration center.
astro-ph
astro-ph
On the oscillations in Mercury's obliquity E. Bois a and N. Rambaux b,c aObservatoire de la Cote d'Azur, UMR CNRS Cassiop´ee, B.P. 4229, F-06304 Nice Cedex 4, France bRoyal Observatory of Belgium, 3 Avenue Circulaire, B-1180 Brussels, Belgium cDepart. of Mathematics, Facult´es Univ. N.D. de la Paix, 8 Rempart de la Vierge, B-5000 Namur, Belgium Email address : [email protected] Email address : [email protected] Telephone Number : + 33 [0]4 92 00 30 20 (E. Bois) Telephone Number : +32 [0]2 373 67 33 (N. Rambaux) Number of manuscript pages: 30 Number of manuscript tables: 3 Number of manuscript figures: 8 7 0 0 2 p e S 7 ] h p - o r t s a [ 1 v 1 1 1 1 . 9 0 7 0 : v i X r a Preprint submitted to Elsevier Science 31 May 2018 Eric Bois Observatoire de la Cote d'Azur, UMR CNRS Cassiop´ee, B.P. 4229, F-06304 Nice Cedex 4, France Telephone Number : + 33 [0]4 92 00 30 20 Fax Number : + 33 [0]4 92 00 31 18 Email address : [email protected] Running title: Librations in Mercury's obliquity 2 Abstract Mercury's capture into the 3:2 spin-orbit resonance can be explained as a result of its chaotic orbital dynamics. One major objective of MESSENGER and BepiColombo spatial missions is to accurately measure Mercury's rotation and its obliquity in order to obtain constraints on internal structure of the planet. Analytical approaches at the first order level using the Cassini state assumptions give the obliquity constant or quasi-constant. Which is the obliquity's dynamical behavior deriving from a complete spin-orbit motion of Mercury simultaneously integrated with planetary interactions? We have used our SONYR model (acronym of Spin-Orbit N -bodY Relativistic model) integrating the spin-orbit N -body problem applied to the solar System (Sun and planets). For lack of current accurate observations or ephemerides of Mercury's rotation, and therefore for lack of valid initial conditions for a numerical integration, we have built an original method for finding the libration center of the spin-orbit system and, as a consequence, for avoiding arbitrary amplitudes in librations of the spin-orbit motion as well as in Mercury's obliquity. The method has been carried out in two cases: (1) the spin-orbit motion of Mercury in the 2-body problem case (Sun-Mercury) where an uniform precession of the Keplerian orbital plane is kinematically added at a fixed inclination (S2K case), (2) the spin-orbit motion of Mercury in the N -body problem case (Sun and planets) (Sn case). We find that the remaining amplitude of the oscillations in the Sn case is one order of magnitude larger than in the S2K case, namely 4 versus 0.4 arcseconds (peak-to- peak). The mean obliquity is also larger, namely 1.98 versus 1.80 arcminutes, for a difference of 10.8 arcseconds. These theoretical results are in a good agreement with recent radar observations but it is not excluded that it should be possible to push farther the convergence process by drawing nearer still more precisely to the libration center. We note that the dynamically driven spin precession, which occurs when the planetary interactions are included, is more complex than the purely kinematic case. Nevertheless, in such a N -body problem, we find that the 3:2 spin- orbit resonance is really combined to a synchronism where the spin and orbit poles on average precess at the same rate while the orbit inclination and the spin axis orientation on average decrease at the same rate. As a consequence and whether it would turn out that there exists an irreducible minimum of the oscillation amplitude, quasi-periodic oscillations found in Mercury's obliquity should be to geometrically understood as librations related to these synchronisms that both follow a Cassini state. Whatever the open question on the minimal amplitude in the obliquity's oscillations and in spite of the planetary interactions indirectly acting by the solar torque on Mercury's rotation, Mercury remains therefore in a stable equilibrium state that proceeds from a 2-body Cassini state. Key words: Mercury, rotational dynamics, spin-orbit resonance, celestial mechanics Email addresses: [email protected] (E. Bois), [email protected] (N. 3 1 Introduction From radar measurements of the rotation of Mercury obtained by Pettengill & Dyce (1965), Colombo (1965) showed the non-synchronous rotation of Mer- cury and its 3:2 spin-orbit state (the rotational and orbital periods are 56.646 and 87.969 days respectively). The stability of the 3:2 spin-orbit resonance de- pends on a significant orbital eccentricity and the fact that the averaged tidal torque trying to slow the planet further is considerably less than the maxi- mum averaged torque acting on the permanent axial asymmetry that causes the planet spin to librate about the resonant 3:2 value. Moreover, Correia & Laskar (2004) have shown more recently that Mercury's capture into the 3:2 resonance can be explained as a result of its chaotic orbital dynamics. The upcoming missions, MESSENGER (Solomon et al. 2001) and BepiCo- lombo, with onboard instrumentations capable of measuring the rotational parameters of Mercury (see e.g. Milani et al. 2001), stimulate the objective to achieve a very accurate theory of Mercury's rotation. One aim of these missions is to determine the existence or not of a liquid core by measuring accurately both the rotation and the gravity field of Mercury. The current method for obtaining constraints on the state and structure of Mercury's core is based on the assumptions introduced by Peale in 1972 and revisited in 2002 (Peale et al. 2002). By measuring the amplitude of the 88-day libration in longitude, called ϕ hereafter, the obliquity ¯η (a constant derived from the Cassini state assumption for Mercury's rotation and playing the role of a mean obliquity), and the degree-two gravitational coefficients (2¸0 and 2. 2), it is possible to determine the radius of the expected fluid core. Let A, B, and C be the principal moments of inertia of Mercury, Cm the mantle's moment of inertia, M the mass of Mercury, and R its equatorial radius. The information on the core size is inferred from the following relation (Peale et al. 2002) : (cid:18) Cm B − A(cid:19)(cid:18)B − A MR2 (cid:19) MR2 C ! = Cm C ≤ 1 (1) where a ratio of Cm/C equal to 1 should indicate a core firmly coupled to the mantle and most likely solid. In the other hand, if the interior of Mercury is partly fluid, the ratio is smaller than 1. The objective of the rotation parameter measurements consists in restraining each factor of the expression (1). The first factor is inversely proportional to the ϕ rotation angle (Peale et al. 2002). The second factor in (1) is equal to 42. 2 while the third factor is related to ¯η by the following relation (Wu et al. 1995) : MR2 C = µ n sin (I + ¯η) sin ¯η [(1 + cos ¯η) G201C22 − (cos ¯η) G210C20] (2) Rambaux). 4 2e − 123 where n is the mean orbital motion of Mercury and µ a constant of uniform precession of the orbit pole. G201 and G210 are the eccentricity functions of Kaula (Kaula, 1966). G201 = 7 16 e3 represent the first two terms in a series expansion where e is the eccentricity while G210 = (1 − e2)− 2 . The formula (2) expresses the Cassini state of Mercury (Colombo 1966; Peale 1969) where I is the inclination of the orbit pole of Mercury relative to the Laplace plane pole around which the orbit pole precesses at an uniform rate. Let us recall that in a Cassini state, the orbital and rotational parameters are indeed matched in such a way that the spin pole, the orbit pole, and the Laplacian pole remain coplanar while the spin and orbital poles on average precess at the same rate. 3 Table 1 is here The rotational motion of Mercury is mainly characterized by a 3:2 spin-orbit resonance and a spin-orbit synchronism where axes of precession cones, spin and orbital, on average precess at the same rate. In the 2-body problem (Sun and Mercury), by using the 3:2 spin-orbit resonance and a few corollary ap- proximations in the classic Euler equations of the 3-D rotational motion of Mercury disturbed by the solar torque, one may get two characteristic fre- quencies, namely the one in longitude Φ, which is solution of the third Eule- rian equation and the other in latitude of the body Ψ, which is solution of the two first Eulerian equations. A Hamiltonian approach of the rotational mo- tion of Mercury has been expanded by D'Hoedt & Lemaıtre (2004). The two resulting frequencies Φ and Ψ, usually called proper frequencies, calculated by this analytical approach and those obtained with our SONYR model are in a good agreement (see Table 1). By the present paper, we have also improved the internal accuracy of our values (see Table 1). Table 2 is here Let us call η the instantaneous obliquity of Mercury, defined as the angle bet- ween Mercury's Oz spin axis and the normal to its orbital plane; let < η > be its mean value. Until the Radar observations recently obtained by Margot et al. (2006), 1 the value of Mercury's mean obliquity was not truly constrai- ned by measurements; many values were available in literature from zero to 7 arcminutes (amin hereafter). On the one hand, from the formula (2) computed with realistic Mercury's parameters, ¯η can not be equal to zero. On the other hand, 7 amin given by Wu et al. (1995) seemed too high for Mercury. In our previous paper (Rambaux & Bois 2004), using (2) we have estimated a range of ¯η from 1.33 to 2.65 amin (for I ∈ [5, 10] deg, e ∈ [0.11, 0.24], with µ = 2π/Π and n = 2π/Pλ where values of the Π orbital precession period and the Pλ orbital period are given in Table 2). However by using our SONYR model including the full spin-orbit motion of Mercury, for the plausible values of 1 published at the very time when we are finishing the revised paper. 5 parameters we had chosen, the mean value of η is close to 1.66 amin. In the present paper, by improving our method, we give a more relevant value for < η > that proves to be in a very good agreement with the observations. We obtain the dynamical behavior of the hermean obliquity by the way of the following relation : cos η = cos i cos θ + sin i sin θ cos(Ω − ψ) (3) where the spin-orbit variables are computed with SONYR. i and Ω are respec- tively the inclination and the longitude of the ascending node of Mercury's orbital plane relative to the ecliptic plane. The angles ψ and θ are the preces- sion and nutation angles of the Eulerian sequence 3-1-3, while the third angle ϕ is called the proper rotation. This angular sequence is used to describe the evolution of the body-fixed axis system Oxyz (centered on the Mercury's cen- ter of mass) relative to a local reference system, dynamically non-rotating but locally transported with the translational motion of the body. Let OX1Y1Z1 be the local reference system for the Mercury's rotation (see complete definitions of reference systems in Section 3). In our previous work as well as in the present one, using our SONYR model, we have numerically integrated the complete equations of the N -body problem (8 planets and the Sun) including the full spin-orbit motion of Mercury. More precisely, in our previous paper (Rambaux & Bois 2004), we presented (i) an accurate theory of Mercury's spin-orbit motion, (ii) the proper frequencies related to the spin-orbit resonance, (iii) an analysis of main librations, and (iv) a first attempt for characterizing the dynamical behavior of Mercury's obliquity. The present paper is wholly devoted to stress and highlight the fourth point. For lack of current accurate observations or ephemerides of Mercury's rotation, and therefore for lack of valid initial conditions for a numerical integration, we have built an original method for finding the libration center of the spin-orbit system and as a consequence for avoiding arbitrary amplitudes in librations of the spin-orbit motion as well as in Mercury's obliquity. Initial conditions of the rotational and orbital modes have indeed to be consistent between themselves with respect to a same spin-orbit reference system. As a consequence, using the SONYR model, the method requires in a first step to build mean initial conditions verifying geometrical conditions of a Cassini state for Mercury, that is to say a spin-orbit equilibrium state. In a second step, these mean initial conditions are fitted in order to place the spin-orbit system at its center of libration. The method has been carried out in two cases: (1) the spin- orbit motion of Mercury in the 2-body problem case (Sun-Mercury) where an uniform precession of the Keplerian orbital plane is kinematically added at a fixed inclination, (2) the spin-orbit motion of Mercury in the N-body problem (Sun and planets). With the spin-orbit system located at its center of libration in both cases above, the reference system of the first case is put at the average of perturbations of the second case. As a consequence, we make not only in 6 evidence the strict dynamical behavior of Mercury's obliquity in both cases but may compare the respective oscillations. Our theoretical results are in a good agreement with recent radar observa- tions. However, it is not excluded that it should be possible to push farther the convergence process by drawing nearer still better to the libration center. In the present framework, we note that the dynamically driven spin precession, which occurs when the planetary interactions are included, is geometrically more complex than the purely kinematic case. In particular, we highlight that in the full spin-orbit motion of Mercury, the asynchronous rotation of the planet is really combined with a synchronism in precession variables as well as a second one between secular changes of the orbit inclination and the spin axis orientation. As a consequence and supposing that it could exist an irre- ducible minimum of the oscillation amplitude, we show how the quasi-periodic oscillations found in Mercury's obliquity could be geometrically understood. We discuss briefly the question to know whether the minimal amplitude has to theoretically be of zero value or not. In the meantime of such a forthcoming work, we describe Mercury's rotation and the resulting equilibrium state that proceeds from a 2-body Cassini state. 2 The Spin-Orbit N -bodY Relativistic model For obtaining the real motion of Mercury, we have used our model of solar System integration including the coupled spin-orbit motion of the Moon. This model (built by Bois, Journet, & Vokrouhlick´y and called BJV), expanded in a relativistic framework, had been previously built in accordance with the requirements of the Lunar Laser Ranging observational accuracy (Bois 2000; Bois & Vokrouhlick´y 1995). We extended the BJV model by generalizing the Moon's spin-orbit coupling to the other terrestrial planets (Mercury, Venus, the Earth, and Mars). The model is at present called SONYR (acronym of Spin-Orbit N -BodY Relativistic model). As a consequence, the SONYR model gives an accurate simultaneous integration of the spin-orbit motion of Mer- cury. The approach of the BJV and SONYR models consists in integrating the N -body problem (including translational and rotational motions) on the basis of the gravitation description given by the general relativity. The equations have been developed in the DSX formalism presented in a series of papers by Damour, Soffel, & Xu (1991, 1992, 1993). For purposes of celestial mechanics, to our knowledge, it is the most suitable formulation of the post-Newtonian (PN) theory of motion for a system of N arbitrary extended, weakly self- gravitating, rotating and deformable bodies in mutual interactions. The DSX formalism, derived from the first PN approximation level, includes then the translational motions of the bodies as well as their rotational motions with respect to locally transported frames with the bodies. To each body is then associated a local reference system, centered on its center of mass and dynam- ically non-rotating. These local reference systems are freely falling in the local 7 gravity field (the Sun) and are referred to a global reference system O′X ′Y ′Z ′ for the complete N-body problem. A description of the SONYR model is given in our previous paper (Rambaux & Bois 2004). The principle of the model allows the analysis of Mercury's spin- orbit motion and the identification of internal mechanisms such as the direct and indirect planetary interactions. In addition, it gives also the interrelations between the parameters involved in Mercury's dynamical figure and the planet librations. 3 Reference systems and their relations O′X ′Y ′Z ′ is centered on the Solar System barycenter, the O′X ′Y ′ plane is parallel to the J2000 ecliptic plane. The X ′-axis is oriented to the equinox J2000. Let OXY be the mean orbital plane of Mercury and OZ the mean orbital pole. Let us call P the instantaneous orbital plane of Mercury. The rotational motion of Mercury is evaluated from a body-fixed coordinate axis system Oxyz centered on Mercury's center of mass relative to a local dy- namically non-rotating reference system, OX1Y1Z1. The Ox, Oy, Oz axes are defined by the three principal axes of inertia a, b, and c of the planet according to the best fitting ellipsoid such as a > b > c. The corresponding principal moments of inertia are then ordered as follows : A < B < C. We used the Euler angles ψ, θ, ϕ related to the 3-1-3 angular sequence (3 represents a ro- tation around a Z-axis and 1, around an X-axis) to describe the evolution of the body-fixed axes Oxyz with respect to axes of the local reference system OX1Y1Z1. These angles are defined as follows : ψ is the precession angle of the Oz polar axis around the OZ1 reference axis, θ is the nutation angle represent- ing the inclination of Oz with respect to OZ1, and ϕ is the rotation around Oz and conventionally understood as the rotation of the largest rotational energy. This angle is generally called the proper rotation. The Oz principal axis around which is applied the proper rotation is called the axis of figure and defines the North pole of the rotation (Bois 1995). The proper rotation ϕ is around the smallest axis of inertia; it is called hereafter the spin angle and Oz the spin axis. Figure 1 is here In SONYR, building up the reference frames from the reference systems is re- lated to the choice of the reference frames transported by the initial conditions (IC hereafter). This is of practical use when we have valid IC (for instance coming from ephemerides or derived from observations). In such conditions, the local reference frame for the rotation of a planet may be the one of the ephemeris. It is not actually the case for Mercury. True IC of Mercury's ro- tation are unknown (with respect to any reference system). Moreover, ad hoc attempts for determining instantaneous IC of the rotation produce a shift with 8 those of the orbit and therefore a shift of the libration center of the spin-orbit resonance. Inconsistency between rotation's and orbit's IC drives necessarily to arbitrary amplitudes and phases in the spin-orbit librations. As a conse- quence, we have built (in Section 4) mean initial conditions (MIC hereafter) coming from geometrical conditions of the mean spin-orbit motion of Mercury. The latter is related to (i) the 3:2 spin-orbit resonance in cyclic variables (ϕ and λ such as < dϕ/dt > = (3n/2) + < dω/dt >), (ii) the synchronism in precession variables (ψ and Ω), and (iii) a second synchronism in nutation and inclination variables (θ and i) (see Rambaux & Bois 2004). In a second step of the method (Section 5), the MIC are accurately corrected by taking into account a fit between the central axis of the precession cone described by the spin pole and the theoretical axis of reference. For these final MIC, the spin-orbit system is at its center of libration (Γ hereafter) without arbitrary amplitudes. Moreover, we consider two cases : (1) The spin-orbit motion of Mercury in the 2-body problem framework (Sun- Mercury) where a uniform precession, which we denote by Π, of the Keplerian orbital plane is kinematically added at a fixed inclination IK with respect to OX ′Y ′. Let us call S2K this construction bearing the geometrical Cassini state. Let us call PK the associated orbital plane (a Keplerian plane in this case), and LK the fixed plane of reference OX ′Y ′ (LK plays the role of a Laplace plane for PK) (see Fig. 1). (2) The spin-orbit motion of Mercury in the N-body problem framework (Sun and planets). Let us call Sn this case. Let us call Pn the associated mean orbital plane of Mercury. The instantaneous orbital plane P has an instantaneous inclination i relative to OX ′Y ′ while the inclination of Pn is < i >. Let us include the S2K construction into Sn by putting PK on Pn. As a consequence, IK = < i >. Over short periods, Pn remains equivalent to PK. The associated Laplace plane Ln is an instantaneous plane which follows the oscillations of P so that i remains quasi-constant relative to Ln. LK coıncides with the average of the oscillations of Ln, i.e. < Ln > ≡ LK (see Fig. 1). Our construction of reference systems verifies in the end the three following relations : Pn ≡ PK < i > ≡ IK < Ln > ≡ LK (4) which permit to accurately evaluate the obliquity's oscillations (i) not only with respect to the center of libration of the full spin-orbit motion of Mer- cury, but (ii) with respect to the included Cassini state. In other words, our construction is such as the mean Cassini state of the full spin-orbit motion of Mercury, on a given interval of time, coıncides with an exact Cassini state derived from the 2-body problem. Let us add that local dynamically non-rotating frames show in general relativ- ity a slow (de Sitter) rotation with respect to the kinematically non-rotating frames. As a consequence, the OX1Y1Z1 local reference frame for Mercury's rotation is affected by a slow precession of its axes transported with the trans- lational motion of Mercury. In the Earth's case, the de Sitter secular preces- 9 sion of the Earth reference frame is close to 1.9"/cy (see Fukushima 1991; Bizouard et al. 1992; Bois & Vokrouhlick´y 1995). Consequently, the real rota- tion of Mercury has not to be expressed in an inertial system fixed in space, but in a local dynamically non-rotating frame fallen down in the gravitational field of the Sun. However, in the framework of the present paper where only Newtonian-like terms of SONYR are activated, OX1Y1Z1 remains parallel to OX ′Y ′Z ′. 4 Building the mean initial conditions For lack of accurate observations or ephemerides of Mercury's rotation, and therefore for lack of valid IC for a numerical integration of the full spin-orbit motion of Mercury, a method is required for: (i) fitting the different parameters or constants involved in the spin-orbit of Mercury so that the SONYR model be self-consistent; (ii) removing arbitrary amplitudes in librations of Mercury's spin-orbit motion and therefore in its obliquity. In this section, we present the construction of MIC related to the spin-orbit coupling of Mercury. (1) We start from the geometry of the spin-orbit coupling problem defined in Goldreich & Peale (1966): the spin axis of Mercury is considered normal to its orbital plane and the orbit is assumed to be fixed and unvarying. At this level, for expressing these assumptions, as well as that the long axis is pointed toward the Sun at a perihelion position t0, analytical relations giving the corresponding set of initial conditions, namely C0, can be written as follows : (C0) ψ0 = Ω0 θ0 = I ϕ0 = ω0 + 3 2M0 dt!0 dψ = 0 dθ dt!0 = 0 dϕ dt!0 = 3 2 n   where ω0 is the mean argument of periastron, and M0 the mean anomaly with Ω0 = ω0 = M0 = 0 at t = t0. For such a Mercury's spin-orbit coupling based on a 2-body problem (with the asphericity factor α = q3(B − A)/C = 0.0187 and the eccentricity e = 0.206), we have plotted a Poincar´e section in Rambaux & Bois (2004, Fig. 2), making in evidence the quasi-periodic librations around the 3:2 rotation state. (2) In the S2K framework, the condition of the precession synchronism drives to the geometrical property < dψ/dt > = 2π/Π where Π is the constant of uniform precession. The principle of our method requires to obtain the value of Π from the Sn case such as 2π/Π = < dΩ/dt >. In the same way, the value of I = IK is obtained from the Sn case such as IK = < i >, and as a consequence the value of I is taken equal to < i > since the S2K case. The mean initial 10 conditions C2K for the S2K case are then written as follows : (C2K) ψ0 = Ω0 θ0 = I = IK ϕ0 = ω0 + 3 2M0 dψ dt !0 = 2π Π dθ dt!0 = 0 dϕ dt!0 = 3 2 n with Ω0, ω0, and M0 equal to zero at t = t0. (3) In the Sn framework, the orbit precession of the rotating body is dy- namically generated by the planetary interactions but its rate is nonuniform. However, 2π/Π has been rightly defined as < dΩ/dt >. Moreover, the 3:2 spin- orbit resonance in the Sn case involves the condition dϕ/dt = (2π/Λ) + (3n/2) where Λ is obtained such as 2π/Λ = < dω/dt >. In addition, as shown further, θ and i on average change at the same rate. This expresses a proximity to a Cassini state, which involves a third geometrical condition for building MIC, namely < dθ/dt > = < di/dt > = 2π/Θ. The mean initial conditions Cn for the Sn case are then written as follows :     (Cn) ψ0 = Ω0 θ0 = < i > ϕ0 = ω0 + 3 2M0 dt !0 dψ = 2π Π dθ dt!0 = 2π Θ dϕ dt!0 = 2π Λ + 3 2 n with Ω0, ω0, and M0 equal to zero at t = t0. The dynamical behaviors of θ and i have the same slopes (see Fig. 2, top panel). As a consequence, secular parts of θ and i decrease at the same rate equal to Θ. Although it is a question of a N-body problem, the spin keeps its proximity to a Cassini state during the slow change in i. But with this signature of the proximity to the Cassini state made in evidence in the Sn framework, < dθ/dt > = < di/dt > may be understood as a second synchronism between the respective inclinations of the orbit and the spin axis relative to a same reference system (OX ′Y ′Z ′). As a matter of fact we find that ξ = θ −i librates and rightly according to the Ψ proper frequency in latitude (see Fig. 2, bottom panel). Figure 2 is here 5 Fitting the mean initial conditions Let us call hereafter SONYR(Si, Ci) the integration of SONYR in the mode Si with the initial conditions Ci. Let us note that under the assumptions of the present paper, SONYR(Si, Ci) is a conservative dynamical system. Let S be the Oz spin pole of Mercury. S describes a large cone with oscillations of polhodie; the projection of the cone in the OX ′Y ′ reference plane gives a 11 large quasi-circle of center A (with oscillations). Due to slightly heterogeneous initial spin-orbit positions, A does not exactly coıncide with the projection of OZ ′. By setting the latter in A, the system is placed at its Γ libration center. This can be achieved by slightly moving OX ′Y ′Z ′ or in SONYR by fitting a little bit the MIC. But because of large periods of precession, building and fitting the large circles is a very heavy computational process. Nevertheless, the reference frame that satisfies the whole system in Γ is the one where all librations (related to small and large cones of precession) are at their respective right centers of libration. As a consequence, we have determined equations expressing the departure from Γ within smaller circles of libration. Let us note SX " and SY " the coordinates of S projected onto a reference plane OX"Y " rotating with the uniform motions Π and Θ; they are written as follows : SX " = sin [ψ − ( 2π SY " = − cos [ψ − ( 2π Π )t] sin [θ − ( 2π Θ )t] Π )t] sin [θ − ( 2π Θ )t] (5)   where 2π/Θ = 0 in the S2K case. For each mode of SONYR (S2K and Sn), S describes a small precession cone. Figure 3 gives the motion of S in the mode SONYR(S2K, C2K). Let us call Oζ the central axis of the small cone of precession whose coordinates are (ζ1, ζ2) in OX"Y ". Figure 3 is here The fitting equations are then written as follows : SX " = ζ1 SY " = ζ2   After substitutions in orbital elements alone, the system (6) becomes : (6) (7) IK + δIK = arcsin (qζ 2 Ω0 + δΩ0 = arctan (−ζ1/ζ2) 1 + ζ 2 2 )   where one gets δIK and δΩ0 as functions of IK, Ω0, ζ1, and ζ2. In the mode SONYR(S2K, C2K), we obtain then the two following fitted initial conditions : ψ0 = Ω0 + δΩ0 θ0 = IK + δIK (8) C2K becomes C ′ 2K and one gets on the same graphic frame as previously a little circle (like a point at this scale, see Fig. 4). The little circle expresses a considerable decreasing of the degree of opening of the cone by the above process. Figure 4 is here 12 The same principle is applied to the Sn case with an analogous result. Taking into account in Cn (becoming C ′ n) the two fitted initial conditions ψ0 and θ0 coming from (8) according to the same process, Figure 5 shows the significant decreasing of the degree of opening of the precession cone. Figure 5 is here Table 3 is here 6 The resulting obliquities In the S2K mode, computations of η give resulting obliquities before and after fitting the MIC at the libration center Γ, as shown in Figure 6a. Figure 6b shows a zoom of the obliquity purged of arbitrary amplitudes after the fit- ting process. However, η appears not constant but quasi-constant. In the S2K mode indeed, i.e. without planetary interactions, due to the gravitational so- lar torque acting on the dynamical figure of Mercury with three unequal axes of inertia, the equations of the rotational motion solely form a non-integrable system whose regular solutions are quasi-periodic librations. As a consequence, the η obliquity that depends on θ and ψ (see formula (3) while i = IK and Ω = Ω0 + (2π/Π)t where IK, Ω0 and Π are constant values) follows in part this behavior. Nevertheless, with such faint amplitudes (peak-to-peak 0.4 arc- seconds, 'as' hereafter), the obliquity may be reported quasi-constant in the S2K case, which is usually characteristic of a Cassini state derived from the 2-body problem (see Fig. 6b). After removing the amplitude of the high fre- quency fluctuations (a modulation of a beat frequency with a period of 295 years whose peak-to-peak amplitude varies between a maximum near 0.28 as and a minimum near 0.14 as), the peak-to-peak amplitude of the resulting long period modulation of 1066.4 years is only 0.12 as. Figure 6 is here In the same way, in the Sn mode, computations of η give resulting obliquities, before and after the fitting process, as shown in Figure 7a. Figure 7b shows a zoom of the obliquity after the fitting process and then purged of arbitrary amplitudes. In both modes, S2K and Sn, the period of the main oscillation is Ψ, namely the second proper frequency of the resonance. A frequency analysis gives 1066.81 years in the Sn case and 1066.44 years in the S2K one, let be a difference lesser than 0.4 years. In the Sn case, after removing the amplitude of the high frequency fluctuations (about 0.6 as), the peak-to-peak amplitude of the resulting long period modulation of 1066.8 years is 3.4 as. Figure 7 is here 13 Both sets of results, from S2K and Sn cases are globally consistent with those of Peale (2006) and implicitly with those of Yseboodt & Margot (2006). We note that the dynamically driven spin precession, which occurs when the plan- etary interactions are included, is geometrically more complex than the purely kinematic case. We find that the 3:2 spin-orbit resonance is not only combined with a first synchronism where the spin and orbit poles on average precess at the same rate but also with a second one where the rotational nutation desi- gnated by θ and the i orbit inclination on average decrease at the same rate (see Fig. 2 where is shown the libration of ξ = θ − i within the Ψ period). These two synchronisms reflect the proximity to the Cassini state and their signatures express its conserving in the N-body problem case. Besides, the conceptual question to know whether the ultimate minimal amplitude could be of zero value or not is discussed further. However, regardless of that, inside the present numerical process, the amplitudes in the Sn case are related to a spin-orbit coupling. As shown below with the equation (9), it is about an indirect effect due to the solar torque, bearing the planetary interactions, and acting on the planet's rotation. Reference frames of the S2K and Sn cases being linked by the conditions (4) (see Fig. 1) and values of the free constants required for SONYR(S2K, C2K) being determined from SONYR(Sn, Cn), SONYR(S2K, C2K) has been built at the heart of SONYR(Sn, Cn) in such a way the quasi-periodic solution of SONYR(Sn, Cn) surrounds the vicinity of the quasi-periodic solution of SONYR(S2K, C2K). Arbitrary amplitudes are not only avoided in the oscilla- tions of η, but a direct comparison of η coming from Sn with η coming from S2K is possible, which makes in evidence the effect of the planetary interactions. In the Sn mode indeed, the gravitational solar torque, bearing the planetary interactions and acting on the dynamical figure of Mercury, induces more size- able librations related to the disturbed Euler angles as well as the variables i and Ω. At this phase of analysis, the obliquity can be no longer said quasi- constant. The oscillation amplitude of period Ψ is indeed equal to 4 as (total peak-to-peak amplitude), let be one order of magnitude larger than in the S2K case (or 28 times larger when comparing the peak-to-peak amplitudes of the strict Ψ modulation without high frequency fluctuations) (see Fig. 6b and Fig. 7b). The comparison of the S2K and Sn modes, respectively derived from the 2-body and N-body problems, both computed and plotted at their respective libration centers Γ, shows how Mercury's obliquity and its dynami- cal behavior are modified when the planetary interactions are fully introduced (see Fig. 8). The mean obliquity is larger as well as the oscillation amplitude. < η > = 1.98 amin in the Sn case while < η > = 1.80 amin in the S2K case. This spin-orbit mechanism is rightly an indirect effect due to the solar torque, bearing the orbital planetary interactions and acting on Mercury's rotation, as shown by the classic general equation of the rotational motion of a solid celestial body (Bois 1995) : 14 ∂[(I)ωωω] ∂t + ωωω × [(I)ωωω] = (cid:18)3Gm r(cid:19)3 a3 (cid:19)(cid:18)a uuu × (I) uuu (9) where ωωω is the instantaneous rotational vector of Mercury and (I) its tensor of inertia. The right hand side of the equation represents the second degree torque exerted by a point mass m (the Sun) of position ruuu with respect to the body (Mercury) when it is reduced to three oblateness coefficients. a is the mean distance between the two bodies while r is the instantaneous one. ruuu is solution of the N-body problem in SONYR(Sn). Let us note that the direct planetary torques are about two orders of magnitude smaller than the indirect effects acting by the solar torque. Figure 8 is here Due to the planetary interactions, Mercury's orbit is no longer Keplerian and rate of precession of Mercury's orbital plane is no longer uniform. As a conse- quence, the obliquity's dynamical behavior is no longer quasi-constant. More generally, it is not possible for the obliquities to be constant since the orbit precession occurs at non-uniform rates (Bills 2005). In the case of Mercury, according to Bills & Comstock (2005), if the orbit pole precession rate were uniform, dissipation would have driven Mercury into a Cassini state, in which the spin pole and orbit pole remain coplanar with the Laplace pole, as the spin pole precesses about the moving orbit pole. However, according to these authors and in accordance with our results, due to the nonuniform orbit pre- cession rate with several distinct modes of oscillation, this simple coplanar configuration is no longer globally maintained. 7 Discussion We have significantly reduced amplitudes of the oscillations in Mercury's obli- quity, which were exposed to discussion since a previous paper (Rambaux & Bois 2004). For doing that, we have built an original method for finding the libration center of the spin-orbit system and as a consequence for avoiding ar- bitrary amplitudes in librations of the spin-orbit motion as well as in Mercury's obliquity. We have converged and the amplitudes are in very good agreement with most recent observations (Radar observations by Margot et al. 2006). This validates not only our theoretical SONYR model applied to Mercury but also our method of convergence. Our centering method has converged but it is not excluded that it may re- main a residual amplitude of arbitrary oscillations. It ought to be possible to push farther the convergence process by drawing nearer yet more precisely to the libration center (in the framework of the paper, i.e. in a conservative dynamical system). Since our amplitudes are compatible with current observa- tions, we have concluded that this question could wait for MESSENGER and 15 BepiColombo data in order to directly fit the SONYR model to observations. Nevertheless, apart from values of residual amplitudes, it remains a conceptual question. Does there exist a theoretical possibility to reach a pure point of equilibrium where the oscillation amplitude would be absolutely of zero value or, on the contrary, does there exist an irreducible minimum, even very small but of non-zero value? In the Sn case, existence of distant resonances of high orders that would excite the 3:2 resonance is not excluded a priori. None mathematical theorem does deny it but the demonstration of this would take us along to a purely mathematical ground. Consequently, we will say that the question remains open (both in conservative and dissipative cases). However, we plane some further works in this connection. The occurrence of a perfect or rigid equilibrium without oscillations seems to us unlikely for natural systems. 8 Conclusion Using our SONYR model of the solar System including the spin-orbit cou- plings of Earth-like planets, and setting the system at its libration center in order to eliminate arbitrary amplitudes (at the internal accuracy level of the model), we have obtained the complete obliquity for the full spin-orbit mo- tion of Mercury. Our construction of reference systems permits to accurately evaluate the obliquity's oscillations (i) not only with respect to the libration center of the full spin-orbit motion of Mercury (Sn), (ii) but also with respect to a Cassini state derived from the 2-body problem (S2K). The construction is such as the exact Cassini state derived from S2K coıncides with the mean Cassini state of Sn on a given interval of time. We have obtained (1) a mean value of Mercury's obliquity that is in a very good agreement with very recent observations (Radar observations by Margot et al. 2006), namely 1.98 amin with SONYR for 2.1 ± 0.1 amin by observations, (2) amplitudes of oscillations in Mercury's obliquity of the order of 2 as, which is compatible with the results of these observations where the deviation with respect to the 2-body Cassini state is constrained lesser than 6 as. These observational data validate not only our theoretical prediction with the SONYR model applied to Mercury but also our method of convergence for finding a libration center in a complex spin-orbit system. Being dependent on the question of existence of an irreducible minimum in the oscillation amplitude, we find that the equilibrium state of Mercury computed in the framework of a conservative N-body system squares to a quasi-periodic oscillation without secular drifts. In other words, in spite of the planetary interactions indirectly acting by the solar torque on Mercury's rotation, Mer- cury remains in a stable equilibrium state that proceeds from a 2-body Cassini state. Besides, by the way of our method, we have made in evidence the richness of a dynamical precession related to planetary interactions (N-body problem) rel- ative to a purely kinematical orbit precession added to a 2-body problem (Sun 16 and Mercury). Whatever the open question on the minimal amplitude in the obliquity's oscillations, the approach of the 2-body Cassini state is necessarily limited to reach a constant or quasi-constant behavior of Mercury's obliquity (where moreover < η > = 1.80 amin versus 1.98 amin, for a difference of 10.8 as). The planetary interactions generate a forced dynamical precession of Mercury's orbital plane with a nonuniform rate. This precession is kine- matically taken into account in the conventional construction of a Mercury's Cassini state as well as in our S2K case. But in the Sn case, due to the first synchronism where the spin and orbit poles on average precess at the same rate (as well as the second one where the spin axis orientation, namely the θ nutation, and the orbit inclination i on average change at the same rate), the difference angle between these two poles, i.e. η, (and θ − i as well) li- brates necessarily. As a consequence and whether it would turn out that there exists an irreducible minimum of the oscillation amplitude, so these resulting quasi-periodic oscillations in the obliquity ought to be understood as librations related to these two synchronisms where the (ψ − Ω) and (θ − i) respective differences librate (according to the Ψ proper frequency in latitude). In the same way, in multi-planet systems, when the longitudes of periapse on average precess at the same rate, the difference angle between the two apsidal lines librates (see e.g. Lee & Peale 2003 or Bois et al. 2003). Under this assumption and according to the obtained dynamical behavior, the obliquity of Mercury should be characterized by a quasi-periodic oscillation according to the period Ψ, with an amplitude Aη and a mean obliquity < η >, in such a way that Mercury's obliquity η could be written as follows (in first approximation) : η = < η > +Aη cos(cid:18) 2π Ψ t + α(cid:19) (10) where α is a phase factor. In the end, for lack of current and accurate observations or ephemerides of n) set up at the libration center Mercury's rotation, but with SONYR(Sn, C ′ of Mercury's spin-orbit motion, we may conclude that our method is relevant to link physical causes and their respective effects. As a consequence, with accurate observations of Mercury's rotational motion, that should be being achieved by MESSENGER and BepiColombo, SONYR(Sn, C ′ n), and not a 2- 2K), ought to be fitted to these body Cassini state model as SONYR(S2K, C ′ observations. Acknowledgments The authors thank the anonymous referee and Bruce Bills for their constructive requirements. NR acknowledges the ROB and FUNDP for their fellowship supports. References [1] Bills, B.G., 2005, Free and forced obliquities of the Galilean satellites of Jupiter, Icarus 175, 233 17 [2] Bills, B.G. & Comstock, R.L., 2005, Forced obliquity variations of Mercury, Journal of Geophysical research 110, 1 [3] Bizouard, C., Schastok, J., Soffel, M., Souchay, J., 1992, in : N. Capitaine (ed.), Journ´ees 1992 : Syst`emes de r´ef´erence spatio-temporels, Observatoire de Paris, 76 [4] Bois, E., 1995, Proposed Terminology for a General Classification of Rotational Swing Motions of the Celestial Solid Bodies, A&A 296, 850 [5] Bois, E., 2000, Knowledge of the lunar librations at the Lunar Laser Ranging experiment epoch, C. R. Acad. Sci. Paris, t. 1, S´erie IV, 809 [6] Bois, E., & Vokrouhlick´y, D., 1995, Relativistic spin effects in the Earth-Moon system, A&A 300, 559 [7] Bois, E., Kiseleva-Eggleton, L., Rambaux, N., & Pilat-Lohinger, E., 2003, Conditions of Dynamical Stability for the HD 160691 Planetary System, Astrophysical Journal 598, 1312 [8] Lee, M.H., & Peale, S., 2003, Secular Evolution of Hierarchical Planetary Systems., 2003, Astrophysical Journal 592, 1201 [9] Colombo, G., 1965, Rotational period of the planet Mercury, Nature 208, 575 [10] Colombo, G., 1966, Cassini's second and third laws, AJ 71, 891 [11] Correia, A., & Laskar, J., 2004, Mercury's capture into the 3/2 spin-orbit resonance as a result of its chaotic dynamics, Nature 429, 848 [12] Damour, T., Soffel, M., & Xu, Ch. 1991, General-relativistic celestial mechanics. I. Method and definition of reference systems, Phys. Rev. D 43, N. 10, 3273 [13] Damour, T., Soffel, M., & Xu, Ch. 1992, General-relativistic celestial mechanics. II. Translational equations of motion, Phys. Rev. D 45, N. 4, 1017 [14] Damour, T., Soffel, M., & Xu, Ch. 1993, General-relativistic celestial mechanics. III. Rotational equations of motion, Phys. Rev. D 47, N. 8, 3124 [15] D'Hoedt, S., & Lemaıtre, A., 2004, The spin-orbit resonant rotation of Mercury: a two degree of freedom Hamiltonian model, Celest. Mech. and Dyn. Astron. 89, 267 [16] Fukushima, T., 1991, Geodesic nutation, A&A 244, L11 [17] Goldreich, P., & Peale, S., 1966, Spin-orbit coupling in the solar system, AJ 71, N. 6, 425 [18] Kaula, W.M., 1966, Theory of Satellite Geodesy; Applications of Satellites to Geodesy, Blaisdell, Waltham, MA [19] Laskar, J., & Robutel, P., 1993, The chaotic obliquity of the planets, Nature 361, 608 18 [20] Margot, J.-L., Peale, S.J., Slade, M.A., Jurgens, R.F., & Holin, I., 2006, Obervational proof that Mercury occupies a Cassini state, AAS, DPS Meeting #38, #49.05 [21] Milani, A., Vokrouhlick´y, D., & Bonanno, C., 2001, Gravity field and rotation state of Mercury from the BepiColombo Radio Science Experiments, Planet. Space Sci. 49, 1579 [22] Moons, M., 1982, Analytical theory of the libration of the moon, Moon and the Planets 27, 257 [23] Peale, S.J., 1969, Generalized cassini's laws, AJ 74, 483 [24] Peale, S.J., 1972, Determination of parameters related to the interior of Mercury, Icarus 17, 168 [25] Peale, S.J., 2006, The proximity of Mercury's spin to Cassini state 1 from adiabatic invariance, Icarus 181, 338 [26] Peale, S.J., Phillips, R.J., Solomon, S.C., Smith, D.E., & Zuber, M.T., 2002, A procedure for determining the nature of Mercury's core, Meteo. and Planet. Sci. 37, 1269 [27] Pettengill, G.H., & Dyce R.B., 1965, A Radar determination of the rotation of the planet Mercury, Nature 206, 1240 [28] Rambaux, N., & Bois, E., 2004, Theory of the Mercury's spin-orbit motion and analysis of its main librations, A&A 413, 381 [29] Solomon, S.C., 20 colleagues, 2001, The MESSENGER mission to Mercury: scientific objectives and implementation, Planet. Space Sci. 49, 1445 [30] Wu, X., Bender, P.L., & Rosborough, G.W., 1995, Probing the interior structure of Mercury from an orbiter plus single lander, Planet. Space Sci. 45, 15 [31] Yseboodt, M., & Margot, J.L., 2006, Evolution of Mercury's obliquity, Icarus 181, 327 19 Proper Frequencies Φ (years) Ψ (years) Rambaux & Bois 2004 D'Hoedt & Lemaıtre 2004 15.847 15.857 1066 1066.68 This paper 15.824 ± 0.024 1066.8 ± 0.4 Table 1 Proper frequencies related to the spin-orbit motion of Mercury (in longitude: Φ, in latitude: Ψ). Comparaison of results. 20 Basic periods Pϕ (rotation period) Pλ (orbital period) = = 58.646 days 87.969 days Π (orbital precession) = 279 000 years Table 2 Basic periods of the spin-orbit motion of Mercury (Rambaux & Bois 2004) 21 S2K Sn δΩ0 (amin) 0.0 -6.564 δIK (amin) -1.806 -1.823 Table 3 Numerical values of δΩ0 and δIK in the S2K and Sn cases. 22 Fig. 1. Planes of reference of the S2K and Sn cases linked by the conditions (4) (see Section 3). 23 ) g e d ( i , θ ) n i m a ( ξ 7.04 7 6.96 6.92 6.88 1.84 1.82 1.8 1.78 1.76 1.74 0 200 400 600 800 1000 1200 1400 1600 T (years) 0 200 400 600 800 1000 1200 1400 1600 T (years) Fig. 2. Second synchronism of the spin-orbit motion of Mercury. θ and i are plotted in the top panel over 1600 years. The difference ξ = θ − i plotted on the bottom panel librates within the proper frequency in latitude, namely Ψ = 1066.8 years. Degrees (top panel) and arcminutes (bottom panel) are on the respective vertical axes. Years are on both horizontal axes. 24 " Y S -0.122 -0.1222 -0.1224 -0.1226 -0.1228 -0.123 -0.0006 -0.0004 -0.0002 0.0002 0.0004 0.0006 0 SX" Fig. 3. Projection of Mercury's spin pole S in mode SONYR(S2K , C2K ) on the OX"Y " plane (i.e. rotating with Π). 25 (1) (2) " Y S -0.122 -0.1222 -0.1224 -0.1226 -0.1228 -0.123 -0.0006 -0.0004 -0.0002 0.0002 0.0004 0.0006 0 SX" Fig. 4. Projection of Mercury's spin pole S on the OX"Y " plane. The large circle (1) obtained in mode SONYR(S2K , C2K ) is the one of Fig. 3. In mode SONYR(S2K , C ′ 2K ), it is a broadly reduced circle (2). 26 -0.1218 -0.122 -0.1222 -0.1224 -0.1226 -0.1228 -0.123 " Y S (1) (2) -0.1232 -0.001 -0.0008 -0.0006 -0.0004 -0.0002 0 0.0002 0.0004 SX" Fig. 5. Projection of Mercury's spin pole S on the OX"Y " plane. The large circle (1) is obtained in mode SONYR(Sn, Cn) whereas the broadly reduced circle (2) in mode SONYR(Sn, C ′ n). 27 ) n i m a ( η ) n i m a ( η 4 3.5 3 2.5 2 1.5 1 0.5 0 (a) 0 200 400 600 800 1000 1200 1400 1600 1.808 1.807 1.806 1.805 1.804 1.803 1.802 1.801 T (years) (b) 0 200 400 600 800 1000 1200 1400 1600 T (years) Fig. 6. (a) Resulting obliquities by SONYR(S2K, C2K ) and SONYR(S2K, C ′ 2K ). (b) Zoom of η in the SONYR(S2K, C ′ 2K) case, i.e. without arbitrary amplitudes. η is characterized by a main oscillation according to the second proper frequency Ψ within a peak-to-peak amplitude lesser than 0.4 as. In both panels, arcminutes are on the vertical axes and years on the horizontal ones. 28 ) n i m a ( η ) n i m a ( η 4 3.5 3 2.5 2 1.5 1 0.5 0 2.01 2 1.99 1.98 1.97 1.96 1.95 1.94 1.93 (a) 0 200 400 600 800 1000 1200 1400 1600 T (years) (b) 0 200 400 600 800 1000 1200 1400 1600 T (years) Fig. 7. (a) Resulting obliquities by SONYR(Sn, Cn) and SONYR(Sn, C ′ n). (b) Zoom of η in the SONYR(Sn, C ′ n) case, i.e. without arbitrary amplitudes. η is charac- terized by a clear oscillation according to the second proper frequency Ψ while the peak-to-peak amplitude is near 4 as. Arcminutes are on the vertical axes and years on the horizontal ones. 29 ) n i m a ( η 2.05 2 1.95 1.9 1.85 1.8 0 200 400 600 800 1000 1200 1400 1600 T (years) Fig. 8. Comparison of the obliquity's dynamical behavior in SONYR(Sn, C ′ n) and SONYR(S2K, C ′ 2K ) modes, i.e. with and without planetary interactions, both com- puted and plotted at their respective libration center Γ. < η > = 1.98 amin in the Sn case while < η > = 1.80 amin in the S2K case. 30
0811.2902
1
0811
2008-11-18T14:03:32
Interstellar Extinction and Long-Period Variables in the Galactic Center
[ "astro-ph" ]
We use the Spitzer IRAC catalogue of the Galactic Center (GC) point sources (Ramirez et al. 2008) and combine it with new isochrones (Marigo et al. 2008) to derive extinctions based on photometry of red giants and asymptotic giant branch (AGB) stars. This new extinction map extends to much higher values of Av than previoulsy available. Our new extinction map of the GC region covers 2.0 x 1.4 degree (280 x 200 pc at a distance of 8 kpc). We apply it to deredden the LPVs found by Glass et al. (2001) near the GC. We make period-magnitude diagrams and compare them to those from other regions of different metallicity. The Glass-LPVs follow well-defined period-luminosity relations (PL) in the IRAC filter bands at 3.6, 4.5, 5.8, and 8.0 micron. The period-luminosity relations are similar to those in the Large Magellanic Cloud, suggesting that the PL relation in the IRAC bands is universal. We use ISOGAL data to derive mass-loss rates and find for the Glass-LPV sample some correlation between mass-loss and pulsation period, as expected theoretically.The GC has an excess of high luminosity and long period LPVs compared to the Bulge, which supports previous suggestions that it contains a younger stellar population.
astro-ph
astro-ph
Astronomy & Astrophysics manuscript no. paperastroph August 10, 2018 c(cid:13) ESO 2018 Interstellar Extinction and Long-Period Variables in the Galactic Center M. Schultheis1,2, K. Sellgren3, S. Ram´ırez4, S. Stolovy5, S. Ganesh6, I.S. Glass7, and L. Girardi8 8 0 0 2 v o N 8 1 ] h p - o r t s a [ 1 v 2 0 9 2 . 1 1 8 0 : v i X r a 1 Observatoire de Besanc¸on, 41bis, avenue de l'Observatoire, F-25000 Besanc¸on, France 2 Institut d'Astrophysique de Paris, CNRS, 98bis Bd Arago, F-75014 Paris, France e-mail: [email protected] 3 Astronomy Department, Ohio State University, Columbus, OH 43210, USA e-mail: [email protected] 4 IPAC, Caltech, Pasadena, CA 91125, USA e-mail: [email protected] 5 Spitzer Science Center, Caltech, Pasadena, CA 91125, USA e-mail: [email protected] 6 Physical Research Laboratory, Astronomy & Astrophysics Division, Ahmedabad, India e-mail: [email protected] 7 South African Astronomical Observatory, PO Box 9, Observatory 7935, South Africa 8 Osservatorio di Padua, Italy received:??; accepted ?? ABSTRACT Aims. To derive a new map of the interstellar extinction near the Galactic Center (GC) extending to much higher values of AV than previously available. To use the results obtained to better characterise the long-period variable star population of the region. Methods. We take the Spitzer IRAC catalogue of GC point sources (Ram´ırez et al. 2008) and combine it with new isochrones (Marigo et al. 2008) to derive extinctions based on photometry of red giants and asymptotic giant branch (AGB) stars. We apply it to deredden the LPVs found by Glass et al. (2001) (Glass-LPVs) near the GC. We make period-magnitude diagrams and compare them to those from other regions of different metallicity. Results. Our new extinction map of the GC region covers 2.0 o × 1.4 o (280 × 200 pc at a distance of 8 kpc). The Glass-LPVs follow well-defined period-luminosity relations (PL) in the IRAC filter bands at 3.6, 4.5, 5.8, and 8.0 µm. The period-luminosity relations are similar to those in the Large Magellanic Cloud, suggesting that the PL relation in the IRAC bands is universal. We use ISOGAL data to derive mass-loss rates and find for the Glass-LPV sample some correlation between mass-loss and pulsation period, as expected theoretically. Theoretical isochrones for a grid of different metallicities and ages are able to reproduce this relation. The GC has an excess of high luminosity and long period LPVs compared to the Bulge, which supports previous suggestions that it contains a younger stellar population. Key words. ISM: dust, extinction -- Galaxy: stellar content -- Infrared: stars 1. Introduction The variable stars of the Galactic Center (GC) region are of great interest, both as population and distance indicators. Long- period, large-amplitude variables situated on the asymptotic gi- ant branch (AGB), comprising Miras and OH/IR stars, are the easiest objects to detect thanks to their high luminosities. Thus, they are one of the few stellar populations that can be observed in their entirety towards the GC. As is well-known, in the inner- most parts of the Galaxy, surveys are hampered by high inter- stellar extinction (e.g. Frogel et al. 1999; Schultheis et al. 1999); observations must therefore be carried out in the infrared where Aλ can be as low as 0.04 AV for 4µm < λ < 8µm (Indebetouw et al 2005). Sensitive surveys at the latter wavelengths have only recently become possible thanks to the Spitzer satellite. For these reasons, Glass et al. (2001) conducted a K-band (2.2 µm) survey for variable stars covering 24 × 24 arcmin2 (56 × 56 pc at a distance of 8 kpc) and centered on the GC in a study spanning 4 years. The majority of the variable sources they Send offprint requests to: M. Schultheis found were, as expected, Miras and OH/IR stars with periods ranging from 150 d to about 800 d. Uncertainty as to the fore- ground extinction unfortunately precluded any detailed compar- ison of their luminosities with similar populations in other well- studied areas, such as the the solar neighbourhood, the Baade's Windows and the Magellanic Clouds, where period-luminosity relations have been determined. The inner Galaxy has also been searched intensively in the radio region for OH sources (see e.g. Lindqvist et al. 1992, Sjouwerman et al. 1996, Wood et al. 1998, Blommaert et al. 1998, Vanhollebeke et al. 2006). From these surveys, it is clear that a number of the large amplitude variables remained unde- tected in the Near-infrared, owing to the extremely high extinc- tion in some regions. Nevertheless, from OH and SiO observa- tions, (see e.g. Messineo 2004 and Deguchi et al. 2008) radial velocities are available for many of these sources, making them extremely valuable for studies of stellar kinematics near the GC. Because the large-amplitude variables have very high mass- loss rates they also play an important role in the chemical evo- 2 Schultheis et al.: Interstellar Extinction and Long-Period Variables in the Galactic Center lution of the Galaxy. It is therefore desirable that they should be characterised as fully as possible. Previous studies in the near-IR using 2MASS or DENIS data have yielded extinction maps with a typical resolution of several arcmin (see e.g. Schultheis et al. 1999, Marshall et al. 2006). These maps, however, were limited by the sensitivities of the near-IR surveys and are only realistic in regions where AV is less than 25 magnitudes. Unfortunately, this limit is generally exceeded towards the GC. In this paper we show that an improved map of the inter- stellar extinction around the GC can be made by fitting mid- IR data on AGB/RGB stars obtained by the IRAC camera on Spitzer to isochrones (see also Ganesh et al., in preparation). Extinctions have been determined for much more heavily ob- scured areas than was possible previously, especially towards the GC itself. We apply the new extinction values to a discus- sion of the long-period variables in the region, concentrating on the period-magnitude relations that they obey in the IRAC bands and their mass-loss rates. 2. The data set 2.1. The Spitzer IRAC Point Source Catalog of the Galactic Center (GALCEN) The central 2.0 o × 1.4 o of the GC have been mapped with Spitzer/IRAC between 3.6 and 8.0 µm (Stolovy et al. 2006 and Stolovy et al. 2008, in preparation). Ram´ırez et al. (2008) per- formed point-source extraction on the IRAC data and published a confusion-limited catalogue of point sources that also included photometry from 2MASS. The IRAC magnitudes are referred to as [3.6], [4.5], [5.8] and [8.0], corresponding to their central wavelengths. The average confusion limits are 12.4, 12.1, 11.7 and 11.2 magnitudes, respectively, but can vary by 2 or 3 mag- nitudes within the survey. The whole catalogue (referred to here as GALCEN, see Ram´ırez et al. 2008 for more details) consists in total of about one million sources. Most of these show the characteristics of red giants or AGB stars, but there are several hundreds of extremely red sources which may be massive Young Stellar Objects (YSOs). A major problem that must be faced in the study of the LPV component is the bright-star limit of the IRAC camera. According to the Spitzer Observer's Manual (Version 7.1), saturation starts at 7.9, 7.4, 4.8 and 4.8 magni- tudes respectively. Saturated sources are however retained in the GALCEN catalogue but if a flux is greater than the limit it is flagged with the number '3' (see Ram´ırez et al. 2008 for a more detailed discussion). We have excluded photometry with flux flag values of 3 from our analysis. IRAC by the processed reduction and source 2.2. GALCEN vs GLIMPSE-II GALCEN Recently, observations have been independently the GLIMPSE-II http://www.astro.wisc.edu/sirtf/docs.html). team (see Different data extraction proce- dures from those of the GALCEN survey were used (see http://www.astro.wisc.edu/glimpse and Benjamin et al. 2003). These data were kindly made available to us in advance of pub- lication by the GLIMPSE-II team. As discussed in the online document, they found systematic offsets between the GALCEN and GLIMPSE-II photometry that was larger than the combined uncertainty in both observations, at the bright and faint ends of the observed range. To look for any possible systematic effects on our work we have also calculated the extinction maps that we discuss in Sect. 3 using the GLIMPSE-II results. We find that the differences in the values of the extinctions derived from GALCEN and from GLIMPSE-II are smaller than the uncertainties in the determinations. The locations of the LPVs in the IRAC colour-colour dia- grams, the IRAC colour-magnitude diagrams, and the period- luminosity relations (see Sections 4 -- 5) also differ slightly between the GALCEN and GLIMPSE-II photometry. The GALCEN and GLIMPSE-II photometry differ by ∼0.05 mag in the mean, with a standard deviation of 0.12 mag, over the brightness range of the LPVs. We find no statistically signifi- cant dependence of the difference between the GALCEN and GLIMPSE-II photometry on magnitude, or on the periods of the LPVs (which can be regarded as a proxy for their de-reddened magnitudes). The scatter in this difference, however, is large and we conservatively place a limit of 0.2 mag on any possible varia- tion in its value over the brightness range occupied by the LPVs. Our conclusions in this paper are, however, unaffected by photo- metric uncertainties at this level. 2.3. Long-Period Variables The Glass et al. (2001) survey covered an area of 24 × 24 arcmin2 around the GC, as mentioned, and found about 400 pe- riodic variables with Mira-like amplitudes and an average pe- riod of 427 d. (Among these are 64 OH/IR stars included in the Wood et al. 1998 sample). We refer to the survey list hereafter as the Glass-LPV catalogue. Compared to the well-explored SgrI Baade's window of low extinction which lies at l = −1.4◦ and b = −2.6◦, the GC field contains more than ten times the number of variables per arcmin2. The average period in the SgrI field is much lower, at 333 d. Due to the limited photometric precision of the Glass et al. (2001) survey, small amplitude variables such as semi-regulars were not detected. 2.4. ISOGAL observations The details of the ISOGAL observational procedure with ISOCAM (Cesarsky et al. 1996) and the general processing of the data are described in Schuller et al. (2003). In this paper we will discuss only the ISOGAL FC−00027 − 00006 field (here- after called FC−027), centered at (l, b) = (−0.27◦,−0.06◦). It was observed with the narrow filters LW5 (6.5 -- 7.1 µm) and LW9 (13.9 -- 15.9 µm), using 3(cid:48)(cid:48) × 3(cid:48)(cid:48) pixels, in order to avoid sat- uration. More specific details about this field can be found in Schuller et al. (2006). The parts of the region closest to the GC itself could not be observed by ISOGAL due to saturation prob- lems even when the narrow-band filters were used. Nevertheless, 25% of the Glass-LPV area is overlapped by the ISOGAL survey at 7 and 15 µm, mainly towards the south-western corner. 2.5. Cross-identifications Cross-identification of stellar catalogues in these extremely crowded regions, which are often confusion limited, is a very difficult issue. In order to minimize spurious associations we used the following method when comparing the different cata- logues mentioned above. We first calculated the offsets of the 400 brightest sources between a given pair of catalogues and determined the second-order distortion matrix. Following appli- cation of the offset and distortion corrections, we made the sub- sequent cross-identifications using a search radius of 2 arcsec: Schultheis et al.: Interstellar Extinction and Long-Period Variables in the Galactic Center 3 Fig. 1. Upper panel: Interstellar extinction map of the central 280 × 200 pc of the Galactic Center with spatial resolution 2(cid:48), derived using 2MASS J − Ks data. Lower panel: As above, except using the [3.6]−[5.8] colour-magnitude data for point sources from the Ram´ırez et al. (2008) GC catalogue. -- First, we cross-identified the LPVs of Glass et al. (2001) with the GALCEN catalogue. By limiting the GALCEN cat- alogue to the area covered by the Glass-LPV survey, it was reduced to a total of 46314 sources. Of the 421 LPVs, 410 were found in the GALCEN list (97%). Many of them are, however, saturated, especially at [3.6] and [4.5]. (From the Glass-LPV sample there were 162, 178, 36, and 58 saturated sources at [3.6], [4.5], [5.8], and [8.0], respectively). -- Second, we cross-identified the Glass-LPV catalogue with ISOGAL. We found 183 matches, despite the small overlap. 010203040501.0000.5000.000359.500359.0000.8000.6000.4000.2000.000-0.200-0.400-0.600-0.800-1.000Galactic longitudeGalactic latitudeAv(2MASS)010203040501.0000.5000.000359.500359.0001.0000.8000.6000.4000.2000.000-0.200-0.400-0.600-0.800Galactic longitudeGalactic latitudeAv(IRAC) 4 Schultheis et al.: Interstellar Extinction and Long-Period Variables in the Galactic Center 3. Interstellar extinction 3.1. Derivation of the extinction values As stated, interstellar extinction remains a serious obstacle to the interpretation of stellar populations in the GC region. The present work follows the method of Ganesh et al. (in preparation) who recently mapped the extinction towards the inner Bulge with a spatial resolution of 2(cid:48). GLIMPSE-II and 2MASS data are compared to the latest isochrones for evolved stars (Marigo et al. 2008) in order to derive AV. Several studies of this type, using only the 2MASS data, have been undertaken before (e.g. Schultheis et al. 1999, Marshall et al. 2006). In the method of Ganesh et al, colour-magnitude diagrams are constructed within sampling boxes of 2(cid:48) square and the amount by which each individual data point has to be de- reddened for it to fall on the isochrone is determined. The ex- tinction values for AV > 20 are derived using the [3.6]−[5.8] colour, while 2MASS J and Ks data are used for AV < 20. With the large number of filter combinations available (three from 2MASS and four from GLIMPSE-II) several possible com- binations had to be considered. The [5.8] vs. [3.6]−[5.8] colour- magnitude diagram was selected since it was found to yield the smallest dispersion, i.e. the [3.6]−[5.8] colour excess was found to be the most suitable one for determining the interstellar ex- tinction (see e.g. Indebetouw et al. 2005). This pair of filters has, in fact, the advantage of simultaneous observations (the other two IRAC colours were observed at different times) and also has better completeness (more stars detected through high extinc- tion). Further, at 3.6 µm and 5.8 µm, the observed flux is dom- inated by stellar photospheric emission while at 8.0 µm inter- stellar PAH emission becomes prominent. The models indicate that metallicity effects are negligible in the [3.6]−[5.8] colour, in contrast to J − Ks, where the RGB/AGB branch of a metal-rich stellar population shifts to redder J − Ks colours (Schultheis et al. 2004). For more detailed information see Ganesh et al. (in preparation). In the following, we took the GALCEN catalogue of Ram´ırez et al. (2008) and limited ourselves to stars with [3.6] < 12.5 and [5.8] < 12. We have also excluded sources with [3.6] -- [4.5] > 0.8 and [5.8] -- [8.0] > 0.5 from our extinction cal- culation, because sources with these colors are typical of YSOs (see Allen et al. 2004) and the inclusion of their extreme col- ors would bias the results. We again chose a sampling box of 2(cid:48) square in order to get a sufficient numbers of stars for the isochrone fitting. For each sampling box the average extinction value was taken. The values of Aλ/AV that we use to make the extinction maps were derived by using the infrared colour-colour diagram J − Ks vs. Ks−[IRAC] as proposed by Jiang et al. (2003), Indebetouw et al. (2005) and Ganesh et al. (2008). We used the J − Ks vs. Ks−[3.6] and the J−Ks vs. Ks−[5.8] diagrams to determine A[3.6] and A[5.8]. We assumed that most of the sources are luminous RGB stars or AGB stars with moderate mass-loss and similar in- trinsic colours. We further adopted the values AJ = 0.256 × AV = 0.089× AV as used by Schultheis et al. (1999). Fitting and AKS straight lines in these diagrams gives us the following values: A[3.6]/AV = 0.0498 ± 0.0015 and A[5.8]/AV = 0.0308 ± 0.0015. These values were determined using the whole GALCEN cat- alog. Ganesh et al. (2008) will discuss more in detail how the extinction coefficients vary with different line of sights using the whole GLIMPSE-II dat set. Our A[3.6]/AV value agrees within the errors with those from Indebetouw et al. (2005), Lutz (1999), Flaherty et al. (2007) and Rom´an-Z´uniga et al. 2007, while our derived A[5.8]/AV is a litte bit lower, but still lies inside the errors of the quoted authors. Lutz et al. (1996) and Lutz (1999) measured a flat mid-IR extinction curve towards the GC. However Flaherty et al. (2007) and Rom´an-Z´uniga et al. 2007 have shown that extinction coeffi- cents may vary somewhat depending in which line of sight one is looking (e.g. especially towards star forming regions). Recently, Chapman et al. (2008) have also studied the mid-infrared extinc- tion law in three molecular clouds and found that as the over- all extinction increases, the curve flattens out. The coefficients evidently vary according to whether the dust is located in the general ISM or in dense molecular clouds. Figure 1 shows the AV map derived from the 2MASS filters (upper panel) and that from the IRAC [3.6] and [5.8] filters of GALCEN (lower panel). There is a great similarity in the overall distribution of the two extinction maps. Where 2MASS gives lower limits for the extinction or is unable to probe the extinction due to lack of sources, the GALCEN extinction map exhibits high values for the extinction (AV > 40). Within Figure 1 there are regions of low extinction according to the 2MASS data but which are seen to have very high extinction when looked at by IRAC (e.g. the region around l = 0.2 o and b = −0.5 o). These areas are mostly associated with infrared dark clouds. 2MASS gives artificially low values of extinction in such cases because it detects only those stars located in front of the dark cloud whereas IRAC penetrates to much greater depth. Fig. 2. Extinction values derived from 2MASS compared to those derived from IRAC colours. The straight line gives the identity relation. The dotted line indicates the maximum value of AV to which 2MASS is sensitive. Since the [3.6] -- [5.8] colour is less than J -- KS , for a given amount of reddening the precision it yields in AV is not as high. The loss of accuracy at low AV can, however, be avoided by us- ing the near-infrared derived values for mapping when they are available. At larger AV, the results from the [3.6] and [5.8] map are, of course, to be preferred. Schultheis et al.: Interstellar Extinction and Long-Period Variables in the Galactic Center 5 Fig. 3. [5.8] vs. [3.6]−[5.8] colour-magnitude diagrams towards regions of low (left panel), intermediate (middle panel) and high (right panel) extinction (see text) in the GALCEN survey. The corresponding isochrones are displayed. The horizontal dotted line indicates the approximate confusion limit. The corresponding fields (from left to right) are α = 265.93596 o, δ = -28.44771 o; α = 266.16275 o, δ = -28.44771 o; and α = 267.027222 o, δ = -28.967888 o. The size of each field is 2(cid:48). The region with AV = 70m is located at α =267.027222 o and δ = -28.967888 o and is associated with the infrared dark cloud IRDN3 of Dutra & Bica (2001) who list it as an opaque region at Ks with an angular diameter of 3(cid:48). Different ages or metallicities for the Bulge population do not significantly affect the mid-infrared colours of RGB stars, as will be seen. The sequences in the sampling fields are well defined in the case of low extinction while there is some scat- ter when it is high. This scatter arises from variable extinction on small scales within the Bulge and along the line of sight. Towards regions with very high extinction (AV > 60), the num- ber densities in the CMD can drop significantly due to the loss of the fainter stars. Such points have relatively large errors. Figure 2 compares the extinction values derived from J and Ks 2MASS data and those derived from the [3.6] − [5.8] IRAC colours. There is a tendency towards a linear relation between AV(2MASS) and AV(IRAC) up to AV = 20 with a dispersion of AV (cid:39) 3. For higher AV values, 2MASS can no longer detect the most reddened stars at the GC and is biased towards less red- dened foreground stars. For those regions where 2MASS gives low AV values and IRAC high ones, different optical depths are obtained. The values of AV derived from J − Ks are representa- tive of the foreground population rather than the regions hidden by high extinction, while the IRAC bands relate to stars that are deeper within the clouds. Figure 3 shows the CMDs towards three regions of different extinction in the GALCEN survey. While for the low extinction field (left panel) the red giant branch is well defined and the dis- persion is low, the CMD is more scattered in the high extinction regions. Despite the high extinction, however, we get a sufficient number of sources to determine AV. 3.2. The combined extinction map Figure 4 shows the combined extinction (2MASS and IRAC) map as well as an intrinsic uncertainty map derived from the isochrone fitting. As explained before, for AV(IRAC) < 20 the 2MASS J − Ks colour was used, while for AV(IRAC) > 20 the IRAC [3.6]−[5.8] colour was used. The highest AV values we can detect correspond to about AV (cid:39) 90. These are mostly asso- ciated with clouds that are dark even in the infrared. We will provide in electronic form (at CDS) the three extinc- tion maps as well as the error map in the form of individual FITS files. 3.3. The uncertainty map The intrinsic uncertainty map corresponding to the combined interstellar extinction map is shown in Fig. 4 (lower panel). The uncertainty map is calculated as the standard deviation of the AV distribution. For AV < 20 the intrinsic uncertainty is about 2-3 mag while for AV > 20 it can go up to 6-7 mag. Figure 5 shows the comparison between our AV determinations based on GALCEN and on the GLIMPSE-II catalogue (see also Sect. 2.2). Despite the differences found in the photometry be- tween these two pipelines, no systematic offset is found between AV derived from GALCEN and AV derived from GLIMPSE- II. The typical r.m.s standard deviation between GALCEN and GLIMPSE-II is 6 mag in AV which is the overall uncertainty in the derived extinction map. For very high AV (AV > 40), the differences between GALCEN and GLIMPSE become larger. Howver, for those regions the CMD is less well defined and the intrinsic uncertainty increases (see Fig. 4). It should be noted that Nishiyama et al. (2006) found that the behaviour of near-IR extinction varies slightly though signif- icantly from one direction to another near the GC so that there is no universally valid extinction law for the region. These results were confirmed by Bandyopadhyay et al. (2008). Variations with direction of Aλ/AV in the IRAC bands do not appear to be sig- nificant at the accuracy that is currently available, at least in the 6 Schultheis et al.: Interstellar Extinction and Long-Period Variables in the Galactic Center Fig. 4. Upper panel: Combined interstellar extinction map of the central 280 × 200 pc of the Galactic Center. For AV(IRAC) < 20 the 2MASS J − Ks colour was used, while for AV(IRAC) > 20 the IRAC [3.6]−[5.8] colour was used. Lower panel: Intrinsic error map of the combined interstellar extinction map. The error is the sigma of the AV distribution. fields studied by Indebetouw et al (2005) and Ganesh et al. (in prepration). Gosling et al. (2006) found complex small-scale structure in the infrared extinction towards the GC, with a typical size of 5 -- 15(cid:48)(cid:48), corresponding to 0.2 -- 0.6 pc. They demonstrated that sig- nificant granularity appears to be present in high and variable extinction fields, indicating that extinction at J and H may have significantly higher values on smaller scales than those previ- ously derived for the GC as a whole. 4. The Long Period Variables near the Galactic Center In the remainder of this work we discuss the long-period vari- ables of the Galactic Center region. Figure 7 shows the 2MASS image of the GC region at KS . The known LPVs of Glass et al. (2001) are superimposed and are bright in KS . We note that patchy and filamentary dark clouds (appearing white in Fig. 7) often vary on size scales smaller than the 2(cid:48)sampling boxes used 010203040501.0000.5000.000359.500359.0001.0000.8000.6000.4000.2000.000-0.200-0.400-0.600-0.800Galactic longitudeGalactic latitudeAv(Combined)01234561.0000.5000.000359.500359.0001.0000.8000.6000.4000.2000.000-0.200-0.400-0.600-0.800Galactic longitudeGalactic latitudeError(IRAC) Schultheis et al.: Interstellar Extinction and Long-Period Variables in the Galactic Center 7 Fig. 5. Extinction values derived from GALCEN compared to those derived from GLIMPSE-II. The straight line gives the identity relation. here to calculate the extinction. This can contribute to the pho- tometric errors in the extinction correction towards individual sources. 4.1. Stellar isochrones The newest set of isochrones for the RGB/AGB phase include an improved treatment of the thermally pulsing asymptotic giant branch (TP-AGB) phase (Marigo et al. 2008). These isochrones predict realistic tracks for the TP-AGB, taking into account mass-loss and pulsation. The whole of TP-AGB evolution is now treated in a realistic way, especially the crucial effects of the third dredge-up, hot bottom burning and variable molecular opacities. The dust models incorporated in these isochrones are an ex- tension of the work by Groenewegen (2006). They have been calculated with a 1-dimensional dust radiative transfer code that solves the radiative transfer equation and the thermal balance equation in a self-consistent way. The main physical inputs to the model are the luminosity, distance, photospheric spectrum, mass-loss rate, dust-to-gas ra- tio, expansion velocity, dust condensation temperature and com- position of the dust. For oxygen-rich stars two main species of dust are considered: aluminium oxide (Al2O3) and silicate dust. Blommaert et al. (2006) studied the CVF ISOCAM spectra of a sample of AGB stars with low mass-loss in the Galactic bulge and found that the dust content is dominated by Al2O3 grains. Isochrones for different metallicities, Z=0.008, Z=0.019 and Z=0.038, and ages between 0.5 and 10 Gyr, have been calculated for the Spitzer IRAC filters. In the present work, we consider the isochrones placed at the distance to the GC (8.0 kpc, Reid 1993). Fig. 6. ([3.6] − [8.0])0 vs ([3.6] − [5.8])0 diagram. Open circles indicate the GC LPVs of Glass et al. (2001). Filled red circles denote oxygen-rich LPVs in the LMC (see text). The straight lines are models computed by Groenewegen (2006) using realis- tic stellar atmosphere models. The dashed green line is a model with 100% aluminium oxide dust, the dotted blue line is 60% aluminium oxide and 40% silicate, and the solid magenta line is 100% silicate dust. The extinction vector of AV = 20 is indi- cated. 4.2. LPVs in the Spitzer colour-magnitude and colour-colour diagrams Figure 8 (left panel) shows the [3.6]−[8.0] vs [8.0] diagram. Indicated are also the approximate saturation limits at [3.6] and [8.0]. In total, only 248 known LPVs have reliable (unsaturated) IRAC photometry. On the right panel of Fig. 8, the dereddened [3.6]−[8.0] vs [8.0] diagram, derived using the extinctions given in Fig. 4, shows clearly that the locus of the known AGB stars is above the RGB tip ([8]0 (cid:39) 7.3). As can be seen, the isochrones of Marigo et al. (2008) are a good fit to the dereddened CMD. This agreement confirms the correctness of the derived extinctions. The locations of the isochrones change only marginally with age and/or metallicity (e.g. from 1 to 10 Gyr and 0.008 < Z < 0.03 in these bands. The AGB stars show a large scatter in this dia- gram due to strong variability (we only have single-epoch mea- surements) and also to the infrared excesses arising from heavy mass-loss (see Sect. 6). A similar diagram can be drawn for LPVs at [4.5] where we find only 84 AGB variables to be un- saturated. 4.2.1. Comparison with Spitzer observations of LMC LPVs Cioni et al. (2001) examined a field of about 0.5 square de- grees around the optical center of the Large Magellanic Cloud. 8 Schultheis et al.: Interstellar Extinction and Long-Period Variables in the Galactic Center Fig. 7. 2MASS image (reversed grayscale) of the Galactic Center region at KS . The known LPVs from Glass et al. (2001) are superimposed with red circles. Fig. 8. Left panel: [3.6]−[8.0] vs [8.0] colour-magnitude diagram of the GALCEN field, with the known LPVs (red crosses) and OH/IR stars (blue open circles) superimposed. The extinction vector for AV =20 is indicated. The dotted horizontal line indicates the saturation limit at [8.0] while the dotted diagonal line shows the saturation limit at [3.6]. Right panel: Same diagram but dereddened (see text). Superimposed is the isochrone for 10 Gyr and Z=0.02 put at a distance of 8 kpc. Light curves in the EROS V and R bands were obtained for 334 variables, most of which are of long period (including Miras and semiregulars). In addition they obtained spectroscopy of about 100 AGB variables whose visible spectra were classi- fied as carbon- or oxygen-rich. We used their sample of known AGB stars and cross-identified their catalogue with the SAGE (Surveying the Agents of a Galaxy's Evolution) MIPS and IRAC imaging survey (Meixner et al. 2006). In the latter work, a field of 7 sq degrees of the LMC was surveyed between 3.6 and 160 µm. In total, 331 objects were found in the SAGE catalogue. For further discussion, we will refer to this as the LMC-AGB sam- ple. Schultheis et al.: Interstellar Extinction and Long-Period Variables in the Galactic Center 9 Figure 6 shows the dereddened [3.6]−[8.0] vs [3.6]−[5.8] colour-colour diagram for both the Galactic Center and the LMC AGB LPVs. We have superimposed for comparison theoretical dust models by Groenewegen (2006). These utilise several types of dust for the O-rich AGB stars: 100% aluminium oxide, a com- bination of 60% aluminium oxide and 40% silicate, and pure silicate dust (see Groenewegen 2006 for a discussion). All three models fail to match the observed colours and a significant offset is clearly visible. Our comparison is restricted to oxygen-rich AGB stars (283 stars). It should be noted that we split the LMC-AGB sample into oxygen-rich and carbon-rich AGB stars using only the J− K colour. Though Cioni et al. (2001) showed that carbon stars can in general be eliminated by avoiding colours in the range 1.4 < J − Ks < 2.0, they warn (see their Fig. 3) that stars selected in this way will contain a certain fraction of carbon stars. Figure 6 shows that the GALCEN variables and LMC oxygen-rich variables only partly overlap in the colour-colour diagram. The GALCEN LPVs show much redder colours in [3.6]−[8.0] and [3.6]−[5.8] than the LMC sample. These stars also have a higher proportion of long periods (Groenewegen & Blommaert 2005) relative to the LMC and the remainder of the Galactic Bulge. The redder colours in the IRAC bands may in- dicate that mass-loss rates of AGB stars in the GC are higher in general than in the LMC. However, the differing circumstellar dust-to-gas ratios between the LMC and GC makes direct com- parison difficult. The differences can also be a manifestation of the strong relation between period and mass-loss (see Sect. 6). 5. The Period-magnitude relations AGB variable stars in the Galactic Bulge follow a well-defined period-magnitude relation in the near infrared. Glass et al. (2001) showed that the mean K magnitude for the Glass-LPVs, measured over a complete cycle, showed some tendency to fol- low a log P vs. K relation in spite of a very high dispersion due to the large and unknown variations in AV. Some of the scatter may also have arisen because of crowding effects or insufficient ob- servations having led to incorrect periods, especially at the long (log P > 2.8) and short ends (log P < 2.3) of the range. The Glass et al. (2001) sample contains a higher fraction of long pe- riods than e.g. the SgrI field. Thus, circumstellar reddening due to heavy mass-loss affects the log P vs. K relation by depressing the emergent fluxes at longer wavelengths. Figure 9 shows the IRAC magnitude vs. log P relations for the Glass-LPVs, after correction of the photometry for interstellar extinction. We find following relations for the Glass-LPV sample: [3.6] = log P × -5.21 +18.94 [4.5] = log P × -5.71 +19.98 [5.8] = log P × -6.10 +20.70 [8.0] = log P × -6.73 +21.96 The typical r.m.s deviation for the four [IRAC] bands are 0.71 , 0.62, 0.60 and 0.61 mag in the period range between 2.2 < log P < 2.8. This is smaller than at K0 where the r.m.s is 0.95 mag (corrected according to the extinctions determined in this work) as the IRAC bands are less affected by any residual errors in the extinction. Even within the IRAC bands, the disper- sions at the longer wavelengths are less than that at 3.6 µm. The Glass-LPVs with periods greater than 900 days do not follow the log P vs. [IRAC] relations. Their periods could be incorrect as these only have a few observations (see Glass et al. 2001) and in addition these stars suffer heavy mass-loss which depress their [IRAC] fluxes. Superimposed on Fig. 9 is the LMC-AGB star sample. Again, we have chosen only oxygen-rich Mira variables in or- der to make direct comparisons with the Glass-LPV sample. We have put the LMC LPVs at the distance of the GC using a dis- tance modulus difference of ∆m = 4.0 mag. Within the errors, the Glass-LPVs do follow the LMC relation. As mentioned, the scatter in the period-magnitude relations for the Glass-LPVs is much higher than in the LMC-AGB sam- ple. This is due to the uncertainties in the extinction determina- tion. No attempt has been made to compensate for the triaxial distribution of stars in the Bulge. Recent studies of the central region of our Galaxy suggest a smaller bar-like structure within the main bar (see e.g. Alard 2001, Nishiyama et al. 2005) which might be partly responsible for the observed scatter. However, the parameters of the central bar (such as its angle, length and orientation) are still poorly known. In contrast to the light curves obtained by Glass et al. (2001), we have only single epoch measurements in the IRAC bands. The K amplitude distribution of the Glass-LPVs peaks around 0.8 mag, going up to 1.5 mag for the OH/IR stars. In order to es- timate the expected amplitude of LPVs in the IRAC filter bands, we used the calculated colours of Galactic oxygen-rich AGB stars by Marengo et al. (2006). Several such objects were ob- served by ISO at different epochs. Mira variables, such as Z Cyg, can show variations up to 1 mag in each IRAC channel, while in the most extreme cases, such as the dust-enshrouded OH/IR star V1300 Aql, variations of 1.4 mag may occur in each filter. Up to now, period-magnitude relations have been mainly studied in the intermediate Bulge (such as Baade's Window) and the LMC, where the log P vs KS relations seem to be in- dependent of the galactic environment (see e.g. Schultheis et al. 2004, Glass & van Leeuwen 2007). Groenewegen & Blommaert (2005), by using OGLE data, also compared the period- magnitude relations in the Galactic bulge and the Magellanic Clouds and concluded that there is no difference in their slopes. Rejkuba (2004) found that even Miras in Cen A follow a similar relation. The Galactic Center, however, has up to now been excluded from this comparison. In asking whether substantially different behaviour should be expected we first summarize recent abun- dance studies in order to place the GC in context with the Bulge ([Fe/H] (cid:39) −0.1; alpha-enhanced) and the LMC ([Fe/H] (cid:39) −0.4; no alpha-enhancement). Studies have shown that in the GC the mean [Fe/H] is nearly solar; [Fe/H] (cid:39) 0.1 (Ram´ırez et al. 2000; Najarro et al. 2004, 2008; Cunha et al. 2007). On the other hand, Ram´ırez et al. (2000) found that the distribution in metallicity is significantly narrower than in the Bulge, emphasizing the differ- ence between its stellar population and that of the GC. Figure 9 shows clearly that the slopes of the period- magnitude relations at the IRAC wavelengths for the Glass- LPVs are similar to those of the LMC. This suggests that any dependence of the log P vs. IRAC relations on abundance, if present at all, must be small. Within the uncertainties, there is no evidence for a significant difference between the period- magnitude relations in the LMC and in the GC. This is in agree- ment with Whitelock et al. (2008) who found, after reanalyz- ing published lightcurves of AGB variables in the LMC, similar zero points in the period-MK relations for systems with differ- ent metallicities. They did not, however, include in their analysis the variables towards the GC, which is thought to be the most metal-rich environment. Thus, our analysis shows that the appar- ent universality of the PL relations extends to the IRAC bands. 10 Schultheis et al.: Interstellar Extinction and Long-Period Variables in the Galactic Center Fig. 9. Dereddened IRAC magnitudes (using the extinction map as described in Sect. 3) vs. log P for LPVs in the Glass et al. (2001) GC field (Glass-LPVs; black dots), where P is the period. Open magenta squares are LMC-AGB stars (oxygen-rich Mira variables). The straight red line is a least-squares fit to the LMC-AGB data set. 6. Mass-loss AGB stars contribute more than 70% towards the enrichment of the dust component of the interstellar medium (ISM) in the so- lar neighbourhood (Sedlmayr 1994) and hence it is important to study their mass-loss in other parts of the Galaxy. One of the most promising tools for determining mass-loss rates is the combination of near-IR and mid-IR colours such as the IRAS K0 − [12] or the ISOGAL K0 − [15] colours (see e.g. Whitelock et al. 1994; Habing 1996; Le Bertre & Winters 1998; Omont et al. 1999; Jeong et al. 2002, Ojha et al. 2003, 2007). We used the dust radiative transfer code for oxygen-rich AGB stars from Groenewegen (2006) to derive theoretical mass- loss rates. The procedures described by Ojha et al. (2007) to de- Schultheis et al.: Interstellar Extinction and Long-Period Variables in the Galactic Center 11 rive observed mass-loss rates based on the (KS − [15])0 colour (see Equ. 1). log M = −8.6171 + 0.85562 x −0.064143 x2 + 0.0018083x 3(1) where x = (K − [15])0 As in Ojha et al. (2007) we are considering an AGB star model with Teff = 2500 K and 100% silicate composition. Note that the empirical values determined by Groenewegen (2006) are for mass-loss rates smaller than 2 × 10−5 M(cid:12) yr−1. For the high mass-loss end we extrapolate the empirical relation. The distribution of mass-loss rates is displayed in Fig. 10. The derived values of M range from 10−8 to 10−5 M(cid:12) yr−1. The distribution, however, is definitely incomplete for M < 1 × 10−7 M(cid:12) yr−1. We emphasize that the determination of these rates is very uncertain as discussed by Schultheis et al. (2003). The un- certainties could easily reach a factor of 3. Fig. 11. Histogram of luminosities of the GC Glass-LPV sample (solid line). The luminosities for an inner Bulge sample (Ojha et al. 2007) is shown for comparison (dotted lines). The vertical lines mark the tip of the RGB (at fainter Mbol) and the tip of the AGB (at brighter Mbol). We used an approximate relation between BCK and (K -- [15])0 derived by Ojha et al. (2003) for AGB stars, using data from Frogel & Whitford (1987). We refer here to Ojha et al. (2003) for a more detailed description of the method. These authors con- clude that the rms uncertainty in the luminosities of the Bulge sources that they observed with ISO is about 0.5 mag. Fig. 11 shows the luminosity function of the Glass-LPVs. These include high luminosity stars, with Mbol < −6, which are not observed in the inner Bulge AGB stars (Ojha et al. 2007). However, one should note that the sample of Ojha et al. (2007) excludes the GC region and are only single-epoch measure- ments. The luminosity distribution of the Glass-LPVs, compared to that of the inner Bulge AGB stars (Ojha et al. 2007), is also broader and less strongly peaked. Due to the variability of our sources, the luminosity distribution will be widened. Thus, com- pared to the Bulge sample of Ojha et al., we find an excess in the GC of luminous stars. Such an excess of luminous stars in the GC, compared to Bulge fields such as Baade's Window, has previously been found in the infrared stellar luminosity function by Blum et al. (1996), Narayanan et al. (1996), Launhardt et al. (2002), Genzel et al. (2003), Van Loon et al. (2003), and Figer et al. (2004). In ad- dition, Blommaert et al. (1998) and Wood et al. (1998) found a similar excess of high luminosity OH/IR stars in the GC com- pared to the Bulge. All these authors conclude that this is gen- erally evidence for recent star formation in the GC. Blum et al. (1996), Figer et al. (2004), and Maness et al. (2007), among oth- ers, have derived detailed star formation histories for the GC by fitting models to the observations. Their studies suggest that star formation has occurred within the last 10 Myr within the cen- Fig. 10. Distribution of mass-loss rates for the GC Glass- LPV sample. The mass-loss rates are inferred from models (Groenewegen 2006) using the dereddened K−[15] colour. There is some correlation between mass-loss rates and period, in the sense that longer periods show higher rates. Theoretical isochrones by Marigo et al. (2008) which include mass-loss and pulsation periods predict that there should be such a correlation; however there is still some disagreement with the observational data indicating that self-consistent dust models of oxygen-rich AGB stars are in need of further improvement. 7. Luminosities We have derived a bolometric magnitude (Mbol) for each Glass- LPV from its de-reddened K magnitude, an assumed bolometric correction (BCK), and a distance modulus of 14.5 for the GC. 12 Schultheis et al.: Interstellar Extinction and Long-Period Variables in the Galactic Center tral 30 pc, and that it has been more or less continuous over the Hubble time. Groenewegen & Blommaert (2005) studied the Mira period distribution of six fields at similar longitudes but spanning lat- itudes −5.8 < b < −1.2 o. They found for all these fields a similar period distribution, consistent with the average period in Baade's Window of ∼333 d (Glass et al. 1995). By comparing their fields with the LPVs of Glass et al. (2001) they found that there is a significant overabundance of LPVs with periods >∼500 d in the GC compared to Bulge fields at higher latitudes. This excess of LPVs with long periods in the GC is not surprising, considering the excess of luminous sources in the GC generally (see Fig. 11), when combined with the period-luminosity rela- tion. Groenewegen & Blommaert (2005) explain this difference in the period distribution as due to a younger population in the GC with a higher initial mass in the range of 2.5 -- 3 M(cid:12) compared to 1 -- 2 M(cid:12) for higher latitude Bulge fields. We conclude that the GALCEN-LPVs show an excess of lu- minous, young stars compared to the galactic Bulge. It is un- fortunate that the isochrones in the [IRAC] bands are not very sensitive to metallicity and/or age; thus no estimation of initial masses or metallicities in this wavelength range can be made. 8. Conclusions We have presented a high resolution interstellar extinction map of the Galactic Center (GC), derived from the Spitzer IRAC catalogue of GC point sources (Ram´ırez et al. 2008). It com- bines observations of the RGB/AGB population with the newest isochrones (Marigo et al. 2008). We have improved on previ- ous near-IR extinction maps by using extinctions derived from mid-IR colours to resolve all regions of high AV. The maximum value we found is AV (cid:39) 90. By using this extinction map, we have studied the Spitzer (IRAC) properties of the long-period variables (LPVs) of in the GC. They follow well-defined period- magnitude relations in the IRAC bands and are similar to those observed in the LMC, lending support to the suggestion that the log P vs. [IRAC] relations are universal, independent of metal- licity. Further, mass-loss rates for the Glass-LPVs have been de- termined by using additional data from ISOGAL. These LPVs show some correlation between mass-loss and pulsation pe- riod as predicted by theoretical isochrones (Marigo et al. 2008). Finally, the long periods of the Glass-LPVs as well as their lu- minosity function agree with the suggestion of Groenewegen & Blommaert (2005) that one is dealing with a young and more massive population in the GC. Acknowledgements. We want to thank the referee J. Blommaert for his very fruitful comments. We are thankful to M. Marengo for making the colours of AGB stars in the IRAC bands available to us, and to the GLIMPSE team for sharing data in advance of publication. We would like to thank M. Groenewegen for the fruitful discussion and his comments on the paper This work is based on observations made with the Spitzer Space Telescope, which is operated by the Jet Propulsion Laboratory (JPL), California Institute of Technology under NASA contract 1407. K.S., R.S., and S.S. are grateful for financial support from NASA through an award issued by JPL/Caltech. ISG acknowledges receipt of a travel grant under the CNRS-NRF bilateral agreement. M.S. and L.G. ac- knowledge support by the University of Padova (Progetto di Ricerca di Ateneo CPDA052212). SG's visit to the Observatoire de Besancon was supported by an EARA-Marie Curie fellowship. Work at the Physical Research Laboratory is supported by the Department of Space, Govt of India. This research has made use of the NASA/IPAC Infrared Science Archive, which is operated by JPL, California Institute of Technology, under contract with the National Aeronautics and Space Administration. This publication makes use 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 L.E., Calvet N., Alessio P., et al., 2004, ApJS 154, 363 Alard C., 2001, A&A 379, L44 Bandyopadhyay R.M., Eikenberry S.S., Curtis D., 2008, astro-ph/0803.2721 Benjamin R.A., Churchwell E., Babler B.L. et al., 2003, PASP 115, 953 Blommaert J.A.D.L., van der Veen W.E.C.J., van Langevelde H.J. et al., 1998, A&A 329, 991 Blommaert J.A.D.L.L, Groenewegen M.A.T., Okumura K., 2006, A&A 460, 555 Blum R.D., Sellgren K., DePoy D.L. 1996, ApJ 470, 864 Cesarsky C.J., Abergel A., Agnese P., et al., 1996, A&A 315, L32 Chapman N.L., Mundy L.G., Shih-Ping L., 2008, astro-ph/0809.1106 Cioni M.R., Marquette J.B., Loup C., et al., 2001, A&A 377, 945 Cunha K., Sellgren K., Smith V.V., et al., 2007, ApJ 669, 1011 Deguchi S., Fuji T., Ita Y., et al., 2008, in "Mapping the Galaxy and Nearby Galalxies", Astrophysics and Space Science Proceedings, Springer, p33 Dutra C.M., Bica E., 2001, A&A 376, 434 Figer D.F., Rich R.M., Kim S.S., et al. 2004, ApJ 601, 319 Flaherty K.M., Pipher J.L., Megeath S.T., et al., 2007, ApJ 663, 1069 Frogel, J. A., Tiede, G. P., & Kuchinski, L. E. 1999, AJ, 117, 2296 Frogel J.A., Whitford A.E., 1987, ApJ 320, 199 Ganesh S., Omont A., Baliyan K.S., et al., 2008, astro-ph/0809.1741 Genzel R., Schodel R., Ott T., et al. 2003, ApJ 594, 812 Glass I.S., Whitelock P.A., Catchpole R.M., et al., 1995, MNRAS 273, 383 Glass I.S., Matsumoto S., Carter B. S., Sekiguchi K., 2001, MNRAS 321, 77 Glass I.S., van Leeuwen F., 2007, MNRAS 378, 1543 Gosling A., Blundell K.M., Bandyopadhyay R., 2006, AJ, 640, L171 Groenewegen M.A.T., 2006, A&A 448, 181 Groenewegen M.A.T., Blommaert J.A.D.L., 2005, A&A 443, 143 Habing, H. J. 1996, A&AR 7, 97 Indebetouw R., Mathis J.S., Babler B.L., et al., 2005, ApJ 619, 931 Jeong, K. S., Winters, J. M., Le Bertre, T., et al. 2002, in Mass-losing Pulsating stars and their Circumstellar Matter, ed. Y. Nakasa, M. Honma, & M. Seki Jiang B.W., Omont A., Ganesh S., et al., 2003, A&A 400, 903 Launhardt R., Zylka R., Mezger P.G., 2002, A&A 384, 112 Le Bertre, T., & Winters, J. M. 1998, A&A, 334, 173 Le Bertre T., Tanaka M., Yamamura I., et al., 2003, A&A, 403, 943 Lindqvist M., Winnberg A., Habing H.J., et al., 1992, A&AS 92, 43 Lutz D., Feuchtgruber H., Genzel R., et al., 1996, A&A 315, L269 Lutz D., 1999, in "The Universe as Seen by ISO", ESA-SP 427, p.623 Maness H., Martins F., Trippe S., et al. 2007, ApJ 669, 1024 Marengo M., Hora J.L., Barmby P., astro-ph/0611346 Marigo P., Girardi L., Bessan A., et al., 2008, A&A 482, 883 Marshall D.J., Robin A.C., Reyl´e C., 2006, A&A 453, 635 Meixner M., Gordon K.D., Indebetouw R., et al., 2006, AJ 132, 2268 Messineo M., "Late-type Giants in the Inner Galaxy", PhD thesis, Leiden Najarro F., Figer D.F., Hillier D.J., et al., 2004, ApJ 611, L105 Najarro F., Figer D.F., Hillier D.J., et al., 2008, astro-ph/0809.3185 Narayanan V., Gould A., DePoy D.L. 1996, ApJ 472, 183 Nishiyama S., Nagata T., Baba D., et al., 2005, ApJ 621, L105 Nishiyama S., Nagata T., Matsunaga N., 2006, ApJ 638, 839 Ojha D. K., Omont A., Schuller F., et al. 2003, A&A, 403, 141 Ojha D.K., Tej A., Schultheis M., et al., 2007, MNRAS 381, 1219 Omont A., Ganesh S., Alard C., et al. 1999, A&A, 348, 755 Ram´ırez S.V., Sellgren K., Carr J.S., et al., 2000, ApJ 537, 205 Ram´ırez S.V., Arendt R.G., Sellgren K., et al., 2008, ApJS 175, 147 Reid M.J., 1993, ARA&A 31, 345 Rejkuba M., 2004, A&A 413, 903 Rom´an-Z´uniga C.G., Lada C.J., Muench A., 2007, ApJ 664, 357 Schuller F., Ganesh S., Messineo M.., et al., 2003, A&A 403, 955 Schuller F., Omont A., Glass I.S., et al., 2006, A&A 453, 535 Schultheis M., Ganesh S., Glass I.S., et al., 1999, A&A 349, L69 Schultheis M., Lanc¸on A., Omont A., et al., 2003, A&A 405, 531 Schultheis M., Glass I.S., Cioni M.R., 2004, A&A 427, 945 Sedlmayr, E. 1994, in Molecules in the Stellar Environment, ed. U. G. Jorgensen (Berlin: Springer), 163 Sjouwerman L., van Langevelde H.J., Winnberg A., 1996, A&AS 128, 35 Stolovy S., Ram´ırez S.V., Arendt R.G et al., 2006, Journal of Physics: Conference Series, Volume 54, 176 Stolovy S., et al., 2008, in preparation Vanhollebeke E., Blommaert J.A.D.L, Schultheis M., et al., 2006, A&A 455, 645 van Loon, J., Gilmore G., Omont A., et al., 2003, MNRAS 338, 857 Whitelock P.A., Menzies J.W., Feast M.W., et al., 1994, MNRAS 267, 711 Whitelock P.A., Feast M.W., van Leeuwen F., 2008, MNRAS 386, 313 Wood P.R., Habing H.H., McGregor P.J., 1998, A&A 336, 925
astro-ph/0604099
2
0604
2006-07-13T15:14:02
Unexpected Dynamical Instabilities In Differentially Rotating Neutron Stars
[ "astro-ph" ]
A one-armed spiral instability has been found to develop in differentially rotating stellar models that have a relatively stiff, $n=1$ polytropic equation of state and a wide range of rotational energies. This suggests that such instabilities can arise in neutron stars that are differentially, although not necessarily rapidly, rotating. The instability seems to be directly triggered by the presence of a corotation resonance inside the star. Our analysis also suggests that a resonant cavity resulting from a local minimum in the radial vortensity profile of the star plays an important role in amplifying the unstable mode. Hence, it appears as through this instability is closely related to the so-called ``Rossby wave instability'' \citep{LLCN99} that has been found to arise in accretion disks. In addition to the one-armed ($m=1$) spiral mode, we have found that higher-order ($m = 2$ and $m=3$) nonaxisymmetric modes also can become unstable if corotation points that resonate with the eigenfrequencies of these higher-order modes also appear inside the star. The growth rate of each mode seems to depend on the location of its corotation radius with respect to the vortensity profile (or on the depth of its corotation radius inside the vortensity well). The existence of such instabilities makes the stability criterion for differentially rotating neutron stars non-unique. Also, the gravitational-waves emitted from such unstable systems generally will not have a monochromatic frequency spectrum.
astro-ph
astro-ph
Unexpected Dynamical Instabilities In Differentially Rotating Neutron Stars Shangli Ou and Joel E. Tohline Department of Physics & Astronomy, Center for Computation & Technology, Louisiana State University, Baton Rouge, LA 70803 ABSTRACT A one-armed spiral instability has been found to develop in differentially ro- tating stellar models that have a relatively stiff, n = 1 polytropic equation of state and a wide range of rotational energies. This suggests that such instabil- ities can arise in neutron stars that are differentially, although not necessarily rapidly, rotating. The instability seems to be directly triggered by the presence of a corotation resonance inside the star. Our analysis also suggests that a reso- nant cavity resulting from a local minimum in the radial vortensity profile of the star plays an important role in amplifying the unstable mode. Hence, it appears as through this instability is closely related to the so-called "Rossby wave insta- bility" (Lovelace et al. 1999) that has been found to arise in accretion disks. In addition to the one-armed (m = 1) spiral mode, we have found that higher-order (m = 2 and m = 3) nonaxisymmetric modes also can become unstable if corota- tion points that resonate with the eigenfrequencies of these higher-order modes also appear inside the star. The growth rate of each mode seems to depend on the location of its corotation radius with respect to the vortensity profile (or on the depth of its corotation radius inside the vortensity well). The existence of such instabilities makes the stability criterion for differentially rotating neutron stars non-unique. Also, the gravitational-waves emitted from such unstable systems generally will not have a monochromatic frequency spectrum. Subject headings: neutron stars -- one-armed instabilities -- corotation -- dif- ferential rotation -- vortensity -- Rossby wave instability -- gravitational waves 1. Introduction Rotating stars are subject to a variety of nonaxisymmetric instabilities (Chandrasekhar 1969; Tassoul 1978; Andersson 2003). If a star rotates sufficiently fast -- to a point where the -- 2 -- ratio of its rotational to gravitational potential energy T /W & 0.27 -- it will encounter a dynamical bar-mode instability that will result in the deformation of the star into a rapidly spinning, bar-like structure. This instability has been verified by a number of numerical hydrodynamics investigations (Tohline, Durisen, & McCollough 1985; Durisen et al. 1986; Williams & Tohline 1988; Cazes & Tohline 2000; New, Centrella & Tohline 2000; Brown 2000; Liu 2002). At a slower rotation rate, T /W & 0.14, a star may also encounter a secular bar-mode instability that can promote deformation into a bar-like shape, but only if the star is subjected to a dissipative process capable of redistributing angular momentum within its structure, such as viscosity or gravitational radiation reaction (GRR) forces (Chandrasekhar 1970; Friedman & Schutz 1978; Ipser & Lindblom 1990, 1991; Lai & Shapiro 1995). In reality, viscosity and GRR forces tend to compete with each other and can drive the secularly unstable star along a variety of different evolutionary paths. Recently, Shibata & Karino (2004) and Ou, Tohline & Lindblom (2004) used numerical hydrodynamic techniques to study the nonlinear evolution of the GRR-driven secular bar-mode instability in rotating, "polytropic" neutron stars in the absence of any competing dissipation. The above studies have shown that the critical limits of both the secular and dynamical bar-mode instabilities do not depend sensitively on the degree of compressibility of the equation of state (EOS) or on the differential rotation law of a star. Surprisingly, several recent numerical studies have found that dynamical instabilities that excite m = 1 as well as m = 2 azimuthal Fourier modes can arise in self-gravitating structures having T /W < 0.27, that is, having a rotational energy below the threshold tra- ditionally expected for dynamical nonaxisymmetirc instabilities. Tohline & Hachisu (1990) found an m = 2 instability in self-gravitating rings and tori having uniform specific angular momentum and T /W & 0.16. Centrella et al. (2001) found a one-armed (m = 1) spiral instability in stellar models with a polytropic index n = 3.33, strong differential rotation and T /W ∼ 0.14. Shibata et al. (2002, 2003) found an m = 2 instability in extremely differentially rotating stellar models having T /W on the order of 10−2! Saijo et al. (2003) concluded that one-armed (m = 1) modes only appear in structures with a very soft equation of state (n & 2.5), and that a necessary condition for this instability to arise is the presence of an off-center density maximum, that is, a toroidal-like structure. Ott et al. (2005a) showed that a one-armed spiral instability may arise in the context of a pre-supernova core collapse. The linear analysis of Watts et al. (2004) has suggested that these low T /W instabilities are triggered because corotation points of the unstable modes fall within the differentially rotating structure of the examined models. Saijo & Yoshida (2005) also have carried out a linear stability analysis that shows a connection between the one-armed, m = 1 instability and corotation points in differentially rotating stars. Many earlier studies of disks and tori have also noted an association between the onset -- 3 -- of dynamical instabilities and the existence of corotation points. For example, a number of groups (Papaloizou & Pringle 1984; Goldreich, Goodman, & Narayan 1986; Narayan, Goldreich, & Goodman 1987; Frank & Robertson 1988) have used both linear analyses and nonlinear simulations to study the so-called Papaloizou-Pringle instability, which is triggered by corotation resonances. But the existence of a corotation point within a model is not sufficient to ensure significant growth of an unstable mode. It appears as though a resonant cavity is also required to drive these modes to very large amplitude. In the case of the Papaloizou-Pringle instability, the inner and outer edges of the disk or torus form a resonant cavity in which waves are reflected back and forth; the unstable mode is greatly amplified through multiple passages across the corotation radius. However, it is not obvious how such a resonant cavity can exist in stellar models that have no inner edges. Lovelace & Hohlfeld (1978) and Papaloizou & Lin (1989) have discussed how the radial distribution of vortensity or potential vorticity, which is defined as the ratio between vorticity and surface density, is important in determining the stability of homentropic, self-gravitating rings and disks. Lovelace et al. (1999) and Li et al. (2000) extended this discussion to non- homentropic disks. According to their analysis of the so-called "Rossby-Wave Instability", local minima in the radial vortensity profile of homentropic configurations tend to trap radially propagating waves and, hence, can serve as resonant cavities that will help drive corotation instabilities to large amplitude. This removes the necessity of having disk edges to serve as reflecting boundaries. In their subsequent nonlinear simulations, Li et al. (2001) found that this mechanism can aid in the excitation of nonaxisymmetric instabilities (m = 1, 3, 5) and vortex formation in accretion disks. It is interesting to note that when the initial perturbations in these nonlinear simulations contained the pure eigenfunctions for a given azimuthal mode (for example, m = 3 or 5), the corresponding mode developed; but, when a white noise initial perturbation was used, an m = 1 spiral mode became dominant in the final stage. It is conceivable that a similar resonant cavity mechanism may work to amplify corota- tion resonances in stellar models. This would seem to be especially likely for rapidly rotating models that have toroidal-like structures because the associated off-center density maximum will inevitably result in a dip in the radial vortensity profile of the star. This profile could then act as a resonant cavity to trap radially propagating waves. Alternatively, a maximum in the star's radial entropy profile can also produce such a cavity (Lovelace et al. 1999), but in keeping with most previous studies of the low T /W instabilities we will confine our analysis to homentropic configurations where a dip in the vortensity profile will be required to create a resonant cavity. In this paper, we report the unexpected discovery of an m = 1 spiral-mode instability in -- 4 -- rotating stars with a relatively stiff (n = 1) polytropic EOS that was designed to represent differentially rotating neutron stars. We also show evidence that this instability is directly triggered by the existence inside the star of a corotation radius for the eigenmode. This one- armed spiral instability develops on a timescale that is somewhat longer than the growth time observed in previous studies of configurations with a softer EOS, but the instability appears to be dynamical nevertheless. All of our unstable models exhibit a local minimum in their radial vortensity profile and this seems likely to be the resonant cavity that traps the wave and permits it to amplify. (We have found that such a minimum can exist in config- urations with centrally condensed, rather than toroidal, density structures when the degree of differential rotation is sufficiently strong.) Our results suggest that it is the combination of corotation points and resonant cavities formed by vortensity minima that is responsible for the excitation of low T /W instabilities found in stellar models. Perhaps our most striking discovery is that, in addition to the m = 1 mode, m=2 and 3 modes also come into play as long as their corotation points exist within the structure of the star. Therefore, if the rotational profile of the star is steep enough to contain corotation points of multiple modes, these unstable modes may all arise within one single star. The entire picture of rotational instabilities that might arise in differentially rotating neutron stars becomes more complicated due to the co-existence of these corotation instabilities. The corresponding gravitational-wave signals may exhibit different characteristic frequencies at different or overlapping times, since various unstable modes may set in simultaneously. 2. Initial Models and Numerical Methods 2.1. Initial Axisymmetric Models Using the Hachisu self-consistent-field technique (Hachisu 1986), four simplified axisym- metric neutron star models (models J250, J133, J127, and J068 ) were constructed on a cylindrical grid with resolution 66 × 82 × 128 in the radial, vertical and azimuthal direc- tions, respectively. For each model, we have adopted a dimensionless unit system in which the gravitational constant G, the maximum density ρmax, and the radius of the entire grid Rgrid are set to 1. For comparison purposes, equatorial radii Req were all set to 0.673. All the models were constructed using Newtonian gravity and a polytropic equation of state, p = Kρ1+1/n, where p is the gas pressure, K is a polytropic constant, and the polytropic index n was in all cases chosen to be 1. The angular velocity profiles were specified by the so-called j-constant-like law as, Ω(R) = ωcA2 R2 + A2 , (1) -- 5 -- where R = px2 + y2 is the cylindrical radius, A is a constant, and ωc is the angular velocity at the rotation axis (the Z-axis). The specific parameter values of each model are given in Table 1, where ¯ρ is the mean density of the star, ρc is the central density (which will be less than ρmax if the model exhibits an off-center density maximum), and Rp is the polar radius. 2.2. Analysis Method We used a three-dimensional (3D), Newtonian hydrodynamics code essentially the same as the one employed by Ou, Tohline & Lindblom (2004, hereafter Paper I) to study the evo- lution of our differentially rotating, neutron stars models. The details of the hydrodynamics code can be found in Motl, Tohline, & Frank (2002). In paper I, an artificially enhanced gravitational-radiation-reaction (GRR) potential was introduced into the code to mimic the GRR effect, which will induce the secular bar-mode instability when the T /W value of the star exceeds the critical limit 0.14. In order to monitor the growth of nonaxisymmetric modes, we measured the Fourier amplitude of each mode with azimuthal quantum number m in the following fashion: At any point in time during an evolution, the azimuthal density distribution in a ring of fixed R and z can be decomposed into a series of azimuthal Fourier components via the relation, ρ(R, z, φ) = +∞ Xm=−∞ Cm(R, z)eimφ , where the complex amplitudes Cm are defined by the expression, Cm(R, z) = 1 2π Z 2π 0 ρ(R, z, φ)e−imφdφ . (2) (3) In our simulations, the time-dependent behavior of the magnitude of this coefficient, Cm, is then monitored to measure the growth-rate of various unstable modes. At any point in time, the ratio between the imaginary and real parts of Cm(R, z) gives us a spatially dependent phase angle φm(R, z) from which we also are able to determine the azimuthal structure of each unstable mode m, if it exists. (In particular, the radial dependence of φm tells us, for example, whether an m = 1 mode has a spiral character or an m = 2 mode has a bar-like structure; see Figures 14 and 15, below.) In general, φm(R, z) is a function of time, but an eigenmode of the system becomes identifiable when the spatial structure of the phase angle φm(R, z) appears to be coherent/aligned throughout the star and time-independent, apart from a spatially independent azimuthal eigenfrequency ωm ≡ ∂φm/∂t of the mode. Another way to look at this is that the oscillation period of mode m, Tm ≡ 2π/ωm, is -- 6 -- equivalent to the time it takes the real or imaginary parts of Cm to complete a full cycle from a positive value to a negative value, then back to a positive value. In our nonlinear hydrodynamic simulations, the measurement of ωm is accurate only when the corresponding mode dominates the evolution. By definition, the corotation radius Rcor of mode m is the radial location in the star where the angular velocity of the fluid resonates with the eigenfrequency of the unstable mode, that is, where ω(Rcor) = ωm/m. In our analysis, we also will find it useful to refer to each model's central spin period, Tc = 2π/ωc, and characteristic dynamical time scale, T0 ≡ 2π/ω0, where ω0 ≡ √πG¯ρ. 3. Simulation Results In this section, we show results from a number of independent 3D hydrodynamic evo- lutions. Because our initial goal was to study the secular bar-mode instability induced by GRR forces, rather than the development of a one-armed spiral instability, the initially ax- isymmetric density structure of all models was perturbed with an m = 2 perturbation with different amplitude as they were introduced into the hydrodynamics code. In the discussion that follows, the evolution time of individual simulations has been normalized to the spin period at the rotation axis of each star Tc, which varies significantly for different models, but in terms of the characteristic dynamical time, T0, they were evolved for roughly the same times, except for model J127. Table 2 summarizes all the relevant frequencies and time scales for each of our models. In addition to the time scales and frequencies defined earlier, τm is the measured e-folding (growth) time for a given mode m, and ωs is the angular frequency at the equatorial surface. We note that, for a given model, the eigenfrequencies quoted in Table 2 have been measured at different evolutionary times. Each ωm is measured at a time when the corresponding mode m is dominant, and the corresponding corotation radius is determined from the system's equatorial plane rotational profile at that same time. 3.1. Evolution of model J250 This investigation began as an extension of Paper I (where uniformly rotating neutron stars were studied) and our initial aim was to study the secular bar-mode instability in differentially rotating neutron stars. The first model we evolved was model J250, which was expected to be secularly unstable but dynamically stable to the high T /W bar-mode instability. Because the model had a relatively large value of the parameter A, it also -- 7 -- Table 1. Parameters of Initial Models Model n A/Req Rp/Req Mtot ρc/ρmax ¯ρ ωc J250 J133 J127 J068 1.0 1.0 1.0 1.0 0.94 0.94 0.59 0.44 0.302 0.558 0.558 0.721 0.204 0.257 0.335 0.384 0.976 0.999 0.999 1.000 0.384 0.309 0.363 0.343 1.44 1.19 1.75 1.80 T /W 0.250 0.133 0.127 0.068 unstable modes m=1 - m=1,2,3 m=1,2 Note. -- n is polytropic index, A is a parameter that determines the degree of differential rotation, Req is the equatorial radius, Rp is the polar radius, Mtot is the total mass, ρc is the central density, ρmax is the maximum density, ¯ρ is the mean density, ωc is the angular frequency at the center, T is the rotational energy, W is the gravitational energy. Table 2. Relevant Frequencies of All Models Model ωc ωs ω0 = √πG¯ρ ω1 ω2/2 ω3/3 τ1/T0 τ2/T0 τ3/T0 J250 J133 J127 J068 1.44 1.19 1.75 1.80 0.66 0.54 0.44 0.29 1.10 0.98 1.07 1.04 1.2 - 1.3 1.3 - - 0.9 0.9 - - 1.0 - 9.4 - 29.4 8.4 - - 7.8 4.4 - - 4.2 - Note. -- ωc is the angular frequency at the center, ωs is the angular frequency at the equatorial surface, ¯ρ is the mean density, ωm is the eigenfrequency of the azimuthal mode m, τm is the e-folding time of the azimuthal mode m, T0 = 2π/ω0. As illustrated in Figure 9, for a given model, the eigenfrequencies ωm have been measured at different evolutionary times, when each mode dominates. -- 8 -- was expected to be stable to the low T /W dynamical bar-mode instability identified by Shibata et al. (2002). Model J250 was evolved for many dynamical times with an artificially enhanced GRR potential, in a manner similar to the evolutions described in Paper I. The top panel of Figure 1 shows the time-evolution of the amplitudes of different modes (averaged over the whole volume) throughout this evolution. In the beginning, the m = 2 bar-mode grows exponentially on a timescale governed by the artificially enhanced GRR potential; but it saturates after only a moderately deformed bar structure has formed. This is consistent with the results reported by Shibata & Karino (2004). However, on a considerably longer time scale, an unexpected m = 1 mode arises and becomes the dominant mode. The density contour plots in the bottom right-hand panel of Figure 1 suggest that the m = 1 mode has a one-armed spiral structure similar to the one discovered by Centrella et al. (2001). We note that, late in the evolution, the amplitude of the m = 2 mode increases when the m = 1 mode reaches the nonlinear regime. (This feature is common to all of our models that are unstable to the m = 1 mode.) Measurements of ωm for various modes at the time when the m = 1 mode is dominant confirms that, for example, ω2 ≈ 2ω1 ≈ 2.6 and suggests that, at late times, the m = 2 mode is a harmonic of the m = 1 mode. This discovery is unexpected in the sense that the one-armed, m = 1 instability had previously appeared only in models with a softer EOS (n & 2.5) and with a high degree of differential rotation, i.e., ωc/ωs ≈ 10 (Centrella et al. 2001; Saijo et al. 2003; Saijo & Yoshida 2005), where ωs is the angular frequency at the equatorial surface. In contrast, our model J250 has an n = 1 polytropic EOS and has only a moderate degree of differential rotational with ωc/ωs ≈ 2. In order to test whether or not this instability was caused by the artificially enhanced GRR forces, a second evolution was performed on model J250 with GRR turned off. Figure 2 shows the time-evolution of the amplitudes of different modes for this second evolution. As expected, the m = 2 bar-mode was not unstable in the absence of GRR forces, but surprisingly the m = 1 mode remained unstable, becoming the dominant Fourier component at time & 100 Tc. Compared to models that had a softer EOS (Centrella et al. 2001; Saijo et al. 2003) in which the m = 1 mode grew on a time scale of a few tens of Tc, the growth time here is much longer but nevertheless still dynamical. (We note that the growth time in this pure hydrodynamical evolution is also longer than that from the previous run, in which GRR was turned on. This is probably because GRR tends to increase the degree of differential rotation, hence influence the growth rate of the m = 1 mode.) The top panel of Figure 3 shows the radial eigenfunction of the m = 1 mode obtained by subtracting the initial axisymmetric background density (bottom panel of Figure 3). Its behavior is qualitatively similar to that derived from a linear stability analysis by Saijo & Yoshida (2005) (see their -- 9 -- Figure 9). The appearance of the one-armed spiral instability in our models with a stiff EOS seems to suggest that its stability criterion is independent of compressibility. Now the question is what mechanism is responsible for its growth? According to Watts et al. (2004), Saijo & Yoshida (2005), and related studies of disks and tori (Papaloizou & Pringle 1984; Goldreich, Goodman, & Narayan 1986; Narayan, Goldreich, & Goodman 1987), corotation points may be responsible for the amplification of some unstable modes. But this mechanism of amplification can only operate if there is at least one radial location in the model where the angular frequency of orbiting fluid elements resonates with the eigenfrequency of the identified special mode. A straightforward check of the radial rotational profiles Ω(R) of model J250 (see Figure 4) shows that a corotation radius is present at R ≈ 0.32 for the unstable m = 1 Fourier mode with an eigenfrequency ω1 ≈ 1.2. (This measurement of the eigenfrequency was done for the pure hydrodynamical simulation, in which GRR was turned off.) However, the result from one model is not sufficient for us to draw a definite conclu- sion. To further clarify the relationship between the instability and corotation points, we performed a number of simulations on different polytropic neutron star models. In these additional simulations, we will concentrate only on hydrodynamic simulations with GRR turned off, because GRR-related instabilities generally would set in on a much longer time scale and therefore are irrelevant here. 3.2. Evolution of model J133 So far we have identified a corotation point associated with the one-armed spiral insta- bility in model J250. It would be interesting to find out if the instability disappears when there is no such corotation point. In order to test this, we chose to evolve model J133. This model is rotating slowly enough so that the central angular frequency is well below the eigen- frequency of the m = 1 mode that arose in model J250. By removing the corotation points of the m = 1 mode inside model J133, it was hoped that the one-armed instability would not show up in this evolution. To eliminate the possible influence of the degree of differential rotation, we kept the value of the parameter A the same as in model J250, but increased the polar radius Rp to slow down the rotation of the star. Note that the eigenfrequency of the m = 1 mode for this model is actually unknown1, but we estimate it would not change 1Our studies do not include a method for performing a complete eigenmode analysis of our selected, initially axisymmetric models. Only modes that are unstable and that grow on a timescale short compared -- 10 -- dramatically compared to that of model J250. (In fact, as shown by later simulations, ω1 in model J250 and J127 are very close to each other; see Table 2.) As shown in the top panel of Figure 5, the m = 1 Fourier amplitude C1 does not grow into the nonlinear regime in this case. It develops somewhat, but only hangs on at a very small amplitude. On the other hand, the m = 2 mode seems to gain some strength in the beginning of the evolution, but it quickly saturates at a very low amplitude and never enters into the nonlinear regime. The density contours shown in the bottom panels of Figure 5 also display no indication of nonaxisymmetric structure. The lack of development of the m = 1 spiral mode in this simulation supports our conjecture that corotation points are directly responsible for the excitation of the mode. 3.3. Evolution of model J127 Although the evolution of model J133 makes our argument about corotation points plausible, the fact that its T /W value is significantly lower than model J250 raised the following question: is it the low T /W value that suppressed the instability? To answer this question and as a further check of our conjecture, model J127 was evolved in the same fashion as model J133. This model has a T /W value that is similar to that of model J133, but its degree of differential rotation was made very high by reducing the value of A. With this steep rotational profile, we hoped that a corotation point for the m = 1 mode would exist within the star and, hence, trigger the instability. The final outcome of this evolution is perhaps the most striking and interesting one. As shown in the top panel of Figure 6, the evolution can be divided into three stages: (1) In the first ∼ 50 Tc, an m = 2 mode grows exponentially, but it saturates very quickly; (2) ∼ 30 Tc later, an unexpected m = 3 mode comes into play, and becomes dominant for quite a long time. The pear-shaped m = 3 density distortion is visible in the bottom left-hand panel of Figure 6; (3) after ≈ 300Tc, the originally expected m = 1 mode catches up and wins over. Given previous studies of the low T /W bar-mode instability (Shibata et al. 2002), the appearance of the m = 2 mode here is not surprising. However, the appearance of an m = 3 mode in axisymmetric equilibrium stellar models with intermediate T /W values is totally unexpected. A plot of the rotational profile of model J127 is shown in Figure 7 and reveals the existence of corotation radii for all three of these unstable modes. These results are fully consistent with the linear analysis presented by Watts et al. (2004): Corotation instabilities to our evolution times can be identified using our present nonlinear hydrodynamic tools. -- 11 -- in stellar models are not limited to m = 1 and m = 2 modes. Other modes may become unstable as long as their corotation points exist inside the star. In this simulation, we also analyzed the behavior of the m = 4 (5) Fourier component of the density distribution. It appears to follow the behavior of the m = 2 (3) mode, but it saturates at an amplitude that is two orders of magnitude lower. This suggests that the m = 4 (5) "mode" is actually a harmonic of m = 2 (3). There is no indication of the independent development of either an m = 4 or 5 mode. We suspect that the reason for the lack of independent m=4 and 5 modes is that their corotation radii fall outside the star. Because instabilities associated with different modes may arise in stars with strong differential rotation, the gravitational-wave signals from such an event would exhibit multiple characteristics. Figure 8 shows the computed quadrupole gravitational-wave strain as viewed by an observer located along the +Z axis and assuming the neutron star has a mass of 1.4 M⊙ and an equatorial radius of 12.5 km. The peak signal strength is actually stronger than the bounce signal expected from the axisymmetric core collapse of massive stars, which has been shown to produce signals with rh+ ≈ 300 cm at a frequency f ≈ 400 Hz (Ott et al. 2004). The signal also has some distinct features. There is more than one localized burst, arising from different modes peaking at different times and in different frequency bands: the first, centered at t/Tc ∼ 65, corresponds to rh+ ∼ 500 cm at a frequency f ≈ 1800 Hz emitted by the m = 2 mode; the second, centered at t/Tc ∼ 100, corresponds to rh+ ∼ 600 cm at a frequency f ≈ 1900 Hz associated with the m = 3 mode; the third, centered at t/Tc ∼ 375, corresponds to a rh+ ∼ 900 cm at frequency f ≈ 2600 Hz emitted by the m = 2 harmonic of the m = 1 spiral mode. Figure 9 shows the pattern frequencies of all the dominating unstable modes as a function of time. To make the graph clean, the frequencies of the m = 1 and 2 modes are only shown in the time intervals when they dominate, and all the data at times t < 40Tc have been omitted because they appear to be noisy due to the lack of coherent modes. It is observed that the pattern frequencies of all three modes -- ω1, ω2/2, and ω3/3 -- are fairly constant throughout their peroid of dominance and the uncertainty in these measurements is . 10%. Hence, the relative location of the corotation radii Rcor for these various modes does not vary with time, unless a change in the rotational profile of the star causes Rcor to migrate. Based upon the above hydrodynamic simulations, we can now draw the following con- clusions: (1) The m = 1 instability can be turned on or off by controlling the rotational profile of the star to allow or remove the existence of a corotation point. (2) The instability is not limited to the m = 1 mode; other modes may also arise, as long as their corotation points exist inside the star. -- 12 -- 3.4. Evolution of model J068 According to Shibata et al. (2002, 2003), models with very small values of the rotation- profile parameter A are subject to a low T /W dynamical bar-mode instability. Since the rotational profiles for these types of models are quite steep, it occurred to us that they may also allow corotation points for various modes within their structure. It also seemed likely that even one individual stellar model might be subject to instabilities associated not only with the m = 2 bar-mode, but also other modes. In order to test this idea, we evolved model J068 with A/Req = 0.44 (ωc/ωs ≈ 6), which falls in the instability region where the low T /W dynamical bar-mode instability would be expected to set in. The results of this simulation support our suspicion that an individual model with a very steep rotational profile may be subject to multiple instabilities associated with differ- ent azimuthal modes. As shown in the top panel of Figure 10, at the beginning of this simulation the m = 2 bar-mode grew very quickly, became nonlinear, and dominated the evolution for quite a while, which is consistent with the results presented by Shibata et al. (2002). In addition, however, we found that the underlying m = 1 and 3 modes also grew simultaneously. These odd-numbered modes had growth rates rather slower than the m = 2 mode, but this did not keep the m = 1 mode from finally winning over all the other modes, including the bar-mode, and manifesting itself as the dominant mode, just as it did in our earlier simulations. Notice that, again, the amplitude of the m = 2 mode increases when the m = 1 mode becomes dominant at late times, which suggests that the m = 2 mode is a harmonic of the m = 1 mode during this period. The rotational profile of model J068, shown in Figure 11, reveals corotation points for both m = 1 and 2 modes. (We could not accurately measure the eigenfrequency of the m = 3 mode because it never became the dominant mode in the evolution time followed by our simulation.) 4. Summary and Discussion We have found that a one-armed spiral instability arises in rotating n = 1 "polytropic" neutron stars with moderate and high degrees of differential rotation. This is in contrast to the work of Saijo et al. (2003), which has suggested that such instabilities can arise only in stars with a softer EOS. As has been suggested by Saijo & Yoshida (2005), the instability seems to be associated with the existence of a corotation point of the m = 1 mode that lies inside the star. Given previous results for models with a softer EOS, we conclude that the compressibility and T /W value of the star are not directly responsible for triggering the -- 13 -- instability. We have discovered that, in addition to the m = 1 mode, other higher-order modes also are allowed to come into play within a single model as long as corotation points for these modes also exist within the star. This can happen as long as the rotational profile is sufficiently steep. In our n = 1 polytropic neutron star models, these modes grow rela- tively slowly compared to previous results for softer EOSs (Centrella et al. 2001; Saijo et al. 2003), but nevertheless they appear to be dynamically unstable. Such modes are likely to develop in newly born neutron stars because young neutron stars are expected to have strong differential rotation (Ott et al. 2005b; Akiyama & Wheeler 2005). Also, because they develop in low T /W configurations on a dynamical time scale, these modes are likely to be more important than the often-discussed secular bar-mode instability. This will have important consequences for the detection of gravitational-waves by the Laser Interferome- ter Gravitational-wave Observatory (LIGO), because gravitational-waves generated in such systems would not be monochromatic; our simulations indicate that signals with different characteristic frequencies are likely to peak at different times. We suspect that these corotation-related instabilities have not been identified in previous studies of models that have a stiff EOS because earlier studies have been focused on a search for, or the analysis of, instabilities that develop on a very short time scale. Most previous simulations have been followed through only a few tens of Tc, which is not a long enough time to permit significant amplification of many of the modes we have identified in this study. Only in this study -- which was initially aimed at examining the secular bar-mode instability in stars with differential rotation -- have models been evolved long enough to reveal the corotation instabilities that grow on a longer time scale for this stiff EOS case. Based on theoretical analyses of differentially rotating configurations that have been presented by others, we suspect that, in addition to the existence of corotation points, a resonant cavity is required in order to drive corotation instabilities to very large amplitude. In particular, Lovelace et al. (1999) have suggested that a local minimum in the radial vortensity profile can serve as a resonant cavity to trap waves and, hence, trigger a "Rossby wave" instability. Figures 12 and 13 display the radial vortensity profiles of our models J068 and J127 at different times. (The vortensity plotted here is the ratio between vorticity and volume density, but the same radial profile is preserved even if we define vortensity as the ratio between vorticity and surface density.) As these figures show, a fairly wide dip encompasses the corotation radii of all of the modes that we have found to be unstable in these models. We note that such a dip in the radial vortensity profile is a common feature for all of our unstable models. We suspect that waves are trapped inside these vortensity cavities and are continously amplified through multiple passages across their respective corotation radii. -- 14 -- Because an off-axis density maximum alone will generate a vortensity minimum and hence form a resonant cavity that can trap waves inside, it is perhaps not surprising that toroidal structures frequently have been found to be unstable to a one-armed, spiral instability (Saijo et al. 2003). However, our new results show that a toroidal structure is not necessary to generate a vortensity minimum; strong differential rotation alone can create a vortensity minimum, as indicated for example by our model J068 which does not have a toroidal structure. We have also noticed that an unstable mode whose corotation radius falls at a position that is deeper in the vortensity well has a higher growth rate in any given simulation. Figures 12 and 13, in conjunction with the data provided in Table 2, reveal this correlation between the growth rate of each mode and the depth of its corotation radius inside the vortensity well. This correlation is consistent with the results reported by Li et al. (2000) from a study of the Rossby-wave instability in differentially rotating thin disks. They discovered that the growth rate of the Rossby-wave instability was proportional to the amplitude of density bumps/jumps that were introduced into their models, in other words, proportional to the depth of the vortensity wells that were created by such density jumps. According to Li et al. (2000), requiring a minimum to be present in a star's radial vortensity profile before the star can become susceptible to a corotation instability is a gen- eralization of the Rayleigh inflexion point theorem, which states that a necessary condition for an instability of this type to develop is that there be an inflection point in the velocity profile (Drazin & Reid 1981). Given the j-constant-like rotation law (Eq. 1) adopted in our (as well as previous) studies, the condition ∂2vφ/∂2R = 0 can be evaluated analytically to show that a velocity inflection point arises at R = A/√3. Hence, for a sufficiently shallow rotational profile having A > √3Req, the required inflection point does not exist inside the star. This may also explain why low T /W instabilities only show up in stars with strong differential rotation (smaller values of A). Finally, in an effort to draw a further connection between our results and previous anal- yses, we have examined the azimuthal structure of the unstable eigenmodes in our models. Figure 14 displays the phase angle of the m = 1 mode as a function of R in the equatorial plane of model J250. The pattern shows a slightly trailing segment in the innermost re- gion, then transforms itself into a leading wave outside the density maximum, and finally a loosely wound, trailing segment appears again outside the corotation radius. This par- ticular characteristic is reminiscent of the m = 1 structure that was observed in an earlier study of Papaloizou-Pringle modes by Woodward et. al. (1994), and it is quite similar to the azimuthal structure of the one-armed spiral mode discussed by Saijo & Yoshida (2005). In contrast, Figure 15 shows the phase-angle plot of the m = 2 mode for model J068. No -- 15 -- spiral structure is observed for the m = 2 mode; instead, the phase angles at different radii are roughly aligned along a straight line indicating a purely bar-like deformation. The radial eigenfunction of the m = 1 mode shown in Fig. 3 is not monotonic and has a node between the center and the surface; we suspect it belongs to the family of Papaloizou-Pringle modes that arises in pressure-supported tori and thick accretion disks. The nodeless feature of the radial eigenfunctions for m = 2, 3 modes suggests that they might be f-mode oscillations (this interpretation is consistent with the unstable mode structure presented by Shibata et al. (2003)); but their structures are different in details: the eigenfunction of the m = 2 mode (seen in J068) starts from zero at the center, monotonically increases outward, and reaches a maximum amplitude near the surface; in contrast, the eigenfunction of the m = 3 mode (seen in J127) has a maximum between the center and the surface. We would like to thank Hui Li for insightful and useful discussions at the 47th Meeting of the Division of Plasma Physics of the American Physical Society. We also thank Juhan Frank, Patrick Motl, Shin Yoshida, and Christian Ott for helpful comments and suggestions. It is a pleasure to thank Anna Watts for useful conversations at the Eighth Divisional Meeting of High Energy Astrophysics Division of the American Astronomical Society. We thank an anonymous referee for helpful suggestions. This work was partially supported by NSF grants AST-0407070 and PHY-0326311. The computations were performed on the Supermike cluster and the Superhelix cluster at LSU, and on the Tungsten cluster at National Center for Supercomputing Applications (NCSA). REFERENCES Akiyama, S., & Wheeler, J. C., 2005, ApJ, 629, 414 Andersson, N. 2003, Class. Quant. Grav., 20, 105 Brown, J. D. 2000, Phys. Rev. D, 62, 084024 Cazes, J. E., & Tohline, J. E. 2000, ApJ, 532, 1051 Centrella, J. M., New, K.C.B., Lowe, L., Brown, J.D. 2001, ApJ, 550, L193 Chandrasekhar, S. 1969, Equilibrium Figures of Equilibrium, New Haven, CT: Yale Univ. Press Chandrasekhar, S. 1970, ApJ, 161, 561 Drazin, P. G., & Reid, W. H., Hydrodynamic Stability (Cambridge: Cambridge Univ. Press) -- 16 -- Durisen, R. H., Gingold, R. A., Tohline, J. E., & Boss, A. P. 1986, ApJ, 305, 281 Frank, J., & Robertson, J.A., 1988, MNRAS, 232, 1 Friedman, J., & Schutz, B. F. 1978, ApJ, 222, 281 Goldreich, P., Goodman, J., & Narayan, R., 1986, MNRAS, 221, 339 Hachisu, I. 1986, ApJS, 61, 479 Ipser, J. R., & Lindblom, L. 1990, ApJ, 355, 226 Ipser, J. R., & Lindblom, L. 1991, ApJ, 373, 213 Lai, D., & Shapiro, S. L. 1995, ApJ, 442, 259 Li, H., Finn, J. M., Lovelace, R. V. E., & Colgate, S. A., 2000, ApJ, 533, 1023 Li, H., Colgate, S. A., Wendroff, B., Liska, R., 2001, ApJ, 551, 874 Liu, Y. T., 2002, Phys. Rev. D, 65, 124003 Lovelace, R. V. E., Hohlfeld, R. G., 1978, ApJ, 221, 51 Lovelace, R. V. E., Li, H., Colgate, S. A., Nelson, A. F., 1999, ApJ, 513, 805 Motl, P. M., Tohline, J. E., & Frank, J. 2002, ApJS, 138, 121 Narayan, R., Goldreich, P., & Goodman, J., 1987, MNRAS, 228, 1 New, K. C. B., Centrella, J. M., & Tohline, J. E. 2000, Phys. Rev. D, 62, 064019 Ott, C. D., Burrows, A., Livne, E., & Walder, R. 2004, ApJ, 600, 834 Ott, C.D., Ou, S., Tohline, J.E., Burrows, A., 2005a, ApJ, 625, L119 Ott, C. D., Burrows, A., Thompson, T. A., Livne, E., & Walder, R. 2005b, astro-ph/0508462 Ou, S., ,Tohline, J.E, Lindblom, L. 2004, ApJ, 617, 490 Papaloizou, J.C.B., Pringle, J.E., 1984, MNRAS, 208, 721 Papaloizou, J.C.B., Lin, D. N. C., 1989, ApJ, 344, 645 Saijo, M., Baumgarte, T. W., & Shapiro, S. L. 2003, ApJ, 595, 352 Saijo, M., & Yoshida, S., 2005, astro-ph/0505543 -- 17 -- Shibata, M., & Karino, S. 2004, Phys. Rev. D, in press (astro-ph/0408016) Shibata, M., Karino, S., & Eriguchi, Y. 2002, MNRAS, 334,L27 Shibata, M., Karino, S., & Eriguchi, Y. 2003, MNRAS, 343, 619 Tassoul, J.-L. 1978, Theory of Rotating Stars, Princeton: Princeton University Press Tohline, J. E., Durisen, R. H., & McCollough, M. 1985, ApJ, 298, 220 Tohline, J. E., Hachisu, I., 1990, ApJ, 361, 394 Watts, A. L., Andersson, N., & Jones, D. I. 2005, ApJ, 618, L37 Williams, H. A., & Tohline, J. E. 1988, ApJ, 334, 449 Woodward, J. W., Tohline, J. E., & Hachisu, I., 1988 ApJ, 420, 247 This preprint was prepared with the AAS LATEX macros v5.2. -- 18 -- Fig. 1. -- The top panel shows m = 1, 2, and 3 Fourier amplitudes as a function of time t in the equatorial plane of model J250 when GRR was turned on; t is normalized to the central spin period Tc of the differentially rotating star. The bottom panels show equatorial-plane isodensity contours at two different evolutionary times, with inertial-frame velocity vectors superposed; contour levels are (from the innermost, outward) ρ/ρmax = 0.9, 0.5, 0.1, and 0.01. -- 19 -- Fig. 2. -- Same as Figure 1, but for model J250 when GRR forces are turned off. -- 20 -- Fig. 3. -- The top and bottom panels show, respectively, δρ(R) and ρ(R) for model J250 in the equatorial plane when the m = 1 mode is dominant. -- 21 -- Fig. 4. -- The rotational profile Ω(R) of model J250 is shown at three different evolutionary times. The dash-dotted horizontal line identifies the measured eigenfrequency ω1 of the unstable m=1 mode. -- 22 -- Fig. 5. -- Same as Fig. 1, but for model J133 when GRR forces are turned off. -- 23 -- Fig. 6. -- The same as Figure 1, but for model J127 with GRR turned off. -- 24 -- Fig. 7. -- The angular velocity profile Ω(R) is shown for model J127 at three different evolutionary times. Dash-dotted horizontal lines denote the measured eigenfrequencies ω1, ω2 and ω3 of various unstable modes. ω2/2 is measured at ∼ 50 Tc, ω3/3 is measured at ∼ 80 Tc, and ω1 is measured at ∼ 400 Tc. To determine a corotation radius, we used a rotational profile that is around at the same time when each frequency is measured. -- 25 -- Fig. 8. -- The quadrupole gravitational-wave strain that would be radiated from model J127 as viewed by an observer looking down the Z (rotation) axis. -- 26 -- Fig. 9. -- The pattern frequencies of different modes is displayed as a function of time for model J127. The frequency ω2/2 is measured between 40 and 130 Tc, ω3/3 is measured between 50 and 430 Tc, and ω1 is measured between 220 and 430 Tc. -- 27 -- Fig. 10. -- Same as Figure 1, but for model J068 when GRR is turned off. -- 28 -- Fig. 11. -- The angular velocity profile Ω(R) of model J068 is plotted at two different times during its evolution. Dashed horizontal lines identify measured eigenfrequencies of the unstable m = 1 and m = 2 modes. -- 29 -- Fig. 12. -- The radial vortensity profile is shown for model J068 initially (solid line) and at a time t = 53Tc (dotted line) during its evolution. The horizontal axis is the cylindrical radius, the vertical axis is the vortensity normalized to the value on the Z axis. The two vertical dashed lines locate the corotation radii for m = 1 (left) and m = 2 (right) modes. There seems to be a correlation between the growth rate and the location of corotation radii inside the vortensity well: the m = 2 mode with a "deeper" corotation radius in the initial vortensity well was the first unstable mode to arise in the simulation, whereas the m = 1 mode whose corotation radius lies at a "shallower" location in the initial vortensity well developed later in the evolution. -- 30 -- Fig. 13. -- The radial vortensity profile is shown for model J127 initially (solid line) and at a time t = 240Tc (dotted line) during its evolution. The horizontal axis is the cylindrical radius, the vertical axis is the vortensity normalized to the value on the Z axis. The three vertical dashed lines locate the corotation radii for m = 1 (left), m = 2 (middle), and m = 3 (right) modes. -- 31 -- Fig. 14. -- The dotted curve displays the azimuthal structure φ1(R) of the m = 1 eigenfunc- tion in the equatorial plane of model J250 at t = 95.4Tc when GRR was turned on. The solid circle denotes the corotation radius; the dashed circle denotes the location of density maximum. -- 32 -- Fig. 15. -- Same as Figure 14 but for the m = 2 mode of model J068 when t = 70Tc; there is no dashed circle because the density maximum lies at the center of the star.
0709.2019
1
0709
2007-09-13T08:32:49
Drift Wave Model of Rotating Radio Transients
[ "astro-ph" ]
During the last few years there were discovered and deeply examined several transient neutron stars (Rotating Radio Transients). It is already well accepted that these objects are rotating neutron stars. But their extraordinary features (burst-like behavior) made necessary revision of well accepted models of pulsar interior structure. Nowadays most popular model for RRATs is precessing pulsar model, which is the subject of big discussion. We assume that these objects are pulsars with specific spin parameters. An important feature of our model, naturally explaining most of the properties of these neutron stars, is presence of very low frequency, nearly transverse drift waves propagating across the magnetic field and encircling the open field lines region of the pulsar magnetosphere.
astro-ph
astro-ph
Drift Wave Model of Rotating Radio Transients D.Lomiashvili a,∗ G.Machabeli b I.Malov c aPurdue University, 525 Northwestern Avenue, West Lafayette, IN, USA bTbilisi state university, 3 Chavchavadze Ave., 0128, Tbilisi, Georgia cPushchino Radio Astronomy Observatory, P.N.Lebedev Institute of Physics, 142290, Pushchino, Moscow Region, Russia Abstract During the last few years there were discovered and deeply examined several tran- sient neutron stars (Rotating Radio Transients). It is already well accepted that these objects are rotating neutron stars. But their extraordinary features (burst-like behavior) made necessary revision of well accepted models of pulsar interior struc- ture. Nowadays most popular model for RRATs is precessing pulsar model, which is the subject of big discussion. We assume that these objects are pulsars with spe- cific spin parameters. An important feature of our model, naturally explaining most of the properties of these neutron stars, is presence of very low frequency, nearly transverse drift waves propagating across the magnetic field and encircling the open field lines region of the pulsar magnetosphere. Key words: (stars:) pulsars: individual PSR J1819-1458, PSR J1752+2359, PSRs J1649+2533, stars: magnetic fields, radiation mechanisms: non-thermal PACS: 97.60.Gb, 94.20.Bb 1 Introduction Recently, McLaughlin et al. (2005) reported the discovery of a new class of ra- dio transients from the Parkes Multibeam Pulsar Survey. The current sample includes 11 objects characterized by single, dispersed bursts of radio emission with the durations ranging from 2 to 30 milliseconds. Long-term monitoring ∗ Corresponding author. Email addresses: [email protected] (D.Lomiashvili), g [email protected] (G.Machabeli), [email protected] (I.Malov). Preprint submitted to New Astronomy 30 October 2018 of these objects led to identification of their spin periods (P), ranging from 0.4 to 7 seconds. As McLaughlin et al. (2005) concluded, these objects rep- resent a previously unknown population of rotation-powered neutron stars, which were named as Rotating RAdio Transients (RRATs). Another interest- ing case between normal pulsars and RRATs have been reported earlier (PSRs J1649+2533 and J1752+2359) by Lewandowski et al. (2004). Understanding the physical origin of these objects as well as their relationship with normal pulsars is desirable. There exist few models for this phenomenon: 1) The pre- cession model; 2) The model suggesting that these objects are pulsars located slightly below the radio emission death line, and become active occasionally when the conditions for pair production and coherent emission are satisfied; 3) The third model invoking a radio emission direction reversal in normal pul- sars. In this picture, our line of sight misses the main radio emission beam of RRATs but happens to sweep the emission beam when the radio emission direction is reversed. Last two ideas were suggested by Zhang et al. (2006). Here we propose another model of RRATs proving once again that they are normal pulsars with special values of certain parameters. The paper is organized as follows. Pulsar radio emission mechanism is pre- sented in Section 2. The generation of the drift waves and their influence on the curvature of magnetic field lines are discussed in Section 3. Our proposed model is presented in Section 4. The conclusions are summarized in Section 5. 2 Emission mechanism As it is known the pulsar magnetosphere is filled by a dense relativistic electron-positron plasma. The (e+e−) pairs are generated as a consequence of the avalanche process (first described by Sturrock (1971)) and flow along the open magnetic field lines. The plasma is multi-component, with a one- dimensional distribution function ( see Fig.1 from Arons (1981)) and consists of the following components: the bulk of plasma with an average Lorentz-factor γp ≃ 10; a tail on the distribution function with γt ≃ 104 and the primary beam with γb ≃ 106. The main mechanism of wave generation in plasmas of the pulsar magnetosphere is the cyclotron instability. Generation of waves is possible if the condition of the cyclotron resonance if fulfilled (Kazbegi et al., 1991a): ω − kϕVϕ − kxux + ωB γr = 0, (1) where Vϕ is the particle velocity along the magnetic field, γr is the Lorentz- factor for the resonant particles and ux = cVϕγr/ρωB is the drift velocity of the particles due to curvature of the field lines (ρ is the radius of curvature of 2 the field lines and ωB = eB/mc is the cyclotron frequency). Here cylindrical coordinate system is chosen, with the x-axis directed transversely to the plane of field line, when r and ϕ are the radial and azimuthal coordinates. Generated waves leave the magnetosphere propagating at very small angles to the pulsar local magnetic field lines and reach an observer as pulsar radio emission. These processes take place near the light cylinder where the cyclotron instability occurs (Lyutikov et al. , 1999). 3 Change of field line curvature and emission direction by drift waves It has been shown by Kazbegi et al. (1991b, 1996) that, in addition to the radio waves, very low frequency, nearly transverse drift waves can be excited in the same region. The period of the drift waves Pdr can be written as: Pdr = e BP 2 4π2mc γ (2) Where P is the pulsar spin period, B = Bs(R0/R)3 is the magnetic field in the wave excitement region and γ is the relativistic Lorentz factor of the particles. It appears that the period of the drift wave can vary in a broad range. The magnetic field of drift wave adds with pulsar magnetic field as r component and causes changing of field line curvature ρc. Here and below the cylindrical coordinate system (x, r, ϕ) is chosen, with the x-axis directed transversely to the plane of field line, while r and ϕ are the radial and azimuthal coordinates, respectively. Even a small change of Br causes significant change of ρc. Variation of the field line curvature can be estimated as ∆ρ ρ ≈ kϕr ∆Br Bϕ (3) Here kϕ is a longitudinal component of wave vector and r is distance to the center of pulsar. It follows that even the drift wave with a modest amplitude Br ∼ ∆Br ∼ 0.01Bϕ alters the field line curvature substantially, ∆ρ/ρ ∼ 0.1 Since radio waves propagates along the local magnetic field lines, curvature variation causes change of emission direction. 3 Fig. 1. Geometry of Ω - rotation, K - emission and A - observers axes. Angles δ and ϑ are constants, while β and α are oscillating with time. 4 The model There is unequivocal correspondence between the observable intensity and α (angle between observers line and emission direction (see fig. 1)). Maximum of intensity corresponds to minimum of α. The period of pulsar is time interval between neighboring maxima of observable intensity i.e. minima of α (see fig. 2). According to this fact, we can say that the observable period depends on time behavior of α and as it will appear below it might differ from the spin period of pulsar. cos α = A · K (4) A and K are unit guide vectors of observers and emission axes respectively. In the spherical coordinate system (r, ϕ, θ), combined with plane of pulsar rotation, these vectors can be expressed as: A = (1, 0, δ) K = (1, Ωt, β) 4 (5) (6) where Ω = 2π/P is the angular velocity of the pulsar. δ is the angle between rotation and observers axis, and β is the angle between rotation and emission axis (see fig. 1). From equations (4),(5) and (6) follows that: α = arccos (sin δ sin β cos Ωt + cos δ cos β) (7) In the absence of the drift wave β = β0 = const and consequently the period of α equals to 2π/Ω, while in the case of existence of drift wave β is oscillating with time (Lomiashvili et al. , 2006). Considering that for wavelengths of order of the transverse scale of magneto- sphere, the treatment of the drift waves as plane waves becomes inappropriate the geometry of magnetosphere must be taken into account. A treatment in terms of spherical harmonic eigenmodes with specific l, m is then appropriate (Gogoberidze et al., 2005). Therefore in general case the amplitude ∆β must be defined as follows ∆β = X ∆βn sin (ωn drt + ϕn) (8) Here n is number of eigenmode. According to equations (7) and (8) we obtain α = arccos[sin δ sin(cid:16)β0 + X ∆βn sin (ωn + cos δ cos(cid:16)β0 + X ∆βn sin (ωn drt + ϕn)(cid:17) cos Ωt drt + ϕn)(cid:17)] (β0 − δ) + ∆β sin(cid:18)2πk αk min = (cid:12)(cid:12)(cid:12)(cid:12) ωdr Ω + ϕ(cid:19)(cid:12)(cid:12)(cid:12)(cid:12) (9) (10) Where αk min is the minimum of α after k revolutions of the pulsar, after time reckoning 1 . The parameters of the pulse profile (e.g. width, height etc.) sig- nificantly depend on what the minimal angle would be between emission axis and observers axis when the first one passes the other (during one revolution). If the emission cone does not cross the observers line of sight entirely (i.e. minimal angle between them is more then cone angle ϑ, see inequality (11)) then pulsar emission is unobservable for us. On the other hand, inequality (12) defines condition that is necessary for emission detection. αk min > ϑ αk min < ϑ (11) (12) 1 as zero point of time reckoning is taken detection moment of any pulse. 5 Fig. 2. The oscillating behavior of α with time for β0 = 1.35, δ = 1.0, ∆β = 0.35, ωdr = 2π(1/4.26 + 1/204.4)(s−1), Ω = 2π/4.26(s−1), ϕ = 1.5 and Simulated lightcurve of PSR J1819-1458 (below) Hence for some values of parameters Ω, ωdr, β, ∆β, δ, ϕ and ϑ (Set A) it is possible to accomplish following regime: Once condition (12) becomes fulfilled it stays along k revolutions, therefore pulsar is observable during k periods. While over the the other m − k revolutions condition (11) accomplishes, con- sequently pulsar appears "switched off" (see Fig. 2). It means that burst-like emission occurs. PSR J1752+2359 has been selected for its unusually long nulling periods. It had been observed on several occasions between 2000 and 2002 by Lewandowski et al. 6 Fig. 3. Simulated lightcurve of PSR J1752+2359 (2004). The pulsar spends 70- 80% of the time in a quasi-null state. The on- states occur once every 400-600 periods and last, on average, for ∼ 100 peri- ods. A more detailed inspection of the on-states reveals that these burst-like emissions are quite similar in shape and duration and they decay into a null state in a manner that is quite reasonably described by a te−t/τ function with τ ∼ 45 seconds. PSR J1752+2359 switches off gradually, rather than suddenly. Simulated lightcurve of J1752+2359 is presented in Fig. 3. If we consider these pulsars in the framework of our model their parameters (spin, angular etc.) will get values shown in Table 1. Simulated lightcurve of PSR J1819-1458 (RRAT J1819-1458) is presented in Fig. 2. 5 Conclusions To sum up, drift wave driven model is very convenient since it allows to explain almost every extraordinary feature of known pulsars such as: ex- tremely long periods (Lomiashvili et al. , 2006), cyclic variations of radia- tion intensity (RRATs) and rotational parameters (in prep.), subpulse drift (Gogoberidze et al., 2005), nullings (Kazbegi et al., 1996), etc. Moreover it has potentiality to unfold future discoveries in this field. In long period radio pulsars (Lomiashvili et al. , 2006) relation between peri- ods of pulsar and drift wave along with pulsar geometry satisfy much precise 7 Table 1 The values of pulsar parameters PSR P ∆β J1752+2359 0.41 0.2 β0 1.2 J1819-1458 4.26 0.35 1.35 δ 1 1 ϑ 0.03 0.03 condition than in RRATs. Therefore total number of RRATs should be much more than long period radio pulsars. The precession model is very similar to ours although it has number of dis- advantages. First of all, it is necessary to take large value of wobble angle to interpret observational data of some RRATs. In the case of neutron stars this is hard to explain because of their superfluid interior structure. However, there is a need of distinguishable sign to elucidate which model is true. Zhang et al. (2006) reported that one of the such tests would be search of X-ray counter- part from RRATs. Particularly, presence of a hot-spot thermal component in the possible X-ray spectra will mean that the emission direction reversal and a preferred viewing geometry are likely the agents to make a RRAT. While ab- sence of X-ray radiation indicates that considering pulsars are "not-quite-dead pulsars before disappearing in the graveyard." We suggest to investigate the power spectrum of the single-pulse observational data thoroughly. In our opinion, the power spectrum should be composition of few slightly broadened areas (ranges with peaks) instead of thin lines. Find- ing this kind of signature in the power spectra of RRATs will help to asses wether the origin of the transient nature of these objects is indeed plasmic, as suggested by our model. Acknowledgments This work was partially supported by by Georgian NSF Grant ST06/4-096, Russian Foundation for the Basic Research (project 06-02-16888) and NSF (project 00-98685). References Arons J., 1981a, Plasma Astrophys., 273 Gogoberidze, G., Machabeli, G.Z., Melrose, D.B., & Luo, Q. 2005, MNRAS, 360, 669 Kazbegi A. Z., Machabeli G. Z., Melikidze G. I., Smirnova T. V., 1991a, Afz, 34, 433 8 Kazbegi A. Z., Machabeli G. Z., Melikidze G. I., 1991b, MNRAS, 253, 377 Kazbegi A. Z., Machabeli G. Z., Melikidze G. I., Shukre C., 1996, A&A, 309, 515 Lewandowski W., Wolszczan A., Feiler G., Konacki M., & Soltysinski T., 2004, ApJ, 600, 905 Lomiashvili D., Machabeli G., & Malov I., 2006, ApJ, 637, 1010 Lyutikov M, Blandford R., Machabeli G., 1999, MNRAS, 305, 338L McLaughlin M. A., Lyne A. G., Lorimer D. R., Kramer M., Faulkner A. J., Manchester R. N., Cordes J. M., Camilo F., Possenti A., Stairs I. H., Hobbs G., DAmico N., Burgay M., & OBrien J. T., 2005, (astro-ph/0511587) Sturrock P. A., 1971, ApJ, 164, 529 Zhang B., Gil J., & Dyks J., 2006, (astro-ph/0601063) 9
0704.1114
1
0704
2007-04-09T20:13:36
Identification of Absorption Features in an Extrasolar Planet Atmosphere
[ "astro-ph" ]
Water absorption is identified in the atmosphere of HD209458b by comparing models for the planet's transmitted spectrum to recent, multi-wavelength, eclipse-depth measurements (from 0.3 to 1 microns) published by Knutson et al. (2007). A cloud-free model which includes solar abundances, rainout of condensates, and photoionization of sodium and potassium is in good agreement with the entire set of eclipse-depth measurements from the ultraviolet to near-infrared. Constraints are placed on condensate removal by gravitational settling, the bulk metallicity, and the redistribution of absorbed stellar flux. Comparisons are also made to the Charbonneau et al. (2002) sodium measurements.
astro-ph
astro-ph
Draft version June 15, 2018 Preprint typeset using LATEX style emulateapj v. 7/15/03 7 0 0 2 r p A 9 ] h p - o r t s a [ 1 v 4 1 1 1 . 4 0 7 0 : v i X r a IDENTIFICATION OF ABSORPTION FEATURES IN AN EXTRASOLAR PLANET ATMOSPHERE Lowell Observatory, 1400 W. Mars Hill Rd., Flagstaff, AZ 86001, [email protected] T. Barman Draft version June 15, 2018 ABSTRACT Water absorption is identified in the atmosphere of HD209458b by comparing models for the planet's transmitted spectrum to recent, multi-wavelength, eclipse-depth measurements (from 0.3 to 1 µm) published by Knutson et al. (2007). A cloud-free model which includes solar abundances, rainout of condensates, and photoionization of sodium and potassium is in good agreement with the entire set of eclipse-depth measurements from the ultraviolet to near-infrared. Constraints are placed on condensate removal by gravitational settling, the bulk metallicity, and the redistribution of absorbed stellar flux. Comparisons are also made to the Charbonneau et al. (2002) sodium measurements. Subject headings: planetary atmospheres - extrasolar planets 1. INTRODUCTION The discovery of transiting extrasolar planets has opened the door to direct detections and characteriza- tion of their atmospheres. Observations using the STIS instrument on HST provided the first glimpse of what the photospheric composition is like for a nearby EGP. Charbonneau et al. (2002) measured the relative change in eclipse depth for HD209458b across a sodium doublet (5893A) resulting in the first detection of atomic absorp- tion in an EGP atmosphere. Following the Na detec- tion, Vidal-Madjar et al. (2003) discovered an extended hydrogen-rich atmosphere surrounding HD209458b using a similar technique as Charbonneau et al., but in the UV. At Lyman-α wavelengths, HD209458b is ∼ 3 times larger than in the optical. These detections were made using a technique called transit spectroscopy as the planet passes in front of its star. Transit spectroscopy uses the fact that the wavelength-dependent opacities in the planet's atmo- sphere obscure stellar light at different planet radii lead- ing to a wavelength-dependent depth of the light-curve during primary eclipse. Consequently, searching for rela- tive changes in eclipse depth as a function of wavelength directly probes the absorption properties of the planet's atmosphere with the potential to reveal the presence (or absence) of specific chemical species. In this paper, recent measurements of HD209458b's radius are combined in a multi-wavelength comparison to model atmosphere predictions. Atmospheric molecular and atomic absorption are identified and constraints are placed on the basic atmospheric properties. 2. THE LIMB MODEL Transit spectroscopy probes the limb of a planet which is the transition region between the day and night sides. One would, therefore, expect that, in the presence of a horizontal temperature gradient between the heated and non-heated hemispheres, the temperatures across the limb would be cooler than the average dayside tem- peratures (Barman et al. 2005; Iro et al. 2005). Re- cent Spitzer observations of both transiting and non- transiting hot-Jupiters showing large flux variations with phase have provided strong evidence supporting such a day-to-night temperature gradient (Charbonneau et al. 2005; Deming et al. 2005b, 2006; Harrington et al. 2006). Describing the limb ultimately requires a multi- dimensional model atmosphere solution; however, as is common practice, a simpler one-dimensional model is used here to represent the average properties of the limb, in both a longitudinal and latitudinal sense. To explore a variety of limb temperature structures, the incident stellar flux has been scaled by a parameter α. A model with α = 0.25 represents an average description of the entire planet in the presence of very efficient horizontal energy transport (Barman et al. 2005). Increasing α to 0.5 increases the heating of the model atmosphere mak- ing it more appropriate for just the dayside which, in this work, will represent an upper limit to the plausible mean temperatures at the limb. The basic chemistry across the limb is modeled assum- ing chemical equilibrium, determined by minimizing the free energy while including grain formation. To under- stand the effects of gravitational settling of grains on the chemistry and the transmission spectrum, the removal of grains from the atmosphere is included via two simple approximations that represent opposite extremes. The first is the "cond" approximation used by Barman et al. (2001) and Allard et al. (2001) which simply ignores the grain opacity without altering the chemistry or abun- dances. The second is the "rainout" approximation which iteratively reduces, at each layer, the elemental abundances (by the appropriate stoichiometric ratios) involved in grain formation and recomputes the chem- ical equilibrium with each new set of stratified elemen- tal abundances. This approach is similar to other rain- out models (Fegley & Lodders 1996; Burrows & Sharp 1999), except that the depletion of elements is contin- ued until grains are no longer present. The transmission of stellar fluxes through the limb of HD209458b is determined by solving the spherically sym- metric radiative transfer equation while fully accounting for scattering and absorption of both intrinsic and ex- trinsic radiation (Hauschildt 1992; Barman et al. 2001, 2002). Spherical, instead of the more traditional plane- parallel geometry, naturally accounts for the curvature of the atmospheric layers and changes in chemistry along the slant paths through the upper atmosphere. The planet radius at a given wavelength (Rλ) is obtained by determining the radial depth at which the transmitted 2 Fig. 1.- Monochromatic transit radii over the STIS spectral range for the baseline rainout model with (red) and without (blue) photoionization of Na and K. The solid and dotted red lines are the same, except H2O line opacity is excluded in the latter. Horizontal bars correspond to mean radii across bins with λ-ranges indicated by the width of each bar. STIS measurements by Knutson et al. (2007) are shown in green with 1 σ error bars. Vertical dashed lines mark the narrow λ-range used by Charbonneau et al. (2002). flux is equal to e−1 times the incident starlight along that same path. 3. RESULTS A cloud-free atmosphere with rainout, α = 0.25, and solar abundances is adopted as the baseline model. This model, along with others, is compared to the relative Rλ measurements of Knutson et al. (2007) which have a reported precision high enough to constrain many of the basic atmospheric properties. Since a comparison is being made to relative Rλ values, the model results were uniformly scaled to match the observations in the 4580 to 5120A wavelength bin; this scaling was always less than 0.005RJ up. Overall, the baseline model (red solid line in Fig. 1) reproduces the observed rise in Rλ toward shorter wavelengths, the increase across the Na doublet, and the increase at the far red wavelengths. The baseline model comparison to the data has a χ2 that is 3 times smaller than a constant Rλ. 3.1. Water Absorption Water is predicted to be one of the most abundant species in an EGP atmosphere and, given its broad ab- sorption features in the infrared, plays a crucial role in regulating the temperature-pressure (T-P) profile. The first major H2O absorption band appears between 0.8 and 1 µm, a region covered by the last two wavelength bins of Knutson et al. (2007). As illustrated in Fig. 1, there is excellent agreement between the baseline model and the observations in this part of the spectrum espe- cially across the longest wavelength bin that sits on top of the H2O band. Qualitatively similar water features are seen in the models of Brown (2001) and Hubbard et al. (2001); however these models fall many σ below the ob- servations. The baseline model also predicts mean Rλ- peaks equal to 1.330, 1.343, and 1.341 RJ up for the next three water bands (at λ ∼ 1.15, 1.4, and 1.9 µm). A model that excludes H2O line opacity is also shown in Fig. 1 and is greater than 10 σ below both the observa- tions and the baseline model prediction. Removing H2O opacity also produces a significant drop in Rλ near 0.9 µm that further increases the discrepancy between the model and overall red/near-IR observations. No other opacity source could be responsible for the observed rise in Rλ across this part of the spectrum. 3.2. Photoionization After the reported sodium detection in the atmosphere of HD209458b (Charbonneau et al. 2002), there were several attempts to explain why the strength of this fea- ture was much lower than expected based on earlier mod- els (e.g., by Seager & Sasselov, 2000). Barman et al. (2002) explored departures from local thermodynamic equilibrium (LTE) which are capable of producing an in- version in the cores of the Na doublet line profiles. How- ever, the earlier models of Barman et al. (2001, 2002) were constructed under the cond approximation and, consequently, contained a larger number of free met- als along with TiO and VO molecular absorption com- pared to a rainout model. These additional sources of optical/UV opacity lead to a hotter upper atmosphere and a very shallow photoionization depth for Na (only a 5% reduction of Na across the region probed by tran- sit spectroscopy). Fortney et al. (2003) also explored Na photoionization and found that their atmosphere model (which included rainout) could be brought into reason- able agreement with the Na observations. The models presented here account for the angular dependence of ionziation on the limb's dayside, but not on the night side. However, Fortney et al. (2003) have shown that Na ionization is still present at ∼ 5◦ past the terminator for pressures relevant to the transit spectrum modeled here. Figure 1 compares cloud-free solar abundance rainout models with and without photoionization of Na and K. While these models include photoionization, the LTE ap- proximation on the atomic level populations and line source function is maintained. Ionization (radially) at the limb reaches 50% at P ∼ 0.1 and 1.2 mbar for Na and K, respectively, and stops at P ∼ 2 and 7 mbar, re- spectively, resulting in a reduction of the Rλ-peaks across 3 The rainout model used here reduces the individual metal abundances with depth iteratively until clouds do not form, thus mimicking efficient gravitational settling. However, the removal of metals is not 100% efficient, leaving behind a variety of atoms (in addition to Na and K) to contribute to the line opacity. The impact of rain- out on the planet's atmosphere is made apparent by com- paring rainout and cond models (Fig. 3). In the cond model grain opacities are simply ignored in the transfer equation while the chemistry remains that of a pure equi- librium model without any actual removal of refractory elements. Thus, the cond model represents the minimum impact of grain formation on the stratified abundances and leaves behind a considerable amount of free metals along with molecules like TiO and VO. This leads to much stronger atomic lines, including Na and K along with TiO and VO bands. Apart from a very minor contribution from molecules, the Rλ features at λ < 0.8µm are all due to atomic line opacity 1. These narrow features are due to blended lines from metals like Ca, Al, Fe, Ni, Mn, and Cr. Note that, while Rayleigh scattering does contribute to the opacity in the blue/UV, a removal of atomic line opacity would drop Rλ by ∼ 0.02RJ up for λ < 0.45µm. The top panel of Fig. 3 compares the solar abundance rainout model with photoionization (shown in Fig. 1) to a rainout model with 10× solar abundances. Since the metal abundances were uniformly scaled, the grain formation and corre- sponding removal of refractory elements were also uni- formly enhanced. The additional opacity altered the T-P profile as well as the total atmospheric extension. These factors contribute to a similar Rλ pattern in the opti- cal but enhanced molecular absorption features in the red/near-IR. The increase in metals increases the H2O absorption feature along with the wings of the Na and K doublets (which form deeper in the atmosphere where photoionization is less important). Also, FeH and CrH absorption is larger in the metal-rich atmosphere leading to an increase in Rλ on both sides of the K doublet. The metal-enhanced Rλ spectrum is in good agreement with the observations near 0.9µm, but is noticeably too high across both the K doublet and the H2O band. Transit spectroscopy can also help constrain the tem- peratures across the limb. The lower panel of Fig. 3 compares an α = 0.5 model (i.e. a model with a hotter dayside-like T-P profile) to the cooler α = 0.25 base- line model. In the hotter model, grain formation is less pronounced resulting in a greater concentration of most free metals and, thus, stronger UV/blue absorption lines than in the α = 0.25 model. Higher temperatures at depth also result in equilibrium concentrations of Na and K that are several times lower than found at the cooler temperatures of the α = 0.25 model. The net re- sults are two distinctively different Rλ spectra with the α = 0.25 model being more consistent with the Knutson et al. measurements. 4. SUMMARY 1 Ballester et al. (2007) claim the Rλ rise in the blue/UV is due to the Blamer edge of hot hydrogen in an extended atmosphere. While the present model does not support this explanation, if ver- ified, it could indicate that rainout is more efficient than predicted here. Fig. 2.- Relative flux differences (with time from center of eclipse) between a wavelength band centered on the Na doublet and the mean of two wavelength bands on either side of the Na doublet. The total wavelength range is indicated by vertical lines in Fig. 1. Each model is cloud-free with solar abundances and α = 0.25. See Figure legend and text for the distinguishing characteristics of each model. Symbols are the observations of Charbonneau et al. (2002) with 1 σ error bars. the Na and K lines. For K, this reduction extends out to the line-wings resulting in a significantly smaller mean Rλ, bringing the model into ∼ 3σ agreement with the Knutson et al. measurement. The impact of Na pho- toionization is mostly confined to the core of the doublet leading to only a small reduction of the mean Rλ, but suf- ficient to bring the model into ∼ 1σ agreement with the observations at these wavelengths. Though not included here, ionization past the terminator onto the night side should further improve the agreement across the Na and K doublets. The two narrow features on the red wing of the K doublet are due to Rb, which should also be affected by photoionization due to its very low first ion- ization potential. Photoionization of Rb resulted in a near complete removal of these features, but reduced the mean Rλ for this bin by less than 0.002 RJ up. The broad wavelength bins allowed Knutson et al. to obtain very precise relative Rλ measurements; however, as illustrated by the Na doublet, such broad bins lim- its the constraints that can be placed on atomic absorp- tion features. In contrast, the narrow range analyzed by Charbonneau et al. (2002) resulted in much larger Rλ error-bars, but still precise enough to easily distin- guish between the various models shown in Fig. 1. The flux differences across the Na doublet as a function of time (measured from the transit center) were computed for each model using the same narrow wavelength bins as Charbonneau et al. (2002), indicated in Fig. 1 by vertical dashed lines. Fig. 2 compares model transit curves to the 2002 Na measurements and shows a large discrepancy between the observations and solar abun- dance models with pure equilibrium chemistry (cond or rainout). Including photoionization brings the baseline rainout model into rough agreement with the observa- tions. Note the cond model shown in Fig. 2 does not include photoionization (similar to the LTE model from Barman et al. (2002)) and illustrates how far off the pre- dicted Rλ can be across the Na doublet under simplified assumptions. 3.3. Rainout, Metallicity, and Temperature 4 Fig. 3.- Monochromatic transit radii over the STIS spectral range. Top panel: solar abundance models, cond (black) and rainout(red), are compared to a model with 10× solar metal abundances including rainout (blue). Lower panel: solar abundances baseline rainout model with α = 0.25 (red) compared to a similar model with α = 0.5 (blue). Horizontal bars have the same meaning as in Fig. 1. Photoionziation plays an important role for both Na and K, and potentially many other species. No evidence is found to support a large metal enhancement (e.g., 10× solar), though smaller Jupiter-like enhancements are not ruled out. Furthermore, the agreement between model and observations demonstrated here alleviates the need for substantial cloud coverage along the limb between 0.05 and 0.001 bars, which would otherwise truncate and flatten the Rλ spectrum (Brown 2001). In the baseline model, the predicted location of clouds (e.g., MgSiO3) lies just below the minimum Rλ(∼ 1.315RJ up) across the STIS wavelength range, consequently deep clouds (and also very high clouds) remain a possibility. The models presented here also predict Rλ variations (∼ 0.02 RJ up) across the 2-0 R-branch of CO in the K- band. This is inconsistent (at 2.5σ) with a non-detection of CO based on Keck-NIRSPEC transit spectra taken in 2002 (Deming et al. 2005a). It is likely that using a single T-P profile to represent the horizontal and pole- to-pole variations across the limb averages out too much of the upper atmospheric structure; this simplification might explain the K-band discrepancy. High clouds may also be involved. In addition, the Keck and HST obser- vations were taken between 1 and 2 years apart and time- dependent atmospheric variations along the limb cannot be ruled out (Menou et al. 2003). While Knutson et al. (2007) did not attribute their measured Rλ variations to any absorption features (this was not the focus of their paper), the models presented above clearly show that these measurements are consis- tent with strong water absorption near 1 µm. A detection of water in the limb of HD209458b is nominally at odds with a recent Spitzer IRS spectrum that shows no H2O features for this planet(Richardson et al. 2007). These data were taken during secondary eclipse and directly probe the planet's dayside with negligible contribution from the limb. It is possible that the dayside atmosphere is nearly isothermal (Fortney et al. 2006) which would re- sult in a spectrum with no detectable water absorption features, despite a copious water supply. The transmis- sion spectrum, however, would contain absorption fea- tures independent of an isothermal dayside or limb. The author thanks B. Hansen and the referees for their comments. This research was supported by NASA Ori- gins and Spitzer Theory Grants and made use of NASA's Project Columbia computer system. REFERENCES Allard, F., Hauschildt, P. H., Alexander, D. R., Tamanai, A., & Deming, D., Harrington, J., Seager, S., & Richardson, L. J. 2006, Schweitzer, A. 2001, ApJ, 556, 357 ApJ, 644, 560 Ballester, G. E., Sing, D. K., & Herbert, F. 2007, Nature, 445, 511 Barman, T. S., Hauschildt, P. H., & Allard, F. 2001, ApJ, 556, 885 -. 2005, ApJ, 632, 1132 Barman, T. S., Hauschildt, P. H., Schweitzer, A., Stancil, P. C., Baron, E., & Allard, F. 2002, ApJ, 569, L51 Brown, T. M. 2001, ApJ, 553, 1006 Burrows, A. & Sharp, C. M. 1999, ApJ, 512, 843 Charbonneau, D., Allen, L. E., Megeath, S. T., Torres, G., Alonso, R., Brown, T. M., Gilliland, R. L., Latham, D. W., Mandushev, G., O'Donovan, F. T., & Sozzetti, A. 2005, ApJ, 626, 523 Charbonneau, D., Brown, T. M., Noyes, R. W., & Gilliland, R. L. 2002, ApJ, 568, 377 Deming, D., Seager, S., Richardson, L. J., & Harrington, J. 2005b, Nature, 434, 740 Fegley, B. J. & Lodders, K. 1996, ApJ, 472, L37+ Fortney, J. J., Cooper, C. S., Showman, A. P., Marley, M. S., & Freedman, R. S. 2006, ApJ, 652, 746 Fortney, J. J., Sudarsky, D., Hubeny, I., Cooper, C. S., Hubbard, W. B., Burrows, A., & Lunine, J. I. 2003, ApJ, 589, 615 Harrington, J., Hansen, B. M., Luszcz, S. H., Seager, S., Deming, D., Menou, K., Cho, J. Y.-K., & Richardson, L. J. 2006, Science, 314, 623 Hauschildt, P. H. 1992, JQSRT, 47, 433 Hubbard, W. B., Fortney, J. J., Lunine, J. I., Burrows, A., Deming, D., Brown, T. M., Charbonneau, D., Harrington, J., & Sudarsky, D., & Pinto, P. 2001, ApJ, 560, 413 Richardson, L. J. 2005a, ApJ, 622, 1149 Iro, N., B´ezard, B., & Guillot, T. 2005, A&A, 436, 719 Knutson, H. A., Charbonneau, D., Noyes, R. W., Brown, T. M., & Gilliland, R. L. 2007, ApJ, 655, 564 Menou, K., Cho, J. Y.-K., Seager, S., & Hansen, B. M. S. 2003, ApJ, 587, L113 Richardson, L. J., Deming, D., Horning, K., Seager, S., & Harrington, J. 2007, ArXiv Astrophysics e-prints Seager, S. & Sasselov, D. D. 2000, ApJ, 537, 916 Vidal-Madjar, A., Lecavelier des Etangs, A., D´esert, J.-M., Ballester, G. E., Ferlet, R., H´ebrard, G., & Mayor, M. 2003, Nature, 422, 143 5
astro-ph/0405369
1
0405
2004-05-19T15:48:29
The type IIn supernova 1994W: evidence for the explosive ejection of a circumstellar envelope
[ "astro-ph" ]
We present and analyse spectra of the Type IIn supernova 1994W obtained between 18 and 203 days after explosion. During the luminous phase (first 100 d) the line profiles are composed of three major components: (i) narrow P-Cygni lines with the absorption minima at -700 km/s; (ii) broad emission lines with BVZI ~4000 km/s; and (iii) broad, smooth wings, most apparent in H-alpha. These components are identified with an expanding circumstellar (CS) envelope, shocked cool gas in the forward post-shock region, and multiple Thomson scattering in the CS envelope, respectively. The absence of broad P-Cygni lines from the supernova is the result of the formation of an optically thick, cool, dense shell at the interface of the ejecta and the CS envelope. We model the supernova deceleration and Thomson scattering wings to recover the density, radial extent and Thomson optical depth of the CS envelope during the first month. We reproduce the light curve with a hydrodynamical model and find it to be powered by a combination of internal energy leakage after the explosion of an extended pre-supernova (~10^15 cm) and luminosity from circumstellar interaction. We recover the pre-explosion kinematics of the CS envelope: it is close to homologous expansion with outer velocity ~1100 km/s and a kinematic age of ~1.5 yr. The CS envelope's high mass and kinetic energy, combined with its small age, strongly suggest that the CS envelope was explosively ejected about 1.5 yr before the supernova explosion.
astro-ph
astro-ph
Mon. Not. R. Astron. Soc. 000, 1 -- 22 (2004) Printed 25 November 2018 (MN LATEX style file v2.2) The type IIn supernova 1994W: evidence for the explosive ejection of a circumstellar envelope Nikolai N. Chugai1⋆, Sergei I. Blinnikov2, Robert J. Cumming3, Peter Lundqvist3, Angela Bragaglia4, Alexei V. Filippenko5, Douglas C. Leonard6, Thomas Matheson5,7, Jesper Sollerman3 1Institute of Astronomy, RAS, Pyatnitskaya 48, 109017 Moscow, Russia 2Institute for Theoretical and Experimental Physics, 117218 Moscow, Russia 3Stockholm Observatory, Department of Astronomy, Stockholm University, AlbaNova University Center, SE-106 91 Stockholm, Sweden 4Osservatorio Astronomico di Bologna, via Ranzani 1, 40127 Bologna, Italy 5Department of Astronomy, University of California, Berkeley, CA 94720-3411 USA 6Five College Astronomy Department, University of Massachusetts, Amherst, MA 01003-9305 USA 7Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Cambridge, MA 02138 USA Accepted 12 May 2004. Received 11 May 2004; in original form 29 March 2004. ABSTRACT We present and analyse spectra of the Type IIn supernova 1994W obtained be- tween 18 and 203 days after explosion. During the luminous phase (first 100 days) the line profiles are composed of three major components: (i) narrow P-Cygni lines with the absorption minima at −700 km s−1; (ii) broad emission lines with blue ve- locity at zero intensity ∼ 4000 km s−1; and (iii) broad, smooth wings extending out to at least ∼ 5000 km s−1, most apparent in Hα. These components are identified with an expanding circumstellar (CS) envelope, shocked cool gas in the forward post- shock region, and multiple Thomson scattering in the CS envelope, respectively. The absence of broad P-Cygni lines from the supernova (SN) is the result of the forma- tion of an optically thick, cool, dense shell at the interface of the ejecta and the CS envelope. Models of the SN deceleration and Thomson scattering wings are used to recover the density (n ≈ 109 cm−3), radial extent [∼ (4 − 5) × 1015 cm] and Thomson optical depth (τT > ∼ 2.5) of the CS envelope during the first month. The plateau-like SN light curve is reproduced by a hydrodynamical model and is found to be powered by a combination of internal energy leakage after the explosion of an extended pre- supernova (∼ 1015 cm) and subsequent luminosity from circumstellar interaction. The pre-explosion kinematics of the CS envelope is recovered, and is close to homologous expansion with outer velocity ∼ 1100 km s−1 and a kinematic age of ∼ 1.5 yr. The high mass (∼ 0.4 M⊙) and kinetic energy (∼ 2 × 1048 erg) of the CS envelope, com- bined with small age, strongly suggest that the CS envelope was explosively ejected ∼ 1.5 yr prior to the SN explosion. Key words: supernovae -- circumstellar matter -- stars: supernovae: individual (SN 1994W) 1 INTRODUCTION Recent studies of supernovae (SNe) have made considerable progress in our knowledge of what makes a star explode. From observations of light curves and spectra, theories of pre-supernova evolution are now better constrained than ever. Conventional wisdom says that Type II SNe (SNe II) are caused by core collapse in massive, usually red super- giant stars. In general, theory has no trouble accounting for the main features of the spectra and light curves of these ob- jects. However, some SNe II are remarkably different, and many of their observational properties are not yet under- stood. ⋆ E-mail: [email protected] (NNC); (RJC); [email protected] (PL) [email protected] These events, sometimes known as narrow-line Type II SNe or SNe IIn (Schlegel 1990; Filippenko 1997), show 2 Nikolai N. Chugai, Sergei I. Blinnikov, Robert J. Cumming et al. 1987F (Filippenko 1988Z (Filippenko in early spectra the presence of strong, narrow Balmer emission lines on top of broad emission lines. Exam- ples of prominent SNe IIn include 1978K (Ryder et al. 1993), SNe 1983K (Niemela, Ruiz & Phillips 1985), 1984E (Dopita et al. 1984), 1986J (Rupen et al. 1987; Leibundgut 1989; Wegner & Swanson 1991), 1994), 1991; Stathakis & Sadler 1991; Turatto et al. 1993; Benetti et al. 1998), 1994W (Cumming, Lundqvist & Meikle 1994; Meikle et al. 1994; Sollerman, Cumming & Lundqvist 1998, hereafter SCL98), 1995G (Pastorello et al. 2002) and 1995N (Fransson et al. 2002). More recent examples of SNe II with narrow lines are SNe 1997cy (Germany et al. 1994; Turatto et al. 2000), 1998S (Filippenko & Moran 1998; Bowen et al. 2000; Leonard et al. 2000; Fassia et al. 2001; Chugai et al. 2002; Fransson et al. 2004), 1999E (Filippenko 2000; Rigon et al. 2003) and, most remarkably, SN 2002ic, which has been identified as a likely SN Ia in a dense circumstellar (CS) envelope (Hamuy et al. 2003), are more recent examples of SNe with narrow lines. The general wisdom is that the narrow lines of SNe IIn originate from the ionized, dense circumstellar gas (Henry & Branch 1987; Filippenko 1991). The interaction of SN ejecta with the dense CS gas modifies the SN op- tical spectrum and may fully power the SN IIn luminos- ity (Chugai 1990, 1992). Interaction with the dense CS gas is indicated also by strong radio and X-ray flux detected in some SNe IIn. SNe IIn are diverse (Filippenko 1997), probably reflecting variations in CS gas density and struc- ture (smooth vs. clumpy) and SN ejecta parameters (mass and energy). On top of this there is also the possibility of asymmetry of both the ejecta and CS gas. Nevertheless, the primary factors responsible for the diversity (mass of main-sequence star, explosion mechanism, structure of pre- supernova, mass-loss mechanism and history, and progenitor binarity) are still unknown. All these uncertainties, along with the possibility of using them to probe pre-supernova behaviour prior to core collapse, add to the interest in these phenomena. SN 1994W, discovered on 1994 July 29 (UT dates are used throughout this paper) at the pre-maximum phase (Cortini & Villi 1994), is among the brightest known SNe IIn. Its proximity (25±4 Mpc; SCL98) and early dis- covery made it an ideal case for detailed study. Preliminary interpretation of the spectra showed that SN 1994W ex- ploded in the dense CS envelope with a characteristic ra- dius of ∼ 1015 cm, while the amount of ejected 56Ni was low (< 0.015 M⊙; SCL98). Here we further advance our understanding of both the CS envelope properties and the phenomenon of SN 1994W as a whole by analysing the spectra and photometry discussed by SCL98 as well as other, hitherto unpublished, spectra. In our study we take two different approaches to modelling the data. The first is the simulation of the Hα profile, which pro- vides an efficient and straightforward probe of the Thomson optical depth and density of the CS envelope. The second ap- proach is a hydrodynamical simulation of the SN explosion and the light curve based upon the upgraded version of the code stella (Blinnikov et al. 1998, 2000). An early version of this code proved effective in computing the hydrodynam- ics and light curve for interaction with a dense CS envelope in the case of SN 1979C (Blinnikov & Bartunov 1993). This paper has the following structure. We start with a description of the data (Sections 2 and 3) and the gen- eral picture which arises from a qualitative analysis (Section 4). Next we quantify parameters of the CS envelope using a thin-shell deceleration model and a line-profile simulation (Section 5). We then use the gross parameters of the recov- ered density distribution as input for detailed hydrodynam- ical modelling (Section 6). Our findings are summarized and discussed in Section 7. 2 OBSERVATIONS 2.1 Photometry The photometric data were compiled and described in SCL98. In our light-curve models in Section 3 we use the photometry up to day 197, assuming SN 1994W exploded on 1994 July 14.0 (SCL98). For a discussion of the explosion date, see Section 3. 2.2 Spectroscopy The spectroscopic observations are described in Table 1. Some of these, taken from La Palma, were presented in Ta- ble 2 of SCL98. We supplement them with spectra taken at the Lick, Keck, and Bologna Observatories. The data cover epochs 18 -- 203 days after explosion. We now describe these spectra in detail. The La Palma spectra were reduced using the figaro (Shortridge 1990) and NOAO iraf1 packages. The CCD frames were bias-subtracted, flat-fielded, and corrected for distortion and cosmic rays. Wavelength calibration was car- ried out using arc spectra of copper-neon and copper-argon lamps, and should be accurate to better than 0.1 A for the high-resolution WHT and INT spectra, 0.5 A for the low- resolution WHT and INT spectra, and 1 A for the NOT spectrum. The Bologna spectrum was obtained with the Bologna Astronomical Observatory 1.5-m telescope with the spec- trograph BFOSC on 1994 July 31.9 and reported by Bragaglia et al. (1994). Spectra were also obtained at the Shane 3 m reflec- tor at Lick Observatory on days 21, 49, 57, 79 and 89, and at the W. M. Keck II 10 m telescope on day 197. The Lick observations were made with the Kast double spec- trograph (Miller & Stone 1993). The Keck spectrum was obtained using the Low Resolution Imaging Spectrometer (LRIS; (Oke et al. 1993)) with a 1200 lines mm−1 grating. For details, see Table 1. For the Lick data, all one-dimensional sky-subtracted spectra were extracted in the usual manner, and each spec- trum was then wavelength and flux calibrated, as well as corrected for continuum atmospheric extinction and tel- luric absorption bands (Wade & Horne 1988; Bessell 1999; Matheson et al. 2000). 1 IRAF (Image Reduction and Analysis Facility) is distributed by the National Optical Astronomy Observatories, which are oper- ated by the Association of Universities for Research in Astronomy, Inc., under cooperative agreement with the US National Science Foundation. Table 1. Log of spectroscopic observations Circumstellar envelope ejection in SN 1994W 3 Phase(a) Telescope/ spectrograph Spectral res. (A) Flux standard Seeing Wavelength coverage (A) (arcsec) Observers BD+25◦3941 -- BD+17◦4708 SP1337+705 BD+28◦4211, BD+26◦2606 Grw+70◦5824 Grw+70◦5824 BD+28◦4211, BD+26◦2606, BD+17◦4711 Feige 110, BD+26◦2606 Feige 110, BD+26◦2606 -- HD 19445 Feige 34 1.6 2.6 -- 3.0 2.5 1.1 -- 1.7 4+ 0.9 -- 2.9 1.0 -- 1.5 3 -- 5 2 -- 3 3 -- 4 1.6 1.2 0.9 4600-7800 6507 -- 6675 4250 -- 7020 4290 -- 8005 3120 -- 10400 3300 -- 9360 4255 -- 5060, 6345 -- 7150 3140 -- 9920 AB EZ AJB, AF, CP MB AJB, AF, TM RR RR AJB, AF, TM 3130 -- 8040 AF, LH, TM 3160 -- 8020 AF, TM 5485-7370 5628 -- 6939 5000-10000 RR, NO AF, LH JS, RC 11 2.7 4 4 4 6 6 12 Date (UT) 1994 07 31.9 1994 07 31.9 1994 08 04 1994 08 13.9 1994 09 01 1994 09 08.9 1994 09 08.9 17.9 17.9 21.5 30.9 49.5 56.9 56.9 BAO/BFOSC 13 INT/IDS Lick/Kast WHT/ISIS Lick/Kast 0.25 7 2.7 5 WHT/ISIS WHT/ISIS 1994 09 09 57.5 Lick/Kast 1994 10 01 79.5 Lick/Kast 1994 10 11 89.5 Lick/Kast 1994 11 12.3 1995 01 26.6 1995 02 01.9 121.3 196.6 202.9 INT/IDS Keck/LRIS NOT/LDS Note: (a) Days after 1994 July 14.0. (b) Observers: AB -- A. Bragaglia; AF -- A. Filippenko; AJB -- A. J. Barth; CP -- C. Peng; EZ -- E. Zuiderwijk; JS -- J. Sollerman; LH -- L. Ho; MB -- M. Breare; NO -- N. O'Mahony; TM -- T. Matheson; RC -- R. Cumming; RR -- R. Rutten. The spectra were flux-calibrated by comparison with the standard stars listed in Table 1. The sky position of SN 1994W in late 1994 meant that the observations were made when the supernova was at rather low altitude, and correspondingly high airmass. We corrected for the differ- ence in atmospheric extinction brought about by the differ- ing zenith distances of the supernova and the standard star. Airmasses at different zenith distances were taken from the relation given by Murray (1983), and we used the tables of extinction coefficient versus wavelength for La Palma given by King (1985). With two exceptions, all our spectra were taken at the parallactic angle (Filippenko 1982), minimising slit losses due to atmospheric dispersion. The first exception was the day 18 Bologna spectrum of SN 1994W, which shows con- siderable loss of flux in the blue. We corrected the slope of the spectrum to match the B and V photometry from the same night reported by Bragaglia et al. (1994) using filter functions from Gualandi & Merighi (2001). The second exception was the standard-star observation for the day 31 La Palma spectrum. While the SN 1994W spectrum was taken at the parallactic angle, the standard- star spectrum was taken at an airmass of 1.6, with the slit at position angle 33◦ when the parallactic angle was 119◦. No contemporary filter photometry is available for correcting the supernova spectrum. The effect on the day 31 supernova spectrum depends on how the standard star was positioned in the slit, which in turn depends on the combination of the spectral response of the acquisition camera at the telescope. The acquisition camera at the WHT in 1994 August was a Westinghouse ETV-1625 whose response curve peaked at 4400 A, with half-power points at 3400 A and 6200 A (C. Jackman & D. Lennon, 2004, private communication). The standard star, Grw +70◦5824, is a DA3 white dwarf with a blue spectrum. It therefore seems likely that the slit was positioned so that maximum transmission was at around 4400 A, at the blue end of the spectrum. We have estimated the amount of flux that would be lost with wavelength, and find that the losses would largely have been at the red end of the standard spectrum -- as much as 80 per cent of the flux at 8000 A. The effect on the calibrated supernova spectrum was thus to make it ap- pear redder than it really was. Since the day 31 spectrum is already remarkably blue (Section 3.1.3), we have made no further correction to the spectrum; the calibration can be regarded as a conservative estimate of the already extreme continuum slope. No observations of flux standards were available for the INT spectra taken on days 18 and 121. A very rough flux- calibration of the day 18 INT spectrum was made by assum- ing that the flux in the wings of the Hα line (at the red and blue edges of the spectrum) declined at the same rate as the visual magnitude estimates between days 18 and 31, i.e., by a factor of about 1.6 in fλ. A similarly rough calibration of the day 121 spectrum was made by comparing with our R-band photometry from day 123 (SCL98). The spectrum taken on day 197 was scaled so that the flux in the narrow Hα feature matched that in the day 203 spectrum. The continua then match well in the overlap re- gion. Spectra from days 18, 57, 121 and 203 were displayed in SCL98 (their Figs. 2 and 3). They also showed the portion of the spectrum on day 31 covering Hα. As a final correction, we noted that all the spectra cover the entire V band, and we scaled each spectrum to match contemporary V -band photometry. The procedure was as follows. We assumed that the day 57 La Palma spectrum was correct. Next, we interpolated the V magnitudes in the light 4 Nikolai N. Chugai, Sergei I. Blinnikov, Robert J. Cumming et al. Figure 1. Spectra of SN 1994W: days 18 -- 203. The spectra have been shifted vertically for clarity. No heliocentric correction has been applied to the wavelength scale. The ticks on the right-hand side mark the zero level for each spectrum. The spectra from day 79 and later have been multiplied by a constant, noted in parentheses. The La Palma spectra from days days 18 and 57 and the day 197 Keck spectrum are not shown. Telluric absorption lines in the day 31 spectrum are marked. curve presented by SCL98. This gave us relative fluxes in V for all the epochs. We then integrated all the plateau spec- tra (days 21 -- 89), multiplying them by a standard V filter function (Allen 1973) with transmission values interpolated between points at 100-A intervals. The whole spectrum was then scaled linearly in flux so that the integrated V magni- tude corresponded to the V -band photometry in SCL98. the spectral development for each epoch in turn, and then concentrate on the individual features and ionic species. In this section, quoted line fluxes (see Table 3) were measured by interpolating the neighbouring continuum across the line in question. The line fluxes are net fluxes and do not correct for the narrow P-Cygni absorption com- ponents. 3 RESULTS 3.1 Spectral evolution 3.1.1 Day 18 During the period of our observations, the spectrum of SN 1994W (Figs. 1 and 2) was dominated by narrow Balmer emission lines, initially with P-Cygni absorption features, and accompanied by a wide variety of weaker lines show- ing broad emission and/or narrow P-Cygni profiles (cf. Filippenko & Barth 1994). Line identifications are presented in Table 2 and in Fig. 3. In the following sections, we examine The low-resolution Bologna spectrum showed a blue contin- uum with prominent lines of Hα and Hβ, which are nar- row with broad wings. The latter have been interpreted as an effect of multiple Thomson scattering of line photons in the CS envelope (Chugai 2001). Our high-resolution spec- trum for this epoch (Fig. 3 in SCL98) covers only the Hα line, but shows a narrow P-Cygni absorption at about −800 Circumstellar envelope ejection in SN 1994W 5 Figure 2. Spectra of SN 1994W: days 18 -- 121. Same as Fig. 1, but with the ordinate stretched to show the weaker features more clearly. The day 197 and 203 spectra are omitted. Telluric absorption lines in the day 31 spectrum are marked. km s−1 not visible in the low-resolution spectrum. The emis- sion peak is close to the velocity of nearby H ii emission, measured at 1249±3 km s−1, which we adopt as our best es- timate of the heliocentric velocity of the supernova. The in- terstellar Na i D lines give a somewhat lower value, 1185±15 km s−1, measured from the day 31 spectrum. 3.1.2 Day 21 The rapid brightening of the supernova at this epoch (SCL98) was reflected in increasingly strong Balmer lines. The spectrum showed a strong blue continuum, and promi- nent lines of Hα, Hβ and Hγ, all exhibiting narrow P- Cygni profiles with broad emission wings. The Balmer decre- ment was remarkably shallow, with flux ratio Hα:Hβ:Hγ = 1.1:1:0.5 (dereddened assuming EB−V = 0.17 mag from SCL98 and the reddening law of Cardelli, Clayton & Mathis 1989). The Balmer lines were accompanied by an asymmet- ric triangular broad emission feature corresponding to He i λ5876, possibly blended with the Na i D doublet, and many narrow P-Cygni lines of Fe ii. 3.1.3 Day 31 By day 31, the blue continuum appeared to have steepened. The calibration of the spectrum is uncertain, as noted in Section 2 above, but the evidence suggests that the spec- trum really was as remarkably blue as it appears to be. The Balmer lines still show a flat decrement, as on day 21, with measured ratios 1.2:1:0.6. The blue wings in Hβ and Hγ ap- pear to be stronger relative to the red wings, compared to day 21 (Fig. 5, right-hand panel). Broad, triangular He i emission lines with full-width at half-maximum velocity width (vFWHM) around 2500 km s−1 were stronger than on day 21 (Figs. 2 and 5). He i λ5876 was accompanied by strong He i λ7065. Broad He i λ6678 appeared to fill out the red wing of Hα, and the triplet line at λ4471 is present as broad emission which underlies neighbouring Fe ii lines. All but He i λ7065 also showed 6 Nikolai N. Chugai, Sergei I. Blinnikov, Robert J. Cumming et al. Figure 3. Line identifications for days 49 (upper spectrum) and 79. The spectra have been continuum-subtracted and shifted for clarity. Each line is indicated by a broken vertical bar marking heliocentric velocities −700 km s−1 and 0 km s−1. Horizontal bars longward of 6700 A mark regions where telluric absorption has been removed. weak, narrow P-Cygni features with minima around −700 km s−1. A similarly triangular emission feature is seen at 7890 A, which we identify with the Mg ii triplet at 7877 -- 7896 A (see Section 3.3.4). Stronger narrow P-Cygni lines with absorptions around −600 km s−1 are seen in O i λ7773, Si ii λλ6347, 6371, and a large number of Fe ii lines. The O i and Si ii lines are dom- inated by absorption, while in Fe ii the equivalent widths of the absorption and emission components are comparable. 3.1.5 Day 57 Eight days later, the Balmer decrement showed no signifi- cant change at 1.5:1:0.6. The spectrum showed ever stronger Fe ii lines. No unambiguous trace of He i remained (Fig. 2), suggesting that the emission feature at ∼6000 A was now only due to Na i D. The emission in this feature broadened and flattened out, extending from −4000 km s−1 to +6000 km s−1 (Figs. 2 and 5). The Mg ii λ7877 -- 7896 feature was still present, but appeared more flat-topped than on day 49. P-Cygni lines of O i λ8446 and the Ca ii near-infrared triplet, also weakly present on day 49, were clearly detected in the red. 3.1.4 Day 49 The day 49 spectrum (Figs. 1, 2 and 3) shows a redder con- tinuum. The Balmer decrement steepened to 1.5:1:0.6. The He i λ5876 and Na i D blend was again asymmetric as on day 21, and appeared to have a broader blue wing. The only other remaining He i line, λ7065, is no longer as prominent as on day 31. The Mg ii λ7877 -- 7896 feature was still clearly visible. Similar broad features were detected at ∼8230 A and ∼9220 A, lending support to the identification with Mg ii. Many new Fe ii lines are apparent, plus Ca ii H & K and possibly the Ti ii multiplet at around 3500 A. The wider wavelength coverage picked up a large number of overlapping narrow P-Cygni lines in the blue; most of the identified ones are H i and Fe ii. 3.1.6 Day 79 The continuum appeared again to have reddened (Fig. 1). The Balmer lines decreased in strength relative to the now very numerous Fe ii P-Cygni lines (Figs. 2 and 3). The Hα/Hβ ratio was now 2.7±0.3, close to Case B. Narrow over- lapping absorption and P-Cygni features completely domi- nated in the blue. In the red, the feature at 7890 A had disappeared, but O i λ8446 and the Ca ii triplet had in- creased in strength. The narrow absorption in Na i D had increased markedly in strength since day 57, and broad Na i emission was no longer apparent. Narrow P-Cygni profiles of O i λ7002 and Sc ii ap- peared, the latter possibly accompanied by a broad emission Circumstellar envelope ejection in SN 1994W 7 Table 2. Line identifications for days 18 -- 121. Symbols are as follows: a -- line seen in absorption only; b -- broad emission, e -- emission; p -- P-Cygni profile; -- (dash) -- no feature detected, : (colon) -- uncertain identification. The Fe ii lines were identified with the help of the line lists of Fuhr, Martin & Wiese (1988) and Sigut & Pradhan (2003). Wavelengths in A are from these references and from van Hoof (1999). Identification 18 21 31 49 57 79 89 121 Identification 18 21 31 49 57 79 89 121 Fe ii 3764.11 +H11 3770.63 H10 3797.90 H9 3835.38 Mg ii 3849.3 H8 3889.05 Fe ii 3914.51 Ca ii 3933.66 Ca ii 3968.47 + Hǫ 3970.07 Hδ 4101.73 Fe ii 4233.12 Fe ii 4296.57 +Fe ii 4303.17 Hγ 4340.46 Fe ii 4351.76 Fe ii 4385.38 Fe ii 4416.83 He i 4471.5 Mg ii 4481.2 Fe ii 4520.21 Fe ii 4549.46 +Fe ii 4555.88 Fe ii 4576.34 +Fe ii 4582.83 Fe ii 4629.34 Hβ 4861.32 Fe ii 4923.94 Fe ii 5018.44 Fe ii 5169.05 Fe ii 5197.59 Fe ii 5234.63 +Sc ii 5239.81 Fe ii 5264-84 Fe ii 5316.62 +Fe ii 5325.54 Fe ii 5362.85 Fe ii 5425.26 Fe ii 5477.66 Sc ii 5526.79 p p p p p -- p pb pb p -- pb -- p p -- p p p p p p p -- p pb pb p p pb p p p -- p p -- p p -- p p p pb pb p p pb p -- p -- -- -- -- -- -- pb -- -- -- -- p -- -- pb -- -- -- ba p -- p p p p p p p pb pb pb p p -- p p -- -- -- -- p -- pb pb pb p -- -- p -- -- -- -- p p pb pb pb p p -- p p -- -- -- -- p p pb pb pb p p p p p p pe -- pb p p pb pb pb p p p p p p pe p pb -- p p -- p p p pb pb p p pb p -- p -- -- p p p pb pb pb p p p p p p pe p pb -- -- -- -- -- -- -- -- -- -- -- -- pb p: p: -- -- -- -- -- -- -- -- -- p: -- -- -- -- -- -- -- -- -- -- -- pb -- -- -- -- -- -- -- -- -- p Sc ii 5657.90 +Sc ii 5658.36 He i 5875.61 Na i 5889.95 +Na i 5895.92 Fe ii] 5990.60 +Fe ii] 5991.38 Fe ii] 6150.10 Sc ii 6245.63 +Fe ii 6247.58 Si ii 6347.11 Si ii 6371.37 Fe ii 6416.91 Fe ii 6456.38 Fe ii 6517.02 Hα 6562.80 He i 6678.15 O i 7002.20 He i 7065.22 [Fe ii] 7214.71 Ca ii] 7291.47 Fe ii 7308.07 +Fe ii 7310.22 Ca ii] 7323.89 Fe ii 7449.34 Fe ii 7462.41 Fe ii 7711.72 O i 7773.8 Mg ii 7877-96 Mg ii 8213.98 +Mg ii 8234.64 O i 8446.36 Ca ii 8498.03 Ca ii 8542.09 Ca ii 8662.14 Mg ii 9218.25 +Pa9 9229.01 +Mg ii 9244.26 Paǫ 9545.97 Paδ 10049.4 +Mg ii 10092.1 -- b -- pb -- b: -- b: b: b: ba: ba: -- p pb p p p p p pb -- -- -- -- -- -- -- -- -- p p b b: p -- p p -- -- -- p p -- -- -- pb p: -- -- -- -- p p -- -- -- pb pb -- b -- -- -- -- -- -- -- p b -- -- pb p p -- -- -- pb -- -- b -- -- -- -- -- -- p p b b: p -- p -- b pb: b: -- ep: -- -- -- -- -- -- -- -- e -- -- -- -- e -- e a -- p ep p pb p p p p p pb -- p -- p -- p -- p p p p -- a -- p ep p pb p p p p p pb -- p -- p -- p -- p p p p -- -- p p p p b pb: b pb: component. Semi-forbidden lines of Fe ii] λλ5991, 6150 were clearly detected for the first time. 3.1.7 Day 89 The spectrum on day 89 was essentially the same as on day 79, but was weaker and showed a redder continuum. The Hα/Hβ ratio increased to 5.2±1.3, and Hγ was now blended with Fe ii lines. Both Na i D and O i λ7002 continued to increase in strength. 3.1.8 Day 121 Our spectrum for day 121 (Figs. 1 and 2) has smaller wave- length coverage and rather poor signal-to-noise ratio, but it is obvious that a dramatic change had occurred since day 89, paralleling the supernova's precipitous fading after day ∼110 (SCL98). As reported by SCL98, three narrow emission lines are unambiguously detected: Hα with a vFWHM of 780 km s−1, plus much fainter Na i D (vFWHM = 775 km s−1) with a P- Cygni absorption component, and Ca ii] λ7291 (vFWHM = 475km s−1) close to the red edge of the spectrum. The emis- sion peak of Hα had moved farther to the red than on day 89, at about +70±100 km s−1. The Sc ii P-Cygni feature at 5530 A appears also to have survived since day 89, showing a sharp absorption feature at −800 ± 100 km s−1. Underlying the narrow features is what appears to be a low, undulating continuum. SCL98 argued that the sharp drop to zero flux at 5620 A was evidence that the contin- 8 Nikolai N. Chugai, Sergei I. Blinnikov, Robert J. Cumming et al. Axelrod 1980). These similarities suggest that SN 1994W had reached its nebular phase at this epoch. 3.1.9 Day 197 Our spectrum, not shown in Figs. 1 and 2, detected only Hα, which was perhaps just resolved. Its velocity width was vFWHM ≈ 625 km s−1, consistent with no change since day 121. Na i emission at the same strength relative to Hα as on day 121 would not have been detected, given the low signal-to-noise ratio of the spectrum. 3.1.10 Day 203 The day 203 spectrum (Fig. 1) confirms that narrow, un- resolved Hα is the only detectable feature from the su- pernova. The galaxy background probably dominates the remaining, noisy continuum, though photometry indicates that the supernova can account for up to 50 per cent of the flux (SCL98). SCL98 put limits on the flux in broad lines of width ∼4500 km s−1. Figure 4. Luminosity evolution of various lines. Line fluxes are detailed in Table 3. A distance of 25.4 Mpc and an extinction of EB−V = 0.17 mag were adopted. 3.2 General spectral appearance 3.2.1 Line profiles The spectra reveal two major types of line profile-forming components (SCL98; Fig. 5): narrow P Cygni profiles and broad lines. However, careful inspection reveals that while the blue velocity at zero intensity (BVZI) of the broad lines does not exceed −5000 km s−1, Hα at the luminous stage shows a red wing extending at least up to +7000 km s−1. This wing is clearly detectable in Hα on days 31, 49 and 57 (Figs. 2 and 5). We believe that this wing is caused by a Thomson scattering effect in the expanding CS envelope (see also Section 4). Since the intensity of the Thomson scatter- ing wing is proportional to the line intensity, the wing may be present but not obvious in other, weaker lines. The intrinsically broad lines typically have a triangular profile. On day 31 they show an apparent skewing toward the blue, possibly due to an occultation effect. The Na i doublet on day 57 shows a flat-topped profile with BVZI≈ 4000 km s−1. No other line shows such a flat-topped profile, nor does Na i doublet at other epochs, though the feature is dominated by He i λ5876 earlier than this. This suggests that the material emitting Na i has a narrower distribution of velocity than the other broad lines (see Section 4.2). 3.2.2 Line strengths Assuming a distance of 25.4 Mpc to NGC 4041 and EB−V = 0.17 mag (SCL98), we have plotted the luminosity evolution of the strongest emission lines in Fig. 4. Note the quasi- exponential decline of the Hα luminosity with exponential lifetime ∼ 21 days. This fast decline is in dramatic con- trast to other SNe IIn, for example SN 1995G, whose nar- row Hα luminosity dropped by only a factor of three during the first 600 days (Pastorello et al. 2002). The fast decay of the Hα luminosity is consistent with a small amount of 56Ni in SN 1994W (SCL98) and also indicates the absence of dense CS gas at large distances from the pre-supernova (see Section 7). Figure 5. Selected line profiles; day numbers are indicated. The left-hand panel shows examples of narrow P-Cygni profiles, the middle panel shows broad lines (note the extended wings in Na i on day 57), and the right-hand panel shows the wings of the H i lines on day 31. The dotted vertical lines mark velocities of 0 and −700 km s−1. uum was rather composed of broad, overlapping emission lines. The steepness of the drop at 5620 A indicates that the velocity widths of such lines can be no greater than ∼5000 km s−1. Similar features have been seen in late-time spectra of the SN IIn 1997cy (Germany et al. 1994; Turatto et al. 2000). A sharp drop at this wavelength is in fact also typ- ical of SNe Ia and some SNe Ib/Ic at late times, when the emission is dominated by lines of Fe ii and [Fe ii] (see Circumstellar envelope ejection in SN 1994W 9 Table 3. Flux measurements of emission lines, in units of 10−14 erg s−1 cm−2. Line Hγ 4340.46 Hβ 4861.32 Fe ii 5018.44 He i 5875.62 +Na i(a) 5889-95 Na i(b) 5889-95 Hα 6562.80 He i 7065.22 Mg ii(a) 7877-7896 Mg ii 9218-44 18 -- 30(10) -- -- -- 25(5) <2 -- -- 21 30(2) 61(3) 31 47(4) 91(5) 49 34(1.5) 60(5) 1.7(0.1) 0.7(0.1) 2.1(0.5) 13.6(2) -- 82(3) -- -- -- 14(4) -- 130(6) 9.7(3) 10(2.5) -- 12(1) -- 110(2) 2.9(0.7) 5.2(0.5) -- Notes: (a) blend, (b) emission component only. Error estimates are 1σ. Upper limits are 3σ. Day 57 24.5(2) 47(2) 2.7(0.7) 79 -- 10(1) 89 -- 4(1) 2.0(0.7) 1.6(0.2) 5(1) <1 <1 121 203 -- -- -- -- -- -- -- -- 0.14(0.05) 0.2(0.02) 0.6(0.1) 0.11(0.03) <0.03 84(2) <0.7 5.0(0.3) 6.3(1) 32(3) <0.5 <0.6 21(1) <0.2 <0.6 7.5(0.8) 1.0(0.3) 2.2(0.2) 0.1(0.025) -- -- -- -- -- -- 3.2.3 Continuum We carried out black-body fits to the continuum. We obtain colour temperatures of 13000±2000 K on day 21, 15000 K on day 31, 10000 K on days 49 and 57, 7200 K on day 79 and 7200 K on day 89. We find that the black-body fits work best for EB−V = 0.15 mag, consistent with our estimate from Na i absorption (0.17 ± 0.06 mag; SCL98). 3.3 Elements identified 3.3.1 Hydrogen We detect the Balmer series from Hα to at least H11, and possibly also members of the Paschen series from Paδ to Pa11 (Fig. 3). The profiles seem to be a combination of ∼ 4000 km s−1 and a narrow broad emission line with BVZI > P-Cygni absorption component (∼ 1000 km s−1). The ra- tio of narrow-to-broad emission component decreases from Hα toward the higher Balmer lines. This indicates a nearly normal Balmer decrement for the narrow component and a flat or possibly inverse Balmer decrement for the broad component (Fig. 5). A flat or inverted decrement suggests strong deviation from the recombination case, related to collisions and ra- diative transitions in the field of the photospheric radia- tion. The observed line ratios suggest an excitation tem- perature for hydrogen's excited levels of around 15000 K. Two different thermalization mechanisms for hydrogen lev- els may operate in this situation: the level populations may be controlled by radiative transitions in the field of the external black-body photospheric radiation, or thermaliza- tion may be primarily collisional. Both mechanisms have been exploited to account for flat or inverted Balmer decre- ments observed in cataclysmic variables (Elitzur et al. 1983; Williams & Shipman 1988). Both thermalization mecha- nisms are plausible in SN 1994W as well. 3.3.2 Helium We observe broad He i lines with triangular emission profiles and narrow P-Cygni features. The He i lines disappear after day 49 (Figs. 2 and 4; Table 2). He i λ5876 and λ7065 appear only in emission. He i λ4471 and λ6678 show weak absorp- tion features. A high ratio I(7065)/I(5876) ≈ 0.5 indicates a large contribution of collisional and/or radiative excitation compared to the purely recombination case. Unfortunately, this fact cannot be used as a straightforward indicator of electron concentration in the line-emitting region. Computa- tions for an extended parameter set (Almog & Netzer 1989) show that this particular ratio may be reached in a wide range of concentrations (ne ≈ 104 − 1014 cm−3), with pro- nounced dependence on the optical depth in He i λ3889. 3.3.3 Oxygen O i λ7773 is present with an absorption-dominated P-Cygni profile from as early as day 18 to day 89. From day 79, we tentatively identify O i λ7002 with a P-Cygni profile. This line has not to our knowledge been seen in a supernova spectrum before, but was identified in the proto-planetary nebula Henize 401 (Garc´ıa-Lario, Riera & Manchado 1999), whose low-ionization spectrum bears some similarity to that of SN 1994W at these phases. 3.3.4 Other metals Ca ii H & K were strong in absorption at all epochs when we covered them. The near-infrared triplet lines were weak on day 49, but strengthened dramatically by day 79 (Fig. 3). This may reflect an increase in the Ca ii/Ca iii ratio with the drop in radiation temperature at late epochs. The Na i D doublet's development is remarkable (Figs. 1, 2 and 5), even taking into account blending with inter- stellar absorption and in the earlier spectra with He i λ5876. Up to day 49, a flat-topped profile emerged as He i faded. From day 57 on, the strength of the circumstellar P-Cygni absorption increased, presumably as a result of increasing Na i ionization fraction. Magnesium lines appear both in absorption and emis- sion. Narrow absorption with weak P-Cygni emission is seen in Mg ii λ4481 from day 21 to day 57. We iden- tify Mg ii λ7877 -- 7896 as the source of the broad feature around 7890 A. As an alternative interpretation, we consid- ered [Fe xi] λ7892. Strong [Fe xi] might be expected to be accompanied by emission in [Fe x] λ6374, and there may 10 Nikolai N. Chugai, Sergei I. Blinnikov, Robert J. Cumming et al. indeed be a broad feature underlying the Si ii absorption lines to the blue of Hα. However, this feature does not per- sist to later epochs as the 7890 A feature does, nor are any other high-ionization lines identified at any epoch. In addi- tion, emission lines with similar widths to the 7890 A feature are seen at wavelengths corresponding to Mg ii transitions around 9230 A, and possibly at 8200 A and 10900 A. Never- theless, we find it somewhat surprising that the broad emis- sion in Mg ii λλ7877 -- 7896 is not accompanied by similar emission in O i λ7773. The doublet of Si ii λλ6347, 6371 is present from days 21 to 89, showing absorption-dominated P-Cygni profiles. The lines weakened relative to other features on days 79 and 89. Si ii λ6240 therefore seems to be an unlikely contributor to the feature at 6260 A (see below), which increases on these dates. Sc ii emission lines at around 5530 A and 6250 A were identified by Pastorello et al. (2002) in the spectrum of the Type IIn SN 1995G. We identify both Sc ii λ5527 and λ6246 in our spectra, and Sc ii may contribute to the feature at 6260 A. In particular, λ5527 shows some evidence of a broad emission component (Figs. 2 and 3). Its narrow P-Cygni core apparently persisted until day 121. We identify a host of Fe ii lines (Fig. 3; Table 2), all of which show narrow P-Cygni profiles. They may also be ac- companied by broad emission as in the case of other species, since the spectra show hints of underlying broad emission in regions of the spectrum where there are many Fe ii lines (Fig. 1). By days 79 and 89, a few lines of forbidden and semi-forbidden Fe ii are identified. These have emission- dominated P-Cygni profiles (Fig. 3). 4 GENERAL PICTURE The previous discussion of spectroscopic and photometric data on SN 1994W (SCL98) led to the conclusion that the narrow lines originate in a dense CS envelope. However, at present, the spectroscopic and photometric results cannot readily be incorporated into any existing model of a SN II interacting with a CS environment. We therefore present first our qualitative view of what we observe in the case of SN 1994W. We emphasize some basic elements of the physi- cal picture that, in our opinion, are crucial for understanding the phenomenon. 4.1 The opaque cool dense shell The smooth continuum of SN 1994W and its lack of broad absorption lines is typical of most SNe IIn at an early phase. It is also reminiscent of the early spectrum of SN 1998S (Leonard et al. 2000). For SN 1998S, the smooth contin- uum and the absence of broad absorption lines was ex- plained as an effect of an opaque cool dense shell (CDS) which formed at the interface of the SN with its circum- stellar medium (Chugai 2001). In this situation, an opaque CDS is physically equivalent to an expanding photosphere with a sharp boundary and without an external, extended SN atmosphere. In comparison, the CDS formed in super- nova remnants during the radiative stage (Pikelner 1954) and in SNe interacting with a moderately dense CS wind (Chevalier & Fransson 1985) are optically thin in the con- tinuum. The large optical depth of the CDS requires a rela- tively large swept-up mass. This may be a natural conse- quence of shock break-out in SNe II with extended stel- lar envelopes (Grasberg et al. 1971; Falk & Arnett 1977; Blinnikov & Bartunov 1993) and/or with extended CS en- velopes of unusually high density. In the model of a pre- supernova with initial radius ∼ 1015 cm (Falk & Arnett 1977), a CDS with a mass of ∼ 2 M⊙ forms and re- mains opaque for about 70 days. We propose that a similar opaque CDS formed in SN 1994W. This suggestion finds support in the exceptionally bright peak of the light curve, MV ≈ −19.5 mag (SCL98, their Fig. 5). According to the theory of SN II light curves, a broad luminous maximum with M ≈ −(19−20) mag requires the explosion of a very ex- tended pre-supernova with an envelope radius of R0 ≈ 1015 cm (Grasberg et al. 1971; Falk & Arnett 1977). 4.2 The broad-line region The deceleration of the CDS in the extended pre-supernova envelope and in the dense CS environment is accompa- nied by the Rayleigh-Taylor (RT) instability (Falk & Arnett 1977; Chevalier 1982). The growth and subsequent fragmen- tation of RT spikes in the dense CDS material results in the formation of a narrow layer (∆R/R ≈ 0.1 − 0.15) composed of dense filaments, sheets and knots embedded in the rar- efied hot gas behind the forward shock (Chevalier & Blondin 1995; Blondin & Ellison 2001). This mixed layer on top of the CDS is likely responsible for the broad emission lines ob- served in SN 1994W. If this is the case, then the expansion velocity of the CDS (vs) should be the same as the velocity of the broad line-emitting gas, i.e., vs ≈ 4000 km s−1. A drawback of this picture of the broad-line formation is that the velocity distribution of the line-emitting gas peaks at the CDS expansion velocity vs, which means the expected line profile should be boxy. With the exception of the Na i profile on day 57, the broad lines are instead triangular. A triangular profile can be produced, however, if the spherically symmetric velocity distribution of the line- emitting matter is close to dM/dv ∝ v. Actually, given this mass-velocity spectrum and constant emissivity per unit mass, and assuming that the velocity field is spherically sym- metric, the luminosity distribution over the radial velocity u (in units of vs) is dL du ∝ Z 1 u 1 v dM dv dv ∝ (1 − u) , (1) which is indeed triangular. The broad velocity spectrum of the dense gas in the mixed layer may naturally arise in two plausible scenarios. In the first scenario the CS gas is clumpy. In this case, an ensemble of dense CS clouds engulfed by the forward post- shock gas could provide a broad velocity spectrum of radia- tive cloud shocks and cloud fragments (Chugai 1997) in the range of 1000 -- 4000 km s−1, where the lower limit is the CS velocity. This model may be characterized as low-velocity clouds in a high-velocity flow. Alternatively, dense high-velocity clumps interacting with the low-velocity CS gas could produce a similar spec- Circumstellar envelope ejection in SN 1994W 11 trum of fragments. This could occur if the forward shock is radiative, forming a thin post-shock layer with ∆R/R < 0.1. In this situation, dense RT spikes, formed at the previous stage of RT instability of the CDS, can penetrate into the pre-shock zone. The interaction of these protrusions with the CS gas in the pre-shock zone leads to further fragmen- tation and deceleration of fragments down to the velocity of the CS gas. As a result, a broad velocity spectrum of dense cool fragments emerges in the velocity range of 1000 -- 4000 km s−1. For this second possibility to hold, a high CS density n is needed to maintain the narrow width of the post-shock cooling region, lc = mpv3 s 32nΛ , (2) where mp is the proton mass, and Λ is the cooling function in the post-shock zone. To reach the required ratio lc/R ≈ 0.1 at radius R = 1015 cm, one needs a CS density of n ≈ 1.7 × 109 cm−3 assuming Λ = 2 × 10−23 erg s−1 cm6 and vs = 4000 km s−1. Both scenarios for the velocity spectrum leave open the question of why the velocity distribution of the line-emitting gas in the mixed layer is close to dM/dv ∝ v. The following naive model seems to provide a hint. Let us consider the interaction of a high-velocity frag- ment with the CS gas. In the rest frame of the initial frag- ment, the fragmentation process in the rarefied CS flow with the velocity ≈ vs may be thought of as a stripping flow with accelerating velocity v1 and in which the "ra- dius" a of fragments progressively decreases as the velocity v1 increases (Klein, McKee & Colella 1994). Assuming mass conservation of fragments in velocity space, Q = dM/dt = (dM/dv)(dv/dt) = constant, the mass-velocity spectrum of fragments may be expressed as dM dv1 = Q(cid:16) dv1 dt (cid:17)−1 , (3) which is essentially determined by the acceleration dv1/dt. The latter can be approximately described as acceleration due to the drag force m dv1 dt = πa2ρ(1 − v1)2 , (4) where the velocity is in units of vs, m = (4π/3)a3ρf is the mass of fragment of the density ρf , and ρ is the density in the rarefied ambient flow. To determine the acceleration we must specify the relation between fragment size and velocity v = 1 − v1 relative to the CS gas. We assume that the fragment size for each new generation of cloudlets is tuned to the Kolmogorov turbulent cascade with spectrum v ∝ a1/3. With this relation, equation (4) leads to dv1/dt ∝ 1/v. The latter combined with equation (3) leads to a mass-velocity spectrum dM/dv ∝ v. A similar analysis carried out for the case of a CS cloud in the SN flow produces a quite different mass-velocity spec- trum, dM/dv ∝ (1−v). This spectrum also results in a broad line but with zero slope at u = 1 and a narrow logarithmic peak: dL/du ∝ (u − 1 − ln u). While this profile resembles the observed Hα profile, it is at odds with the triangular profile of He i lines. Below (Section 5.4) we argue against a large contribution of CS clouds in the formation of Hα. Figure 6. Circumstellar lines on day 79. Overplotted on the ob- servational data (dotted line) are the models for homologous ex- pansion (thick line) and the constant-velocity case (thin line). 4.3 The circumstellar envelope 4.3.1 Expansion kinematics The radial velocities of narrow absorption features in SN 1994W are strikingly persistent between days 18 and 89. This implies that the expansion velocity of the CS envelope (∼ 1000 km s−1) has a pre-explosion origin. An alternative possibility -- acceleration of a slow wind by SN radiation (see also SCL98) -- seems unlikely, since it would require un- realistic fine tuning between the time-dependent luminosity and absorption (scattering) coefficient to arrange the persis- tence of the velocity of accelerated gas. Two extreme options are conceivable for the CS enve- lope's pre-explosion kinematics: constant-velocity flow and homologous expansion (u ∝ r). The former might be the result of either continuous wind outflow or a succession of ejection events with similar ejecta velocities. Homologous expansion, on the other hand, might be produced by a sin- gle explosive ejection. The latter mechanism, though exotic, was in fact proposed by Grasberg & Nadyozhin (1986) to account for the narrow CS lines in SN 1983K. 4.3.2 Envelope size The light curve of SN 1994W is characterized by a broad maximum around day 30 and a subsequent plateau, ter- minated by a sudden drop in luminosity at td ≈ 110 d (SCL98, their Fig. 5). Assuming a CDS expansion velocity of vs = 4000 km s−1, we find that during this period the CDS sweeps up the CS gas within the radius Rd = vstd ≈ 4×1015 cm. Although this estimate refers to t = 110 d, we adopt it as a rough estimate for earlier epochs as well. We now argue that this radius should coincide with the outer radius of the dense CS envelope, Rcs ≈ Rd. If in- stead Rcs > Rd, then the light curve of SN 1994W should not have shown such a steep drop, since the CS interaction 12 Nikolai N. Chugai, Sergei I. Blinnikov, Robert J. Cumming et al. would have augmented the luminosity at the luminosity de- cay phase. Equally, if Rcs < Rd, the narrow lines would have disappeared at epoch t < td, which was not observed. These arguments, therefore, provide strong evidence that the ex- tent of the dense CS envelope is Rcs ≈ Rd ≈ 4×1015 cm. Be- yond this radius, the CS density presumably drops steeply, as indicated both by the broad-band light drop after day 100 and the fast decline of the Hα luminosity (Fig. 4). At larger radii, X-ray observations suggest that other density enhancements may be present (Section 7; Schlegel 1999). 4.3.3 Electron-scattering wings and circumstellar density We attribute the extended smooth wings observed in Hα and Hβ between days 18 and 89 to the effect of Thomson scatter- ing on thermal electrons participating in the bulk expansion of the CS envelope (Chugai 2001). Based on previous com- putations of Thomson scattering effects in SN 1998S, and the strength of the Hα wings in SN 1994W on day 57 (when blending with He i λ6678 was negligible), we believe that the optical depth of the CS envelope to Thomson scattering (τT) at this stage must have been close to unity. Adopting a photospheric radius Rp ≈ vst = 2 × 1015 cm on day 57 (where vs ≈ 4000 km s−1) and an outer radius of the CS envelope Rcs = 4 × 1015 cm, we obtain an estimate of the average electron concentration in the CS envelope assuming τT ≈ 1: ne ≈ τT σT(Rcs − Rp) ≈ 109 cm−3. (5) This value, taken together with the forward shock-wave ve- locity (vsh ≈ 3000 km s−1), after correction for the CS ve- locity, provides an estimate of the density of the CDS mate- rial in the broad-line region (ncds). The pressure equilibrium condition, with sound speed in the CDS matter of cs ≈ 15 km s−1, results in ncds ≈ n(vsh/cs)2 ≈ 4 × 1013 cm−3. This tremendous density provides a natural reason for the strong collisional thermalization of the broad component indicated by the observed inverse Balmer decrement (Section 3.3.1). 4.3.4 Circumstellar density from narrow lines The strength of narrow subordinate lines indicates a high density in the CS envelope, as noted by SCL98. The sim- plest density estimate may be taken from the condition that the optical depth in a line of Fe ii, for example λ5018, is on the order of unity. This follows from the relative intensity of the absorption component of this line (∼ 0.5). Assuming that level populations of Fe ii obey the Boltzmann distribu- tion for T = 104 K, one obtains for solar Fe/H a hydrogen concentration of n ≈ 3 × 107v8r−1 cm−3. Here r15 is the linear scale of the line-forming zone in units of 1015 cm, v8 is the velocity dispersion in units of 108 cm s−1, and f2 is the ionization fraction of Fe ii. Given the scale of the CS en- velope (∼ 4 × 1015 cm), we thus have a lower limit n > 107 cm−3. This lower limit should be considered as a revised version of the estimate reported in SCL98. 15 f −1 2 A somewhat more elaborate estimate of the density may be obtained using several different lines and assuming a quasi-local thermodynamic equilibrium (quasi-LTE) ap- proximation, i.e., the Saha-Boltzmann equations corrected for both geometrical dilution W and dilution of the pho- tospheric brightness ξ. The latter defines the photospheric brightness through the black-body brightness Iν = ξBν (T ). We adopt a simple density distribution in the CS envelope: a plateau with a steep outer drop n = n0/[1 + (r/Rk)12] , (6) where the cutoff radius Rk ≈ Rcs = 4 × 1015 cm. We consider two cases for the CS kinematics: free expansion (v = hr), with outer velocity at r = 6 × 1015 cm equal to vb = 1100 km s−1, and constant-velocity flow with v = 1100 km s−1. The photospheric radius on day 79 is estimated as Rp ≈ vst ≈ 2.7 × 1015 cm, where vs = 4000 km s−1, while the temperature is taken to be Tp = 7200 K according to the value found from the black-body fit to the continuum (Section 3.2.3). The parameter ξ in this case is found to be 0.26. In Fig. 6 we show calculated profiles of Fe ii λ5018, Si ii λ6148, the Na i D1,2 doublet and the O i 7773 A triplet, for solar abundance and hydrogen concentration n0 = 1.5 × 109 cm−3. In the case of Na i we have for the sake of simplicity ignored the doublet structure, which explains why our model line is narrower than the observed one. This density seems to be the optimal one: a factor of 1.5 lower density results in too weak O i and Na i lines, while a factor 1.5 higher density makes all the lines too strong. Although neither kinematic model produces an excel- lent fit, it is clear that free expansion kinematics better pre- dicts the positions of the absorption minima and, therefore, is preferred compared to the constant-velocity case. The dif- ferences between the modelled and observed line profiles in the free-expansion case may be related to the omission of Thomson scattering and the simplicity of our quasi-LTE model. 4.4 Overview of the qualitative model We summarize the main results of the qualitative analysis in a cartoon (Fig. 7). It shows the basic structural elements of SN 1994W which we believe are responsible for the for- mation of the optical spectrum when the supernova's lumi- nosity is high. The ejecta expand with a velocity of ∼ 4000 km s−1 into an extended (∼ 4 × 1014 cm) CS envelope. The characteristic density of the CS envelope is ∼ 109 cm−3. The SN ejecta are enshrouded by the opaque CDS within which the photosphere resides during most of the luminous phase. The opaque CDS precludes the formation of absorption lines from the ejecta. The CDS presumably forms primarily dur- ing shock break-out and is subsequently maintained by both CS material swept up by the radiative forward shock and SN material swept up by the reverse shock. These shocks are not shown in Fig. 7 but they are presumably located at the dis- tances ∆R < 0.1R away from the outer and inner edges of the CDS. Attached to the CDS is a layer populated by dense (n ≈ 4 × 1013 cm−3) fragments supplied by the RT insta- bility of the dense CDS and, possibly, by radiative shocks in CS clouds. This inhomogeneous layer of dense material is the primary site of the broad emission lines with the charac- teristic velocity of the line-emitting gas, 4000 km s−1 (e.g., in He i). Circumstellar envelope ejection in SN 1994W 13 the SN ejecta begin to interact with the extended stellar envelope. The density of the extended stellar envelope is ad- justed in our model by the requirement that a strong initial deceleration of the swept-up thin shell must result in an expansion velocity of ∼ 4000 km s−1 on day 30 when the photospheric radius reaches ∼ 1015 cm. The CS envelope is attached to the stellar envelope at r ≈ 1015 cm. A constant pre-shock velocity (1000 km s−1) is assumed for the CS gas. This is a reasonable approximation for the homologous ex- pansion, bearing in mind additional radiative acceleration. For the SN mass and energy we adopt M = 8 M⊙ and E = 1051 erg -- both values are close to the parameters of the light-curve models B and C of Falk & Arnett (1977). We find that the dynamical model is nearly the same for SN mass in the range 7−9 M⊙. The evolution of the CDS radius and velocity is shown in Fig. 8 for the density profile which is also consistent with the Hα model. In fact, the optimal den- sity is a compromise between the requirements imposed by the Hα profile evolution, the CDS velocity, and the bolomet- ric luminosity at the final stage of the light-curve plateau. The density of the CS envelope is nearly flat (ρ ∝ r−0.4) in the range 1015 < r < 4 × 1015 cm. The ionization of the CS envelope is produced by both X-ray emission from the radiative forward shock wave and by photospheric radiation. The latter primarily operates via hydrogen photoionization from the second level. The X-ray luminosity of the forward shock wave is calculated as Lx = 2πr2ρ(vs − ups)3, (7) where ρ is the pre-shock density, while the pre-shock ve- locity ups is the superposition of the pre-explosion velocity and the accelerated velocity term. Half of this luminosity is directed toward the CDS where it is reprocessed into con- tinuum radiation. This component is not treated explicitly in the model. The other half of the X-ray luminosity of the forward shock is emitted outward and is partially ab- sorbed by the CS gas. The X-ray spectrum of the forward shock is approximated as FE = CE−q exp (−E/Ts), where Ts is the temperature of the forward shock defined by the shock velocity vs − ups. The spectral power index q = 0.5 roughly describes both the Gaunt factor and the contribu- tion of X-ray lines in the low-energy band (Terlevich et al. 1992). The soft X-rays from the reverse shock at the epoch under consideration are fully absorbed by the CDS (cf. Fig. 9 of Fransson, Lundqvist & Chevalier 1996) and by the SN ejecta, and are reprocessed into the optical. The energy of the X-rays absorbed in the CS envelope is shared between heating, ionization and excitation. Ex- citation and ionization by the photospheric radiation are also taken into account. The electron temperature Te in the CS envelope is determined from the energy balance between heating from ionization by X-rays and photospheric radi- ation and cooling due to hydrogen. This approximation is quite reasonable in the temperature range of (1 − 2) × 104 K. The hydrogen atom is treated in the two-level-plus- continuum approximation. The degree of ionization we ob- tain for hydrogen in the CS envelope lies typically in the range 0.8 − 0.95. The Hα emissivity is due to recombination (Case B) and collisional excitation. Collisional de-excitation of Hα is included in the computation of the net emissivity. To allow for uncertainties in the flux, model, distance and reddening we use a fitting parameter Cn for narrow Hα. This (photosphere) Figure 7. A sketch of SN 1994W at the epoch around day 30. We show the principal structural elements involved in the formation of the spectrum. The SN ejecta are bounded by an opaque cool dense shell (CDS), which is responsible for the continuum radi- ation. The broad-line region is a narrow mixing layer attached to the CDS, composed of Rayleigh-Taylor fragments of the CDS matter and possibly of shocked CS clouds. The SN ejecta expand into a dense CS envelope with Thomson optical depth on the or- der of unity. The CS envelope is responsible for both narrow lines and the extended Thomson wings seen in Hα. The CS envelope, which in turn expands with a veloc- ity of ∼ 1000 km s−1, is responsible for the narrow lines. Thomson scattering in the CS envelope results in the emer- gence of broad emission-line wings, which are most apparent in the strong lines such as Hα and Hβ. Because of the large expansion velocity, the red wing is stronger than the blue wing. 5 A MODEL FOR Hα In order to confirm and refine the above qualitative picture of the formation of the SN 1994W spectrum, we have mod- elled the Hα line profile. In many respects our approach repeats the analysis of SN 1998S by Chugai (2001), which involved calculations of the CDS dynamics and Monte Carlo computations of the emergent lines formed outside the CDS. Here we use an updated version of this model, which now in- cludes calculations of CS hydrogen ionization, electron tem- perature and Hα emissivity. In the previous model both of the latter were set rather arbitrarily. 5.1 The model The CDS dynamics are calculated numerically in the thin- shell approximation (Chevalier 1982). We assume that the SN initially expands homologously (v ∝ r) with density dis- tribution ρ = ρ0/[1 + (v/vk)9], where the parameters of the distribution are defined by the ejecta mass M and kinetic en- ergy E. At radius R0 = 1000 R⊙ (a rather arbitrary value) 14 Nikolai N. Chugai, Sergei I. Blinnikov, Robert J. Cumming et al. Figure 9. A typical model of Hα on day 31 (thick line), showing the contribution of the narrow component (thin line). Figure 8. The CS density (top panel) and evolution of the radius and velocity of the cool dense shell (lower panels). parameter is a multiplicative factor in the Hα emissivity in the CS envelope. The broad-line region (Fig. 7) is described by an (on average) homogeneous layer, Rs < r < R1, where R1/Rs = 1.15 and the distribution of the line luminosity dL/dv ∝ v for v1 < v < vs, where v1 is the minimal velocity adopted to be equal to the pre-shock CS gas velocity. We express the average emissivity of the broad component in terms of the pre-shock emissivity, multiplied by a fitting factor Cb, which is restricted by the requirement that the Hα intensity from the broad-line region cannot exceed the black-body value for the local electron temperature of the CS gas in the pre-shock zone. We assume free expansion pre-explosion kinematics (u = hr), though we have also explored the constant-velocity case. The acceleration by the SN radiation is described as an additional CS velocity term ua = u0(Rs/r)2, where u0 is a free parameter. 5.2 Thomson scattering and the broad component In the Hα line profile, the broad emission component and electron-scattering wings overlap and at first sight cannot be disentangled unambiguously. Nevertheless, we found that there is not much freedom in the decomposition procedure. In Fig. 9 we show a model narrow-line component on day 31, with the broad component turned off. To emphasize the dif- ferent behaviours of the broad component and the electron- scattering wings, we show in Fig. 10 models with different Thomson optical depth (0, 1 and 2.8) but with otherwise similar parameters. For optical depth zero, the modelled broad component is strongly skewed toward the blue be- cause of occultation by the photosphere. This asymmetry decreases as τT increases and Thomson scattering of the Figure 10. The same model as in Fig. 9 but for different Thom- son optical depths. Lines with increasing thickness show cases τT = 0, 1 and 2.8, respectively. line emission in the expanding CS envelope produces an in- creasingly strong red wing. These different behaviours of the broad line and electron-scattering wings permit us to disen- tangle the contribution of the two line components in Hα. In passing, we note that the continuum level also depends on τT, since the Thomson optical depth affects the escape probability of photons emitted by the photosphere. 5.3 Hα and the expansion law The sensitivity of the model line profile to the kinematics of the CS envelope is demonstrated in Fig. 11. The plot shows Hα profiles on day 31 computed for similar model parameters -- all that differs is the velocity distribution. For τT = 2.8, homologous expansion without post-explosion radiative acceleration fits the observations quite well. The Circumstellar envelope ejection in SN 1994W 15 Figure 11. Hα profile on day 31 for different velocity distri- butions (shown in the inset in each panel). Overplotted on each observed profile (thin line) are models for CS envelope kinematics displayed in insets. constant-velocity case is much less successful: it produces an emission component which is too broad, and absorption which is too shallow (Fig. 11). This result, taken together with the results of the narrow CS line modelling (Fig. 6), ar- gues against the constant-velocity case. We checked a model with constant velocity and pre-shock acceleration and found, unsurprisingly, that this gives even worse agreement. The homologous model, on the other hand, works just as well when combined with post-explosion acceleration. In the bot- tom panel of Fig. 11 we show such a model profile char- acterized by an amplitude of accelerated velocity term of ua = 400 km s−1. Although homologous expansion kinematics is certainly preferred, this model shows a rather deep absorption compo- nent. This contradiction is not related to the finite resolution since we have smoothed the model profile by a Gaussian with appropriate width. An explanation for the large absorption strength in the model might be hidden in a possible devi- ation of the kinematics from homologous expansion, some clumpiness of the CS envelope, or lower hydrogen excita- tion in the outer layers of the CS envelope than the model predicts. 5.4 Modelling the Hα evolution We have calculated Hα for three epochs (31, 57 and 89 days past explosion) using the density distribution of the CS en- velope (r > 1015 cm), and the CDS radius and velocity as shown in Fig. 8. The density parameter w = 4πr2ρ at r = 1015 cm on day 31 is w = 7.5 × 1016 g cm−1. Free ex- pansion kinematics is assumed with a boundary velocity of 1100 km s−1 at r = 5.4 × 1015 cm on day 31. This implies an age for the CS envelope of tcs = 1.5 yr when the SN Figure 12. Hα at different epochs. Model profiles (thick line) are overlaid on the observations (thin line). explodes. For other values of tcs counted from the moment of envelope ejection, we rescale according to r ∝ tcs and ρ ∝ 1/t3 cs. The computed profiles are shown in Fig. 12. In Table 4 we list the corresponding parameters: photospheric (CDS) radius, photospheric temperature, brightness dilution, CDS velocity, post-explosion velocity increase, and emissivity fit- ting factors for narrow and broad components. The last col- umn displays the calculated Thomson optical depth outside the CDS. On days 31 and 57 we use the approximation of an opaque photosphere. On day 89 we found, however, that this approximation predicts too weak a red part of the profile. The model agrees better with the observations if we suggest that at this late phase the CDS is semi-transparent. To de- scribe this effect in a simple way we allow a photon striking the photosphere to cross it and escape into the CS medium with a finite probability, which we found should be close to p = 0.3. This assumption is qualitatively consistent with the fact that the photospheric brightness on day 89 is diluted (ξ = 0.2) for the optimal continuum temperature. The mod- els for all three epochs demonstrate satisfactory fits to the data. On day 31, the model is consistent with the presence of broad He i λ6678 in the red wing of Hα, while on day 57 the model fit in the red wing does not leave any room for the He i line. This behaviour is consistent with the weakening of other He i lines at this time. We calculated the Hα component from the CS envelope in a unified way with a minimum number of free parameters. For this reason, the fact that the tuning parameter of the narrow-line intensity (Cn) is close to unity (Table 4) means that within the uncertainties (∼ 20 per cent) the electron distribution in the CS envelope recovered from the Thomson scattering effects is also consistent with the luminosity of the narrow Hα component. Because the line luminosity depends on the filling factor as L ∝ 1/f while the Thomson optical depth does not, we conclude that the filling factor in the CS 16 Nikolai N. Chugai, Sergei I. Blinnikov, Robert J. Cumming et al. envelope is close to unity in at least most of the line-forming zone. This last observation is of crucial importance for distin- guishing between the two scenarios for the formation of the broad-line region (Section 4.2), and indicates that CS clouds do not play a significant role in this process. We have studied cases with different density power-law indices s (defined via ρ ∝ rs) in the flat part of the CS envelope. In the case s = 0, the tuning parameter (Cn) must systematically decrease by a factor of 1.25 to fit the profile. For s = −1, Cn should systematically increase with time by a similar factor. Bearing in mind the uncertainties in the modelling and fluxing, we do not rule out that the density gradient lies in the range −1 6 s 6 0. 5.5 Implications for the origin of the circumstellar envelope The model for the Hα evolution presented above for the CS envelope outside r ≈ 1015 cm argues for a total mass of Mcs ≈ 0.4 M⊙ and kinetic energy of Ecs ≈ 2 × 1048 erg. Combined with the derived kinematic age of the CS envelope tcs ≈ 1.5 yr, the average kinetic luminosity of the mass-loss mechanism responsible for the ejection of the CS envelope then becomes L ≈ Ecs/tcs ≈ 4 × 1040 erg s−1, and the average mass-loss rate is ∼ 0.3 M⊙ yr−1. The kinetic luminosity we derive here is enormous com- pared to stellar values. It exceeds by two orders of mag- nitude the radiative luminosity of a pre-supernova with a main-sequence mass of 6 20 M⊙. This certainly rules out a superwind as the mass-loss mechanism responsible for the CS envelope around SN 1994W. The CS envelope around SN 1994W must have been born as a result of a rather vio- lent mass ejection initiated by some energetic explosive event in the stellar interior (E > 2 × 1048 erg) at ∼ 1.5 yr prior to the SN outburst. 6 LIGHT-CURVE MODELLING To model the broad-band photometric light curves, we use the multi-energy group radiation hydrodynamics code stella (Blinnikov et al. 1998, 2000). In the current work, stella solves time-dependent equations for the angular mo- ments of intensity averaged over fixed frequency bands, us- ing 200 zones for the Lagrangian coordinate and up to 100 frequency bins with variable Eddington factors. The transfer of gamma rays from radioactive decay is calculated using a one-group approximation for the non-local deposition of the energy of radioactive nuclei. Except for the latest phases, the gamma-ray deposition is not important for SN 1994W. In the equation of state, LTE ionizations and recombina- tions are taken into account, but radiation is not assumed to be in equilibrium with matter. The effect of line opac- ity is treated as an expansion opacity according to the pre- scription of Eastman & Pinto (1993). In comparison with previous work with stella, here we use an extended spec- tral line list, the same as in the discussion of SN 1998S by Chugai et al. (2002). Figure 13. Composition as a function of interior mass, Mr, for the most abundant elements in the pre-supernova model sn94w58. The mass cut is at Mc = 1.41 M⊙. 6.1 Pre-supernova models SCL98 used analytical fits (Litvinova & Nadyozhin 1983, 1985; Popov 1993) to estimate the pre-SN radius, mass and explosion energy in SN 1994W. However, those fits are only valid for typical SNe II-P, and from the discussion above we know that SN 1994W was instead dominated by circumstel- lar interaction. To determine the pre-supernova parameters for such an unusual event, we therefore need to do detailed numerical modelling. This can be done only for a subset of the pre-supernova parameters, so we used a cut-and-trial ap- proach, guessing the initial conditions and computing light curves in order to meet the constraints posed by the photom- etry and the spectral analysis of Section 5. In the absence of an evolutionary model, we built a sequence of several dozens of non-evolutionary models, converging finally on a reasonably good fit to observations. Our initial models are constructed in hydrostatic equilibrium for the bulk mass, while the outer layers mimic the structure of the circumstel- lar envelope. For a given mass M and radius R0, we obtain a model in mechanical equilibrium assuming a power-law dependence of temperature on density (Nadyozhin & Razinkova 1986; Blinnikov & Bartunov 1993): T ∝ ρα . (8) The hydrostatic configuration thus obtained would be close to a polytrope of index 1/α if it were chemically homoge- neous and fully ionized. The difference from a polytropic model arises due to recombination of ions in the outermost layers and non-uniform composition (Fig. 13). In the centre of this configuration, at the mass cut of the collapsing core, we assume a "point-like" gravitating hard core (with numerical radius Rc = 0.1R⊙ -- much larger than a real core, but much smaller than the radii of mesh zones involved in our hydrodynamic simulations). The den- Table 4. Parameters of the Hα evolution models Circumstellar envelope ejection in SN 1994W 17 Day Rp (1015 cm) T (K) ξ vs ua (km s−1) (km s−1) Cb Cn τT 31 57 89 1.2 2.2 3.3 16000 7760 7000 1 1 0.2 4220 4160 3920 400 400 170 35 1 0.5 1.15 1 1.15 2.8 2.0 0.54 Figure 14. Density as a function of the radius r in the pre- supernova models sn94w43 and sn94w58. Figure 15. Density as a function of the interior mass (Mr) in the pre-supernova models sn94w43 and sn94w58. The mass cut is at Mc = 1.41 M⊙. sity structure found in this way is shown in Figs. 14 and 15. The parameters of the models are given in Table 6.1. The first column is the model label, which is followed by the mass of SN ejecta, pre-supernova stellar radius, 56Ni mass, power-law index α, density of the CS envelope (ρ15) at the radius r = 1015 cm, power-law index (p) of the CS envelope density distribution ρ ∝ r−p, the outer radius of the CS envelope, and the kinetic energy at infinity in units of foe (1 foe = 1051 erg). Each model was exploded by the deposition of heat en- ergy in a layer of mass ∼ 0.06 M⊙ outside of 1.41 M⊙. Since stella does not include nuclear burning, preservation of the same mixed composition in the ejecta is assured. We explored parameter space for the mass (M ) and en- ergy (E) of the SN ejecta and found acceptable fits to the data for masses in the range 6 -- 15 M⊙ and with the en- ergy given by the ratio E/M ≈ 0.15 -- 0.2 foe M −1 ⊙ . In the following sections and in Table 6.1, we present three rep- resentative models which allow us to demonstrate different aspects of the formation of the light curve. Of these three, model sn94w58 provides our best fit to SN 1994W, model sn94w43 shows that a normal SN II-P light curve is not appropriate for SN 1994W, and model sn94w68 illustrates what the supernova light curve might have looked like with- out CS interaction. All three models have E = 1.5 × 1051 ergs. This energy is typical for good light-curve models. The asymptotic kinetic energy of the ejecta is somewhat lower and is given in the Table 6.1. 6.2 Hydrodynamics and shock propagation The modelled light curves are dominated by the diffusion of the trapped radiation generated during the shock wave's propagation in the extended stellar atmosphere and sub- sequent emission of a radiative shock propagating in the dense circumstellar medium. A dense shell is formed which is found in non-equilibrium radiation hydrodynamic mod- elling but often missed in equilibrium diffusion modelling. We identify this dense shell as the CDS discussed in Sec- tion 4.1. The Lagrangian code stella does a good job of resolving the very fine structure of the opaque shell, which contains approximately one solar mass of the material (Fig. 16). 6.3 Light curves In Figs. 17 and 18 we show the changes in the colour tem- perature, Tc, of the best black-body fit to the flux. We compare this to the effective temperature, Teff , defined by the luminosity and the radius of last scattering R through L = 4πσT 4 eff R2 (see Blinnikov et al. 1998 for details of find- ing R and from that Teff ). Our multi-group radiative trans- 18 Nikolai N. Chugai, Sergei I. Blinnikov, Robert J. Cumming et al. Table 5. Parameters of hydrodynamical models Model a Mej (M⊙) R0 (104 R⊙) sn94w43 sn94w58 sn94w64 12 7 7 8 2 2 MNi (M⊙) 0.015 0.015 0.015 α ρ15 (10−15 g cm−3) 0.319 0.31 0.31 5 12 0 p 2 1 - Rw (104 R⊙) 12 6.6 - a pre-supernova mass = Mej + 1.41 M⊙ b kinetic energy at infinity b Ekin (1051 erg) 1.07 1.25 1.33 Figure 16. Density as a function of radius (r) in two models with different amounts of artificial smearing at day 30. The height of the density spike with small smearing (solid line) reaches the theoretical estimate for an isothermal shock wave. fer with hydrodynamics obtains this temperature in a self- consistent way, and no additional estimates of the thermal- ization depth are needed (in contrast to the one-group model of Ensman & Burrows 1993, for example). The large differ- ence between colour and effective temperatures is partly due to a geometric effect (the radius of last scattering is greater than the effective radius of photon creation) and partly due to the blanketing effect of scattering (the average energy of the photons is higher than that corresponding to the value of Teff ; Schuster 1905). The modelled Tc changes in a way sim- ilar to the values derived from black-body fits to the spectra (Section 3.2.3). The light curves are shown in Figs. 19, 20 and 21. The model sn94w43 produces a bright light curve, but the sec- ond half of its plateau is dominated by diffusion, not by the shock, so the behaviour of colours is similar to that of a typical SN II-P. The density of the more powerful wind in the model sn94w58 follows the law ρ ∝ r−1. This model produces a much better fit to the colours at the late plateau phase. The flat CS density distribution is consistent with the modelling Figure 17. Colour and effective temperatures for the model sn94w43. Realistic, scattering-dominated opacity has been as- sumed. The solid line shows the temperature of the best black- body fit to the flux (colour temperature). The dashed line shows the effective temperature defined by the luminosity and the radius of last scattering. of Hα evolution (Section 5.4). Model sn94w64, which is ex- actly the same as model sn94w58 in the bulk mass, but does not have the powerful wind, is much less luminous at the late plateau phase. This shows the importance of CS interaction for the luminosity. The time-dependence of the radius Rs and velocity Vs of the cool dense shell in model sn94w58 is shown in Fig. 22. The two nearly indistinguishable lines in the Vs plot are obtained by taking Vs = dRs/dt numerically, and by taking the mass-averaged speed of matter inside the shell. A com- parison with Fig. 8 shows good agreement with the spectral model, although Vs is somewhat higher in the hydrodynam- ical model. The difference is at most ∼ 10 per cent. To summarize, the hydrodynamical modelling of the light curve of SN 1994W suggests that the optimal light curve is produced in a model which contains a dense CS en- velope at r > ∼ 1015 cm with a relatively flat density distribu- tion and an outer radius of ∼ 4.5 × 1015 cm. This conclusion is fully consistent with the Hα modelling. Circumstellar envelope ejection in SN 1994W 19 Figure 18. Colour and effective temperatures for the model sn94w58. Figure 20. Same as in Fig. 19, but for model sn94w58. Figure 19. Light curves for the model sn94w43. The photometric points are taken from SCL98 and references therein. 7 DISCUSSION AND CONCLUSIONS We have presented and analysed spectra and light curves of SN 1994W, one of the best-observed SNe IIn. During the first three months the spectrum clearly shows the presence of three line components: (i) narrow P-Cygni lines, (ii) in- trinsically broad lines and (iii) extended smooth wings in Hα and Hβ. We attribute these components, respectively, to (i) a dense CS envelope, (ii) shocked, cool, dense gas confined in a narrow layer on top of the photosphere, and (iii) the effects of Thomson scattering in the CS envelope. Our line profile analysis and hydrodynamical light- curve modelling have led us to a coherent picture. SN 1994W Figure 21. Same as in Fig. 19, but for model sn94w64. appears to have been the result of explosion of an extended pre-supernova (∼ 1015 cm) embedded in an extended CS envelope [∼ (4 − 5) × 1015 cm] with Thomson optical depth ∼ 2.8 one month after the explosion. Although we do not rule out the presence of a wind from a normal red supergiant pre-supernova outside the dense CS envelope, such a wind component cannot have been very dense, considering the low luminosity of Hα on day 203 (L ≈ 1038 erg s−1). This luminosity is two orders of mag- nitude lower than for SN 1979C (∼ 1040 erg s−1) at ∼ 1 yr (Chevalier & Fransson 1985). Attributing this difference to the difference in the interaction luminosity (L ∝ 4πρv3 s ), we can estimate the density parameter of the outer wind in SN 1994W. The expansion velocity of the CDS in SN 1994W, ∼ 4000 km s−1, is half that of SN 1979C (∼ 8000 km s−1). 20 Nikolai N. Chugai, Sergei I. Blinnikov, Robert J. Cumming et al. ∼ 11 M⊙ star may produce strong Ne flashes several years prior to the SN explosion and that the strongest flash could eject most of the hydrogen envelope with velocities of ∼ 100 km s−1. Although such flashes do not occur in more recent models with finer zoning (Woosley, Heger & Weaver 2002), we believe that the effect warrants further investigation, es- pecially given the highly complicated nuclear burning regime and hydrodynamics of degenerate O/Ne/Mg cores. Remarkably, the low 56Ni mass (< 0.015 M⊙) estimated from the tail R-band luminosity of SN 1994W (SCL98) seems to be in accord with the suggested mass of such a progenitor, if at the lower end of our estimated mass range. Such stars are expected to eject only small amounts of 56Ni (Mayle & Wilson 1988). Our models are, however, not very sensitive to the zero-age main-sequence mass of the progen- itor, and, as SCL98 point out, the low nickel content could indicate either a low-mass progenitor which may have lost a few solar masses in shell and wind ejections, or a more mas- sive star that lost several solar masses prior to the ejection of the CS shell. Deep X-ray searches for more extended CS ma- terial could help solve this problem, as should more detailed models for stellar evolution just prior to core collapse. How common are events like SN 1994W? Of the SNe IIn that have been observed so far, it is possible that a close counterpart has been seen but not recognised as such due to sparser temporal coverage or greater distance. Never- theless, several objects have shown narrow hydrogen and Fe ii absorption lines with velocities of about 1000 km s−1. This family includes SNe 1987B (Schlegel et al. 1996), 1994aj (Benetti et al. 1998), 1994ak (Filippenko 1997), 1995G (Pastorello et al. 2002), 1996L (Benetti et al. 1999) and 1999el (Di Carlo et al. 2002). We suggest that in at least some of these SNe, a CS envelope was ejected in a violent, explosive manner, as we believe was the case for SN 1994W, and as Chugai & Danziger (2003) have suggested for SN 1995G. Further data and analyses are needed, however, be- fore it can be demonstrated that the energy of the ejected CS envelope is comparable for these events. In any case, we suggest that at least part of the variety of SNe IIn has been accounted for. Some, like SN 1988Z and SN 1998S, have a dense, slow CS envelope formed by a slow superwind. Others, like SN 1994W, seem to have ejected a CS envelope in a violent event a few years before the SN explosion. ACKNOWLEDGEMENTS We thank Aaron J. Barth, Mike Breare, Ren´e Rutten, Luis C. Ho, Neil O'Mahony, Chien Y. Peng and Ed Zuiderwijk for helping take observations for us, and Clive Jackman, Danny Lennon and Marco Azzaro for help assessing the day 31 spectrum calibration. We also thank Itziar Aretxaga, Eddie Baron, Claes Fransson, Seppo Mattila, Peter Meikle, Miguel P´erez Torres, and Luca Zampieri for discussions. This paper is based on observations made with the Isaac Newton Telescope (INT), the William Herschel Telescope (WHT), and the Nordic Optical Telescope (NOT). The INT and WHT are operated on the island of La Palma by the Isaac Newton Group in the Spanish Observatorio del Roque de los Muchachos of the Instituto de Astrof´ısica de Canarias. The NOT is operated on the island of La Palma jointly Figure 22. The evolution of the radius and velocity of the cool dense shell in hydrodynamical model sn94w58. Compare with Fig. 8. Given the SN 1979C wind parameter, w ≈ 1016 g cm−1, the outer wind in SN 1994W should have w ≈ 1015 g cm−1 to account for the difference in the Hα luminosity. This is at least ten times lower than in SN 1979C, and a factor of two lower than in SN 1980K (e.g., Lundqvist & Fransson 1988). In this respect, the detection by Schlegel (1999) of X-rays from SN 1994W at t = 1180 days with luminosity L ≈ 8 × 1039 erg s−1 appears odd when compared with the expected X-ray luminosity ( < ∼ 1038 erg s−1) at that epoch for a case of SN/wind interaction with w ≈ 1015 g cm−1. To resolve this problem, one would need the wind density to be at least of factor of ten higher at ∼ 4 × 1016 cm, i.e., exceeding that of SN 1979C. This could argue for a possible multi-shell ejection scenario prior to the explosion. The recovered kinematics, density and linear scale of the CS matter around SN 1994W imply a kinematic age for the CS envelope of tcs ≈ 1.5 yr, mass Mcs ≈ 0.4 M⊙ and kinetic energy Ecs ≈ 2 × 1048 erg. The enormous average mass-loss rate (Mcs/tcs ≈ 0.3 M⊙ yr−1) and equally enormous kinetic luminosity (Ecs/tcs ≈ 4 × 1040 erg s−1) of the mass-ejection mechanism strongly suggest that the CS envelope has been lost as a result of an explosive event which occurred ∼ 1.5 yr prior to the SN outburst. In their model for the narrow lines in SN 1983K, Grasberg & Nadyozhin (1986) invoked the violent ejection of the hydrogen envelope 1 -- 2 months before the SN explo- sion with a velocity on the order of ∼ 2500 km s−1 and an energy (1−2)×1049 erg. Dramatically high mass loss shortly before explosion has also been derived for the SN IIn 1995G by Chugai & Danziger (2003), on the basis of an analysis similar to the one we have presented here. The mechanism behind such violent ejections might be associated with a nuclear flash in the degenerate core of the pre-supernova. This conjecture, also presented by Chugai & Danziger (2003), is prompted by the prediction of Weaver & Woosley (1979) that the O/Ne/Mg core of an Circumstellar envelope ejection in SN 1994W 21 by Denmark, Finland, Iceland, Norway, and Sweden, in the Spanish Observatorio del Roque de los Muchachos of the Instituto de Astrof´ısica de Canarias. Some of the data pre- sented herein were obtained at the W.M. Keck Observatory, which is operated as a scientific partnership among the Cal- ifornia Institute of Technology, the University of California and the National Aeronautics and Space Administration. The Observatory was made possible by the generous finan- cial support of the W.M. Keck Foundation. This project was supported by the Royal Swedish Academy of Sciences. S.I.B. was supported partly by RFBR 02-02-16500, the Wenner-Gren Science Foundation, MPA Garching and ILE Osaka guest programs. The research of P.L. is sponsored by the Royal Swedish Academy and the Swedish Research Council, and he is a Research Fellow at the Royal Swedish Academy supported by a grant from the Wallenberg Foundation. A.V.F.'s research is supported by National Science Foundation grant AST-0307894. REFERENCES Allen, C. W., 1973, Astrophysical Quantities. University of London, Athlone Press, London Aretxaga, I., Benetti, S., Terlevich, R. J., Fabian, A. C., Cappellaro, E., Turatto, M., Della Valle, M., 1999, MN- RAS, 309, 343 Axelrod, T., 1980, Ph.D. thesis, California Univ., Santa Cruz Cumming, R. J., Lundqvist, P., Meikle, W. P. S., 1994, IAU Circ., No. 6057 Di Carlo, E., et al., 2002, ApJ, 573, 144 Dopita, M. A., Evans, R., Cohen, M., Schwarz, R. D., 1984, ApJ, 287, L69 Eastman, R. G., Pinto, P. A., 1993, ApJ, 412, 731 Elitzur, M., Ferland, G. J., Mathews, S. G., Shields, G. A., 1983, ApJ, 272, L55 Ensman, L., Burrows, A., 1992, ApJ, 393, 742 Fassia, A., et al., 2001, MNRAS, 325, 907 Falk, S. W., Arnett W. D., 1977, ApJS, 13, 515 Filippenko, A. V., 1982, PASP, 94, 715 Filippenko, A. V., 1989, AJ, 97, 726 Filippenko, A. V., 1991, in Danziger, I. J., Kjar, K., eds, Supernova 1987A and Other Supernovae ESO, Garching, p. 343 Filippenko, A. V., 1997, ARA&A, 35, 309 Filippenko, A. V., 2000, in Holt, S. S., Zhang, W. W., eds, Cosmic Explosions, AIP, New York, p. 123 Filippenko, A. V., Barth, A. J., 1994, IAU Circ., No. 6046 Filippenko, A. V., Moran, E. C., 1998, IAU Circ., No. 6830 Fransson, C., 1982, A&A, 111, 140 Fransson, C., 1984, A&A, 133, 264 Fransson, C., Lundqvist, P., Chevalier, R. A., 1996, ApJ, 461, 993 Fransson, C., et al., 2002, ApJ, 572, 350 Fransson, C., et al., 2004, ApJ, submitted Fuhr, J. R., Martin, G. A., Wiese, W. L., 1988, J. Phys. Chem. Ref. Data 17, Suppl. 4 Almog, Y., Netzer, H., 1989, ApJ, 238, 57 Benetti, S., Cappellaro, E., Danziger, I. J., Turatto, M., Gaskell, C. M., 1994, IAU Circ., No. 6048 Garc´ıa-Lario, P., Riera, A., Manchado, A., 1999, ApJ, 526, Patat, F., Della Valle, M., 1998, MNRAS, 294, 448 854 Benetti, S., Turatto, M., Cappellaro, E., Danziger, I. J., Germany, L., Reiss, D. J., Sadler, E. M., Schmidt, B. P., Mazzali, P. A., 1999, MNRAS, 305, 811 Stubbs, C. W., 2000, ApJ, 533, 320 Bessell, M. S., 1999, PASP, 111, 1426 Blinnikov, S. I., Bartunov, O. S., 1993, A&A, 273, 106 Blinnikov, S. I., Eastman, R., Bartunov, O. S., Popolitov, Grasberg, E. K., Imshennik, V. S., Nadyozhin, D. K., 1971, Ap&SS, 10, 28 Grasberg, E. K., Nadyozhin, D. K., 1986, Sov. Astr. Lett., V. A., Woosley, S. E., 1998, ApJ, 496, 454 12, 68 Blinnikov, S., Lundqvist, P., Bartunov, O., Nomoto, K., Iwamoto, K., 2000, ApJ, 532, 1132 Blondin, J. M., Ellison, D. C., 2001, ApJ, 560, 244 Bowen, D. V., Roth, K. C., Meyer, D. M., Blades, J. C., 2000, ApJ, 536, 225 Bragaglia, A., Munari, U., Barbon, R., 1994, IAU Circ., No. 6044 Cardelli, J. A., Clayton, G. C., Mathis, J. S., 1989, ApJ, 345, 245 Chevalier R.A., 1982, ApJ, 258, 790 Chevalier R.A., Blondin J.M., 1995, ApJ, 444, 312 Chevalier R. A., Fransson C., 1985, in Bartel N., ed., Su- pernovae as Distance Indicators, Springer-Verlag, Berlin, p. 123 Chevalier, R. A., Fransson, C., 1992, ApJ, 395, 540 Chevalier, R. A., Fransson, C., 1994, ApJ, 420, 268 Chugai, N. N., 1990, Sov. Astr. Lett., 16, 457 Chugai, N. N., 1992, SvA, 36, 63 Chugai, N. N., 1997, Ap&SS, 252, 225 Chugai, N. N., 2001, MNRAS, 326, 1448 Chugai N. N., Blinnikov, S. I., Fassia, A., Lundqvist, P., Meikle, W. P. S., Sorokina, E. I., 2002, MNRAS, 330, 473. Chugai, N. N., Danziger, I. J., 2003, Astr. Lett., 29, 732 Cortini, G., Villi, M., 1994, IAU Circ., No. 6042 Gualandi, R., Merighi, R., 2001, BFOSC (Bologna Faint Object Spectrograph & Camera) Manuale Utente, version 2.0. Osservatorio Astronomico di Bologna, Bologna Hamuy, M. et al., 2003, Nat, 424, 651 Henry, R. B. C., Branch, D., 1987, PASP, 99, 112 King, D. L., 1985, ING Technical Note, 31 Klein, R. I., McKee, C. F., Colella, P., 1994, ApJ, 420, 213 Leibundgut, B. et al., 1991, ApJ, 372, 531 Leonard, D. C., Filippenko, A. V., Barth, A. J., Matheson, T., 2000, ApJ, 536, 239 Lewis, J. R., et al., 1994, MNRAS, 266, L27 Litvinova, I. Yu., Nadyozhin, D. K., 1983, Astrophys. Sp. Sci. 89, 89 Litvinova, I. Yu., Nadyozhin, D. K., 1985, Pis'ma Astron. Zh. 11, 351 (Sov. Astron. Lett. 11, 145) Lundqvist, P., Fransson, C., 1988, A&A, 192, 221 Matheson, T., Filippenko, A. V., Ho, L. C., Barth, A. J., Leonard, D. C., 2000, AJ, 120, 1499 Meikle, W. P. S., Catchpole, R. M., Cumming, R. J., Geballe, T. R., Lewis, J. R., Martin, R., Walton, N. A., 1994, Spectrum, 4, 7 Mayle, R., Wilson, J. R., 1988, ApJ, 334, 909 Miller, J. S., Stone, R. P. S., 1993, Lick Obs. Tech. Rep., No. 66 22 Nikolai N. Chugai, Sergei I. Blinnikov, Robert J. Cumming et al. Murray, C. A., 1983, Vectorial Astrometry. Adam Hilger, Bristol Nadyozhin, D. K., Razinkova, T. L., 1986, Nauchnye infor- matsii, 61, 29 Niemela, V. S., Ruiz, M. T., Phillips, M. M., 1985, ApJ, 289, 52 Oke, J. B., et al., 1995, PASP, 107, 375 Pastorello, A., Turatto, M., Benetti, S., Cappellaro, E., Danziger, I. J., Mazzali, P. A., Patat, F., Filippenko, A. V., Schlegel, D. J., Matheson, T., 2002, MNRAS, 333, 27 Pikelner, S. B., 1954, Izv. Krymsk. Astroph. Obs. 12, 93 Popov, D. V., 1993, ApJ, 414, 712 Pozzo, M., Meikle, W. P. S., Fassia, A., Geballe, T., Lundqvist, P., Chugai N. N., Sollerman, J., 2004, MN- RAS, in press (astro-ph/0404533) Rigon, L., Turatto, M., Benetti, S., Pastorello, A., Cappel- laro, E., Aretxaga, I., Vega, O., Chavushyan, V., Patat, F., Danziger, I. J., Salvo, M., 2003, MNRAS, 340, 191 Rupen, M. P., et al., 1987, AJ, 94, 61 Ryder, S., Staveley-Smith, L., Dopita, M., Petre, R., Col- bert, E., Malin, D., Schlegel, E., 1993, ApJ, 416, 167 Schlegel, E. M., 1990, MNRAS, 244, 269 Schlegel, E. M., Kirshner, R. P., Huchra, J. P., Schild, R. E., 1996, AJ, 111, 2038 Schlegel, E. M., 1999, ApJ, 527, 85 Schuster, A., 1905, ApJ, 21, 1 Shortridge, K., 1990, Starlink User Note, 86, Rutherford Appleton Laboratory Sigut, T. A. A., Pradhan, A. K., 2003, ApJS, 145, 15 Sollerman, J., Cumming, R. J., Lundqvist, P., 1998, ApJ, 493, 933 Stathakis, R. A., & Sadler, E. M., 1991, MNRAS, 250, 786 Terlevich, R. J., Tenorio-Tagle, G., Franco, J., Melnick, J., 1992, MNRAS, 255, 713 Turatto, M., Cappellaro, E., Danziger, I. J., Benetti, S., Gouiffes, C., Della Valle, M., 1993, MNRAS 262, 128 Turatto, M., Suzuki, T., Mazzali, P. A., Benetti, S., Cap- pellaro, E., Danziger, I. J., Nomoto, K., Nakamura, T., Young, T. R., Patat, F., 2000, ApJ, 534, L57 van Hoof, P., 1999, Atomic Line List v2.04, http://www.pa.uky.edu/∼peter/atomic/ Wade, R. A., Horne, K., 1988, ApJ, 324, 411 Weaver, T. A., Woosley, S. E., 1979, BAAS, 11, 724 Wegner, G., Swanson, S. R., 1996, MNRAS, 278, 22 Williams, G. A., Shipman, H. L., 1988, ApJ, 326, 738 Woosley, S. E., Heger, A., Weaver, T. A., 2002, RvMP, 74, 1015 Woosley, S. E., 1986, in Hauck B., Maeder, A., Meynet, G., eds, Nucleosynthesis and Stellar Evolution, Saas-Fee Advanced Course 16, Geneva Observatory, Sauverny, p. 1 This paper has been typeset from a TEX/ LATEX file prepared by the author.
astro-ph/9904030
2
9904
1999-07-18T02:00:34
Luminosity Profiles of Merger Remnants
[ "astro-ph" ]
Using published luminosity and molecular gas profiles of the late-stage mergers NGC 3921, NGC 7252 and Arp 220, we examine the expected luminosity profiles of the evolved merger remnants, especially in light of the massive CO complexes that are observed in their nuclei. For NGC 3921 and NGC 7252 we predict that the resulting luminosity profiles will be characterized by an r^{1/4} law. In view of previous optical work on these systems, it seems likely that they will evolve into normal ellipticals as regards their optical properties. Due to a much higher central molecular column density, Arp 220 might not evolve such a ``seamless'' light profile. We conclude that ultraluminous infrared mergers such as Arp 220 either evolve into ellipticals with anomalous luminosity profiles, or do not produce many low-mass stars out of their molecular gas complexes.
astro-ph
astro-ph
Draft version September 24, 2018 Preprint typeset using LATEX style emulateapj 9 9 9 1 l u J 8 1 2 v 0 3 0 4 0 9 9 / h p - o r t s a : v i X r a LUMINOSITY PROFILES OF MERGER REMNANTS National Radio Astronomy Observatory1, 520 Edgemont Road, Charlottesville, VA 22903-2475 J. E. Hibbard ([email protected]) and Min. S. Yun National Radio Astronomy Observatory1, P.O. Box 0, Soccoro, NM 87801-0387 ([email protected]) Draft version September 24, 2018 ABSTRACT Using published luminosity and molecular gas profiles of the late-stage mergers NGC 3921, NGC 7252 and Arp 220, we examine the expected luminosity profiles of the evolved merger remnants, especially in light of the massive CO complexes that are observed in their nuclei. For NGC 3921 and NGC 7252 we predict that the resulting luminosity profiles will be characterized by an r1/4 law. In view of previous optical work on these systems, it seems likely that they will evolve into normal ellipticals as regards their optical properties. Due to a much higher central molecular column density, Arp 220 might not evolve such a "seamless" light profile. We conclude that ultraluminous infrared mergers such as Arp 220 either evolve into ellipticals with anomalous luminosity profiles, or do not produce many low-mass stars out of their molecular gas complexes. Subject headings: galaxies: individual (NGC 3921, NGC 7252, Arp 220) -- galaxies: interactions -- galaxies: evolution -- ISM: molecules -- infrared: galaxies 1. INTRODUCTION The merger hypothesis for elliptical galaxy formation, as put forth by Toomre & Toomre (1972; see also Toomre 1977), posits that two spiral galaxies can fall together un- der their mutual gravitational attraction, eventually evolv- ing into an elliptical-like remnant. An early objection to this hypothesis was that the cores of ellipticals are too dense to result from the dissipationless merging of two spirals (Carlberg 1986, Gunn 1987, Hernquist, Spergel & Heyl 1993). An obvious solution to this objection is to include a dissipative (gaseous) component in the progeni- tors (for other solutions, see e.g. Veeraraghavan & White 1985, Barnes 1988, Lake 1989). Numerical experiments in- cluding such a component readily showed that gaseous dis- sipation can efficiently drive large amounts of material into the central regions (Negroponte & White 1983, Noguchi & Ishibashi 1986, Barnes & Hernquist 1991). Indeed, it was considered a great success for early hydrodynamical work to be able to reproduce dense knots of gas within the central regions of simulated merger remnants, similar to the gas concentrations observed in IR luminous mergers (Barnes & Hernquist 1991, Sanders et al. 1988b). Since the inferred central gas mass densities in the observed gas knots are comparable to the stellar mass density seen in the cores of normal ellipticals (∼ 102M⊙ pc−3), it seems natu- ral that they could be the seed of a high surface brightness remnant (Kormendy & Sanders 1992). Subsequent numerical work by Mihos & Hernquist (1994; hereafter MH94) finds that the dissipative response of the simulated gas component is so efficient that the resulting mass profiles of the simulated remnants are un- like those seen in normal ellipticals. In particular, ensu- ing starformation leaves behind a dense stellar core whose surface density profile does not join smoothly onto the de Vaucoleurs r1/4 profile of the pre-existing stellar pop- ulation. Instead, the profiles exhibit a "spike" at small radii, with a suggested increase in surface brightness by factors ∼100. While the predicted break in the mass den- sity profile occurs at spatial scales comparable to the grav- itational softening length of the simulations, making the precise slope somewhat questionable, the conclusion that the profiles should exhibit a clear break was considered firm (MH94). This prediction, if confirmed, offers a means to constrain the frequency of highly dissipative mergers in the past by searching for their fossil remnants in the cores of nearby ellipticals. However, the numerical formalisms used in MH94 to model gaseous dissipation, star forma- tion, and energy injection back into the ISM from mas- sive stars and SNe ("feedback") are necessarily ad hoc in nature. As such, these predictions should be viewed as preliminary (as Mihos & Hernquist themselves note). We investigate this question observationally by convert- ing the observed gas column densities of on-going or late- stage mergers into optical surface brightness by assuming a stellar mass-to-light ratio appropriate for an evolved pop- ulation. This light is added to the observed luminosity profile after allowing it to age passively. The resulting lu- minosity profile is examined for anomalous features such as the sharp break predicted by MH94. We conduct this experiment with the late-stage mergers NGC 3921 and NGC 7252 and with the ultraluminous in- frared (ULIR) merger Arp 220. These systems were chosen because they have been observed in the CO(1-0) molecu- lar line transition with resolutions (full width at half max- 1The National Radio Astronomy Observatory is a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc. 1 2 Luminosity Profiles of Merger Remnants imum) of 2′′ − 2.5′′ (Yun & Hibbard 1999a, Wang et al. 1992, Scoville et al. 1997). The resulting spatial radial resolution (300 -- 400 pc; Ho= 75 km s−1 Mpc−1) is similar to the hydrodynamical smoothing length used in MH94 (∼ 350 pc)2, indicating that the molecular line observations have sufficient resolution to resolve the types of mass con- centrations found in the simulations. The molecular gas surface densities of each of these systems are plotted in Figure 1, converted from CO fluxes by adopting a con- version factor of NH2 /ICO = 3 × 1020 cm−2 (K km s−1)−1 (Young & Scoville 1991). Detailed studies on each of these systems, which fully discuss their status as late stage merg- ers, can be found in Schweizer (1996) and Hibbard & van Gorkom (1996) for NGC 3921; Schweizer (1982) and Hib- bard et al. (1994) for NGC 7252; and Scoville et al. (1997) for Arp 220. 2. RESULTS 2.1. NGC 3921 & NGC 7252 For these moderately evolved merger remnants (ages of ∼ 0.5 -- 1 Gyr since their tidal tails were launched, Hibbard et al. 1994, Hibbard & van Gorkom 1996), the observed gas and luminosity profiles are used to predict the expected luminosity profile of a 2 Gyr old remnant. We assume that all of the molecular gas is turned into stars at the same radii, adopting an exponentially declining starformation history. The present luminosity profile is allowed to fade due to passive aging effects, and the final luminosity pro- file is the sum of these two populations. The molecular gas profiles plotted in Fig. 1 are converted into gas mass den- sities by multiplying by a factor of 1.36 to take into con- sideration the expected contribution of Helium. Optical luminosity profiles have been obtained by Schweizer (1982, 1996), Whitmore et al. (1993), and Hibbard et al. (1994), showing them to be well fitted by an r1/4 profile over all radii, with no apparent luminosity spikes. The gas surface densities (Σgas in M⊙ pc−2) are con- verted to optical surface brightnesses (µB) by dividing by the stellar mass-to-light ratio (M∗/LB) expected for a 2 Gyr old population. We adopt the stellar mass-to- light ratios given by de Jong (1995, Table 1 of ch. 4), which were derived from the population synthesis mod- els of Bruzual & Charlot (1993) for an exponentially de- clining star formation history, a Salpeter IMF, and Solar metallicity (M∗/LB = 0.82 M⊙L−1 ⊙ at 2 Gyr). Noting that 1 L⊙ pc−2 corresponds to µB = 27.06 mag arcsec−2 (adopting MB,⊙ = +5.48), the conversion from gas surface density to optical surface brightness is given by µB(r) = 27.06 mag arcsec−2 − 2.5 × log[Σgas(r)/(M∗/LB)]. The luminosity profiles of the evolved remnants are estimated from the observed B-band profiles (Schweizer 1996, Hibbard et al. 1994), allowing for a fading of +1 mag arcsec−2 in the B-band over the next 2 Gyr (Bruzual & Charlot 1993, Schweizer 1996), and adding in the expected contribution of the population formed from the molecular gas, calculated as above. We emphasize that this should favor the production of a luminous post-merger popula- tion, since it assumes the that none of the molecular gas is lost to stellar winds or SNe and the adopted IMF favors the production of many long-lived low-mass stars. The results of this exercise are plotted in Figure 2. This 2Assuming scaling parameters appropriate for a MilkyWay-like progenitor. plot shows that the observed gas densities in NGC 3921 and NGC 7252, although high, are not high enough to significantly affect the present luminosity profiles. The profile of NGC 3921 is basically indistinguishable from an r1/4 profile. The profile of NGC 7252 does show a slight rise at small radii, but not the clear break predicted by MH94. Therefore the resulting luminosity profiles of these remnants are now and should remain fairly typical of nor- mal elliptical galaxies, and the conclusions of MH94 are not applicable to all mergers of gas-rich galaxies. Since both of these systems also obey the Faber-Jackson rela- tionship (Lake & Dressler 1986) and NGC 7252 falls upon the fundamental plane defined by normal ellipticals (Hib- bard et al. 1994, Hibbard 1995), we conclude that at least some mergers of gas-rich systems can evolve into normal elliptical galaxies as far as their optical properties are con- cerned. 2.2. Arp 220 Since Arp 220 is an extremely dusty object, its opti- cal luminosity profile is poorly suited for a similar analy- sis. Instead, we use a luminosity profile measured in the near-infrared, where the dust obscuration is an order of magnitude less severe. Arp 220 was recently observed with camera 2 of NICMOS aboard the HST (Scoville et al. 1998), and we use the resulting K −band luminosity profile, kindly made available by N. Scoville. Since Arp 220 is presently undergoing a massive starburst, the fading factor is much less certain than for the already evolved sys- tems treated above, and depends sensitively on what frac- tion of the current light is contributed by recently formed stars. We adopt a situation biased towards the production of a discrepant luminosity profile by assuming that the en- tire population was pre-existing, converting the observed K −band profile to an evolved B−band profile by adopting a B −K color of 4, appropriate for a 10 Gyr old population (de Jong 1995). The contribution due to the population formed from the molecular disk is calculated exactly as before. The resulting profile is shown in Figure 2. This figure shows that Arp 220 is predicted to evolve a luminosity profile with a noticeable rise at small radii. This is due to the peak in the molecular gas surface den- sity at radii less than 0.5 kpc (Fig. 1). We conclude that Arp 220 has the potential to evolve a similar feature in its luminosity profile, if indeed all of the current molecular gas is converted into stars. However, the expected rise of ∼2 mag arcsec2 in surface brightness (a factor of ∼6) is con- siderably lower than the two orders of magnitude increase predicted by the simulations (see Fig. 1 of MH94). 3. DISCUSSION From the above exercise, we conclude that neither NGC 3921 nor NGC 7252 are expected to show a significant de- viation in their luminosity profiles, and that the maximum rise expected for Arp 220 is considerably lower than the two orders of magnitude increase predicted by the simula- tions of MH94. We conclude that the numerical formalisms adopted in the simulations to treat the gas and star forma- tion are incomplete. Mihos & Hernquist enumerate var- ious possible shortcomings of their code. For example, their star-formation criterion is extrapolated from studies Hibbard & Yun 3 of quiescent disk galaxies, and may not apply to violent starbursts. Perhaps most importantly, their simulations fail to reproduce the gas outflows seen in ULIR galaxies ("superwinds" e.g. Heckman, Armus & Miley 1990, Heck- man, Lehnert & Armus 1993), suggesting that the numer- ical treatment of feedback is inadequate. In spite of these results, it is still interesting that under some conditions there might be an observational signature of a past merging event in the light profile of the remnant. The question is for which mergers might this be the case? Since ΣH2 is tightly correlated with IR luminosity (Yun & Hibbard 1999b), we infer that only the ultraluminous IR galaxies retain the possibility to evolve into ellipticals with a central rise in their luminosity profiles. While such profiles are not typical of ellipticals in general, they are not unheard of. For example, ∼ 10% of the the Nuker sample profiles presented by Byun et al. (1996) show such anomalous cores (e.g., NGC 1331, NGC 4239). It is there- fore possible that such systems evolve from ultraluminous IR galaxies. This can be tested by careful "galactic ar- chaeology" in such systems to search for signatures of a past merger event (e.g., Schweizer & Seitzer 1992, Malin & Hadley 1997). However, it is not a foregone conclusion that systems like Arp 220 will evolve anomalous profiles. This system presently hosts a very powerful expanding "superwind" (Heckman et al. 1996), which may be able to eject a signif- icant fraction of the cold gas in a "mass-loaded flow" (e.g. Heckman et al. 1999). Such winds are common in ULIR galaxies (Heckman, Armus & Miley 1990, Heckman, Lehn- ert & Armus 1993). Another related possibility is that the IMF may be biased towards massive stars (i.e. "top heavy", Young et al. 1986, Scoville & Soifer 1991). Such mass functions will leave far fewer stellar remnants than the IMF adopted here. A third possibility is that the stan- dard Galactic CO-to-H2 conversion factor is inappropriate for ULIR galaxies, and that the high gas surface densities derived from CO observations (and thus the resulting stel- lar luminosity profile) may be over-estimated (see Downes et al. 1993, Bryant & Scoville 1996). Some support for the idea that central gas cores may be depleted by the starburst is given by a population syn- thesis model of NGC 7252, which suggests that it expe- rienced an IR luminous phase (Fritze-von Alvensleben & Gerhard 1994). While the current radial distribution of molecular gas in NGC 7252 is flat and lacks the central core seen in Arp 220, it appears to connect smoothly with that of Arp 220 in Fig. 1. Therefore one may speculate that NGC 7252 did indeed have a radial gas density pro- file much like Arp 220 but has since lost the high density gas core as a result of prodigious massive star formation and/or superwind blowout. However, the burst parame- ters are not strongly constrained by the available obser- vations, and a weaker burst spread over a longer period may also be allowed (Fritze-von Alvensleben & Gerhard 1994). Further insight into this question could be obtained by constraining the past star formation history in other evolved merger remnants. In conclusion, a comparison of the peak molecu- lar column densities and optical surface brightnesses in NGC 3921 and NGC 7252 suggests that some mergers between gas-rich disks will evolve into elliptical-like rem- nants with typical luminosity profiles, even considering their present central gas supply. For ULIR galaxies like Arp 220 the case is less clear. Such systems will either produce an excess of light at small radii, as seen in a small number of ellipticals, or require some process such as mass- loaded galactic winds or a top-heavy IMF to deplete the central gas supply without leaving too many evolved stars. If the latter possibility can be excluded, then the frequency of such profiles may be used to constrain the number of early type systems formed via ULIR mergers.3 4. SUMMARY • Even under assumptions that favor the production of a luminous post-merger population, the dense molec- ular gas complexes found in the centers of NGC 3291 and NGC 7252 should not significantly alter their luminosity profiles, which are already typical of el- liptical galaxies (Schweizer 1982, Schweizer 1996). Since these systems also obey the Faber-Jackson re- lationship (Lake & Dressler 1986) and NGC 7252 falls upon the fundamental plane defined by normal ellipticals (Hibbard et al. 1994, Hibbard 1995), it appears that at least some mergers of gas-rich sys- tems can evolve into normal elliptical galaxies as far as their optical properties are concerned. • The dense molecular gas complex found in the cen- ter of Arp 220 may result in a moderate rise in the remnants' luminosity profile at small radii. Since the molecular gas column density is a tight function of IR luminosity (Yun & Hibbard 1999b), we conclude that this condition may apply to all of the ultra- luminous infrared galaxies. However, this does not preclude a merger origin for elliptical galaxies since (1) About 10% of the Nuker sample ellipticals (Byun et al. 1996) show such rises in their radial light pro- files, and (2) it is possible that much of the gas in such systems is blown into intergalactic space by the mass-loaded superwinds found emanating from such objects (Heckman et al. 1999, Heckman, Lehnert & Armus 1993). • The maximum expected rise in the luminosity pro- files are considerably lower than the orders of magni- tude increase predicted by the simulations (MH94). We therefore suggest that the numerical formalisms adopted in the simulations to treat the gas and star formation are incomplete. The authors thank N. Scoville for kindly providing the NICMOS K-band profile for Arp 220. We thank F. Schweizer and J. van Gorkom for comments on an earlier version of this paper, and R. Bender and C. Mihos for use- ful discussions, and the referee, J. Barnes, for a thorough report. 3We note that any subsequent dissipationless merging of these cores with other stellar system will tend to smooth out these profiles. 4 Luminosity Profiles of Merger Remnants REFERENCES Barnes, J. E. 1988, ApJ, 331, 699 Barnes, J.E., & Hernquist, L. 1991, ApJ, 370, L65 Bryant, P.M. & Scoville, N.Z. 1996, ApJ, 457, 678 Bruzual, G. & Charlot, S. 1993, ApJ, 405, 538 Byun, Y.-L., Grillmair, C. J., Faber, S. M., Ajhar, E. A., Dressler, A., Kormendy, J., Lauer, T. R., Richstone, D., & Tremaine, S. 1996, AJ, 111, 1889 Carlberg, R. G. 1986, ApJ, 310, 593 Downes, D., Solomon, P. M., & Radford, S. J. E. 1993, ApJ, 414, 13 Fritze-von Alvensleben, U., Gerhard, O. E. 1994, A&A, 285, 751 & 775 Gunn, J. E. 1987, in "Nearly Normal Galaxies: From the Planck Time to the Present", edited by S. M. Faber (Springer, New York), p. 455 Heckman, T. M., Dahlem, M., Eales, S. A., Fabbiano, G., & Weaver, K. 1996, ApJ, 457, 616 Heckman, T. M., Armus, L., Weaver, K. & Wang, J. 1999, ApJ, 517, 130 Heckman, T. M., Armus, L. & Miley, G. K. 1990, ApJS, 74, 833 Heckman, T. M., Lehnert, M. & Armus, L. 1993, in "The Evolution of Galaxies and their Environment", edited by H. A. Thronson and J. M. Shull (Kluwer, Dordrecht), p. 455 Hernquist, L., Spergel, D. N. & Heyl, J. S. 1993, ApJ, 416, 415 Hibbard, J.E. 1995, PhD Thesis, Columbia University Hibbard, J.E., Guhathakurta, P., van Gorkom, J.H., & Schweizer, F. 1994, AJ, 107, 67 Hibbard, J.E., & van Gorkom, J.H. 1996, AJ, 111, 655 de Jong, R. S. 1995, PhD Thesis, University of Groningen Kormendy, J., & Sanders, D.B. 1992, ApJ, 390, L53 Lake, G. 1989, AJ, 97, 1312 Lake, G., & Dressler, A. 1986, ApJ, 310, 605 Malin, D. F. & Hadley, B. 1997, PASA, 14, 52 Mihos, J.C., & Hernquist, L. 1996, ApJ, 437, L47 (MH94) Negroponte, J., White, S.D.M. 1983, MNRAS, 205, 1009 Noguchi, M., & Ishibashi, S. 1986, MNRAS, 219, 305 Sanders, D.B., Scoville, N.Z., Sargent, A.I., Soifer, B.T. 1988b, ApJ, 324, L55 Schweizer, 1982, ApJ, 252, 455 Schweizer, F. 1996, AJ, 111, 109 Schweizer, F., & Seitzer, P. 1992, AJ, 104, 3 Scoville, N.Z., & Soifer, B.T. 1991, in Massive Stars in Starbursts, eds. C. Leitherer, N.R. Walborn, T.M. Heckman, & C.A. Norman (Cambridge Univ. Press: New York), p 233. Scoville, N.Z., Yun, M.S., & Bryant, P.M. 1997, ApJ, 484, 702 Scoville, N.Z., Evans, A.S., Dinshaw, N., Thompson, R., Rieke, M., Schneider, G., Low, F.J., Hines, D., Stobie, B. & Epps, H. 1998, ApJ, 492, L107 Toomre, A. 1977, in "The Evolution of Galaxies and Stellar Populations", edited by B. M. Tinsley and R. B. Larson (Yale University Press, New Haven), p. 401 Toomre, A., & Toomre, J. 1972, ApJ, 178, 623 Veeraraghavan, S. & White, S. D. M. 1985, ApJ, 296, 336 Wang, Z., Schweizer, F., & Scoville, N.Z. 1992, ApJ, 396, 510 Whitmore, B.C., Schweizer, F., Leitherer, C., Borne, K., & Robert, C. 1993, AJ, 106, 1354 Young, J. S., Schloerb, F. P., Kenney, J. D., & Lord, S. D. 1986, ApJ, 304, 443 Young, J.S., & Scoville, N.Z. 1991, ARA&A, 29, 581 Yun, M.S., & Hibbard, J. E. 1999a, ApJ, submitted Yun, M. S., & Hibbard, J. E., 1999b, submitted Hibbard & Yun 5 Fig. 1. -- Radial surface molecular gas mass density profiles (in M⊙ pc−2) derived from the integrated CO flux maps are plotted as a function of radius. The CO fluxes have been turned into mass densities by adopting a conversion factor of NH2 /ICO = 3 × 1020 cm−2 (K km s−1)−1. The data are from Yun & Hibbard (1999a) for NGC 3921 (dotted line); from Wang et al. (1991) for NGC 7252 (solid line); and from Scoville et al. (1997) for Arp 220 (dashed line). 6 Luminosity Profiles of Merger Remnants Fig. 2. -- Predicted luminosity profiles as a function of the fourth-root of the radius for the aged remnants of NGC 3921, NGC 7252, and Arp 220. Dotted lines: expected contribution to the B−band profile by the present stellar component, after aging by 2 Gyr. For NGC 3921 and NGC 7252 this is calculated from the presently observed B−band luminosity profiles (taken from the data in Schweizer 1996 and Hibbard et al. 1994) and fading by 1 mag arcsec−2. For Arp 220 we use the observed K −band luminosity profile (from Scoville et al. 1998) adopting B − K = 4 for a 10 Gyr old population. Dashed lines: expected contribution to the B−band profile from the expected post-merger population if all the molecular gas is turned into stars. Solid lines: the expected B−band luminosity profiles after adding in the aged stellar and post-merger populations. No significant deviation from a pure r1/4 law expected for NGC 3921 or NGC 7252. Arp 220 does not show such a seamless profile, exhibiting a slight rise in at small radii.
astro-ph/0207488
1
0207
2002-07-23T09:24:14
The status of Galactic field lambda Bootis stars in the post-Hipparcos era
[ "astro-ph" ]
The lambda Bootis stars are Population I, late B to early F-type stars, with moderate to extreme (up to a factor 100) surface underabundances of most Fe-peak elements and solar abundances of lighter elements (C, N, O, and S). To put constraints on the various existing theories that try to explain these peculiar stars, we investigate the observational properties of lambda Bootis stars compared to a reference sample of normal stars. Using various photometric systems and Hipparcos data, we analyze the validity of standard photometric calibrations, elemental abundances, and Galactic space motions. There crystallizes a clear picture of a homogeneous group of Population I objects found at all stages of their main-sequence evolution, with a peak at about 1 Gyr. No correlation of astrophysical parameters such as the projected rotational velocities or elemental abundances with age is found, suggesting that the a-priori unknown mechanism, which creates lambda Bootis stars, works continuously for late B to early F-type stars in all stages of main-sequence evolution. Surprisingly, the sodium abundances seem to indicate an interaction between the stars and their local environment.
astro-ph
astro-ph
Mon. Not. R. Astron. Soc. 000, 000 -- 000 () Printed 2 November 2018 (MN LATEX style file v2.2) The status of Galactic field λ Bootis stars in the post-Hipparcos era E. Paunzen1,2, I.Kh. Iliev3, I. Kamp4, I.S. Barzova3 1-Institut fur Astronomie der Universitat Wien, Turkenschanzstr. 17, 1180 Wien, Austria 2-Zentraler Informatikdienst der Universitat Wien, Universitatsstr. 7, 1010 Wien, Austria 3-Institute of Astronomy, National Astronomical Observatory, P.O. Box 136, 4700 Smolyan, Bulgaria 4-Leiden Observatory, Niels Bohrweg 2, PO Box 9513, 2330 RA Leiden, The Netherlands 2 November 2018 ABSTRACT The λ Bootis stars are Population I, late B to early F-type stars, with moderate to extreme (up to a factor 100) surface underabundances of most Fe-peak elements and solar abundances of lighter elements (C, N, O, and S). To put constraints on the various existing theories that try to explain these peculiar stars, we investigate the observational properties of λ Bootis stars compared to a reference sample of normal stars. Using various photometric systems and Hipparcos data, we analyze the valid- ity of standard photometric calibrations, elemental abundances, and Galactic space motions. There crystallizes a clear picture of a homogeneous group of Population I objects found at all stages of their main-sequence evolution, with a peak at about 1 Gyr. No correlation of astrophysical parameters such as the projected rotational velocities or elemental abundances with age is found, suggesting that the a-priori unknown mechanism, which creates λ Bootis stars, works continuously for late B to early F-type stars in all stages of main-sequence evolution. Surprisingly, the sodium abundances seem to indicate an interaction between the stars and their local environment. Key words: Stars: λ Bootis -- Stars: chemically peculiar -- Stars: early-type 1 INTRODUCTION Knowledge of the evolutionary status of the members of the λ Bootis group is essential to put tight constraints on the as- trophysical processes behind this phenomenon. We have cho- sen the following working definition as group characteristics: late B- to early F-type, Population I stars with apparently solar abundances of the light elements (C, N, O and S) and moderate to strong underabundances of Fe-peak elements (see Faraggiana & Gerbaldi 1998 for a critical summary of the various definitions). Only a maximum of about 2% of all objects in the relevant spectral domain are believed to be λ Bootis -type stars (Paunzen 2001). That already suggests either that the mechanism responsible for the phenomenon works on a very short time-scale (106 yrs) or else the general conditions for the development are very strict. We know already of a few λ Bootis stars in the Orion OB1 association and one candidate in NGC 2264 (Paun- zen 2001), for both of which log t ≈ 7.0. The evolutionary status for two λ Bootis -type spectroscopic-binary systems (HD 84948 and HD 171948) has been determined as very close to the Zero Age Main Sequence (ZAMS hereafter) for HD 171948 and to the Terminal Age Main Sequence (TAMS hereafter) for HD 84948 (Iliev et al. 2002). The results for the other Galactic field stars are not clear. In 1995, Iliev & Barzova summarized the evolution- ary status of 20 λ Bootis stars (and Vega) using Stromgren uvbyβ photometric data. They concluded that most of the stars studied are in the middle of their main-sequence evo- lution, with only a few objects near the ZAMS. Paunzen (1997) investigated the parallaxes measured by the Hipparcos satellite for a sample of λ Bootis -type stars in order to derive luminosities, masses and ages for 18 ob- jects in common with Iliev & Barzova (1995). He found no systematic influence of the distance, effective temperature, metallicity and rotational velocity on the difference between photometrically calibrated absolute magnitudes and those derived from Hipparcos parallaxes. Six objects were found to be very close to the ZAMS and a hypothesis was pro- posed that all other stars are in their Pre-Main-Sequence (PMS hereafter) phase. Later on, Bohlender, Gonzalez & Matthews (1999) and Faraggiana & Bonifacio (1999) challenged that hypothesis with plausible arguments such as the unusually vigorous star-forming activity that it implied in the solar neighbour- hood and a statistical analysis of normal-type stars. 2 E. Paunzen, I.Kh. Iliev, I. Kamp, I.S. Barzova Already Gray & Corbally (1998) stated that the λ Bootis phenomenon can be found from very early stages to well into the main-sequence life of A-type stars. That con- clusion was based on the incidence of λ Bootis stars among very young A-type stars, which is not very different from the incidence among Galactic field stars. In this paper we present a much more extensive in- vestigation, including the data of the Hipparcos satellite. With the help of photometric data of the Johnson UBV, Stromgren uvbyβ and Geneva 7-colour systems, different cal- ibrations of the absolute magnitude, effective temperature and surface gravity are applied and compared. From evolutionary models (Claret 1995), masses and ages are estimated. They are compared with those derived by Iliev & Barzova (1995) and Paunzen (1997). As a further step, the proper motions of λ Bootis stars are used to calculate space velocities. That very important information should help further to sharpen the group prop- erties and to sort out probably misclassified stars. Another point investigated is the question as to whether there exists a typical abundance pattern for the λ Bootis group. Heiter (2002) and Heiter et al. (2002) tried to shed more light on the abundance pattern in the context of the proposed theories. We have searched for correlations of the individual abundances with, especially, mass and age. However, we will not comment on our results in the context of the developed theories and models, because they still depend on too many free parameters. The aim of this paper is not to promote one of the suggested theories but rather to find strict observational constraints which should be incorporated into future theoretical investigations. 2 PROGRAM STARS AND THEIR ASTROPHYSICAL PARAMETERS The program stars were taken from the lists of Gray & Corbally (1998) and Paunzen (2001), with the omission of apparent spectroscopic binary systems (e.g. HD 38545, HD 64491, HD 111786, HD 141851 and HD 148638) as well as of HD 191850 for which no β measurement is available so no reliable astrophysical parameters could be derived from the Stromgren photometric system. A further critical as- sessment of the literature was performed in order to reject probable non-members and ill-defined objects. In total, 57 well established λ Bootis stars were chosen. The data were photometric from the General Catalogue (GCPD; http://obswww.unige.ch/gcpd/) from the Hipparcos and Tycho database. If available, averaged and weighted mean values were used. of Photometric Data taken as well as The following calibrations for the individual photomet- ric systems were used to derive the effective temperatures and surface gravities: • Johnson UBV: Napiwotzki, Schonberner & Wenske (1993) • Stromgren uvbyβ: Moon & Dworetsky (1985) and Napi- wotzki et al. (1993) • Geneva 7-colour: Kobi & North (1990) and Kunzli et al. (1997) The reddening and absolute magnitudes were estimated by use of the Stromgren uvbyβ system. The calibrations for the Johnson UBV and Geneva 7-colour system need an a-priori knowledge of the reddening which is not easy to estimate. An independent way to derive the interstellar redden- ing maps is to use data from open clusters as well as Galac- tic field stars. Several different models have been published in the literature (Arenou, Grenon & G´omez 1992, Hakkila et al. 1997). Chen et al. (1998) compared the results from Arenou et al. (1992) and those derived from Hipparcos mea- surements, and found an overestimation in previously pub- lished results for distances less than 500 pc. They conse- quently proposed a new model for Galactic latitudes of ±10◦, but otherwise find excellent agreement with the model by Sandage (1972). We use here the model proposed by Chen et al. (1998) who corrected previous models by Arenou et al. (1992) on the basis of the Hipparcos data. The values from the calibration of the Stromgren uvbyβ and the model by Chen et al. (1998) are in very good agreement. To min- imize possible inconsistencies we have averaged the values from both approaches. The Hipparcos parallaxes were converted directly into absolute magnitudes. The latter were averaged (using weights based on the standard errors) with the abso- lute magnitudes derived from the photometric calibra- tions. With the absolute bolometric magnitude of the Sun (MBol)⊙ = 4.75 mag (Cayrel de Strobel 1996) and the bolo- metric correction taken from Drilling & Landolt (2000), lu- minosities (log L∗/L⊙) were calculated (Table 1). We have also corrected our data for the so-called 'Lutz -- Kelker effect' (Koen 1992), an overestimation of paral- laxes owing to random errors. Oudmaijer, Groenewegen & Schrijver (1998) showed that that bias has to be taken into account if individual absolute magnitudes from Hipparcos parallaxes are calculated. The projected rotational velocities were taken from Ue- sugi & Fukuda (1982), Sturenburg (1993), Abt & Morrell (1995), Holweger & Rentzsch-Holm (1995), Chernyshova et al. (1998), Heiter et al. (1998), Paunzen et al. (1999a,b), Kamp et al. (2001), Solano et al. (2001), Andrievsky et al. (2002) and Heiter (2002). If possible the individual val- ues were averaged with weights in accordance with the listed standard errors. As the next step, we have used the post-MS evolution- ary tracks and isochrones from Claret (1995) to estimate the individual masses and ages for our program stars. The mod- els were calculated with solar abundances. That is justified because the abundance pattern found for the λ Bootis stars must be restricted to the surface only: almost all of them would lie below the Population II ZAMS which would be ap- plicable if they were metal-poor throughout. Table 2 lists the masses together with minimum and maximum ages, which are not necessarily equally spaced owing to the shape of the isochrones. Since the lifetime for a star on the MS is dependent on its mass, we have transformed the ages thus determined into relative ones (trel) in the following way. For our sample we find masses (Table 2) from 1.6 to 2.5 M⊙, which correspond to times on the MS of 2.2 Gyr down to 700 Myr. The relative age is zero for an object which just arrived at the ZAMS and unity for stars at the TAMS. We have taken into account the error of the estimated mass as well as the error box of the calibrated ages. The status of Galactic field λ Bootis stars in the post-Hipparcos era 3 Table 1. Photometric data, stellar parameters and calibrated values for the program stars. In parenthesis are the errors in the final digits of the corresponding quantity. HD HR HIP V [mag] B − V [mag] b − y [mag] AV [mag] v sin i [kms−1] Teff [K] log g(phot) [dex] MV [mag] log L∗/L⊙ 319 6870 7908 11413 13755 15165 23392 24472 30422 31295 35242 54272 74873 75654 81290 83041 83277 84123 87271 90821 91130 101108 102541 105058 105759 106223 107233 109738 110377 110411 111005 111604 120500 120896 125162 125889 130767 142703 142944 149130 153747 154153 156954 168740 168947 170680 175445 183324 184779 192640 193256 193281 198160 204041 210111 216847 221756 12 541 1525 1570 1777 3481 3517 4124 4824 4828 4875 5351 5930 6338 6871 6944 7400 636 5321 6108 8593 10304 11390 17462 18153 22192 22845 25205 43121 43354 46011 47018 47155 47752 49328 51556 56768 57567 58992 59346 59594 60134 61937 61960 62318 62641 67481 67705 69732 72505 78078 81329 83410 83650 84895 90304 90806 92884 95793 99770 7736 7764C 100286 7764A 100288 7959 102962 105819 8203 109306 8437 113351 116354 8947 5.934 7.494 7.288 5.940 7.844 6.705 8.260 7.092 6.186 4.648 6.348 8.800 5.890 6.384 8.866 8.927 8.304 6.840 7.120 9.470 5.902 8.880 7.939 8.900 6.550 7.431 7.353 8.277 6.228 4.881 7.959 5.886 6.600 8.495 4.186 9.849 6.905 6.120 7.179 8.498 7.420 6.185 7.679 6.138 8.123 5.132 7.792 5.795 8.940 4.934 7.721 6.557 5.663 6.456 6.377 7.060 5.576 0.141 0.246 0.272 0.147 0.318 0.333 0.020 0.304 0.190 0.085 0.122 0.261 0.115 0.242 0.332 0.294 0.311 0.297 0.172 0.105 0.109 0.179 0.230 0.183 0.218 0.288 0.255 0.198 0.195 0.077 0.376 0.160 0.131 0.296 0.084 0.241 0.042 0.240 0.293 0.342 0.140 0.284 0.294 0.201 0.196 0.008 0.110 0.084 0.224 0.154 0.213 0.190 0.189 0.161 0.203 0.242 0.095 0.079 0.164 0.192 0.105 0.181 0.191 0.014 0.214 0.101 0.044 0.068 0.214 0.064 0.161 0.254 0.223 0.226 0.235 0.151 0.068 0.073 0.114 0.163 0.129 0.142 0.228 0.192 0.161 0.120 0.040 0.224 0.112 0.068 0.166 0.051 0.206 0.002 0.182 0.201 0.233 0.098 0.199 0.200 0.136 0.172 0.008 0.055 0.051 0.187 0.099 0.116 0.098 0.108 0.092 0.136 0.155 0.056 0.004 0.000 0.000 0.004 0.000 0.010 0.094 0.003 0.014 0.063 0.042 0.000 0.078 0.012 0.124 0.161 0.131 0.040 0.008 0.013 0.000 0.006 0.097 0.009 0.000 0.015 0.048 0.073 0.000 0.045 0.009 0.000 0.017 0.000 0.039 0.108 0.000 0.021 0.013 0.109 0.128 0.020 0.050 0.035 0.116 0.091 0.108 0.083 0.000 0.016 0.063 0.111 0.022 0.026 0.000 0.000 0.043 60 165 125 90 135 115 90 130 45 55 95 20 150 135 90 140 120 90 95 170 165 180 125 115 100 180 50 145 205 90 80 250 95 200 65 55 105 8020(135) 7330(102) 7145(87) 7925(124) 7080(161) 7010(167) 9805(281) 6945(131) 7865(108) 8920(177) 8250(103) 7010(217) 8700(245) 7350(104) 6895(214) 7120(208) 7000(189) 7025(175) 7515(232) 8190(79) 8135(98) 7810(80) 7665(168) 7740(171) 7485(102) 6855(247) 7265(143) 7610(145) 7720(89) 8930(206) 6860(66) 7760(149) 8220(70) 7260(89) 8720(156) 7275(175) 9195(220) 7265(150) 7000(125) 6945(103) 8205(90) 7055(120) 7130(93) 7630(81) 7555(185) 9840(248) 8520(198) 8950(204) 7210(173) 7940(96) 7740(94) 8035(115) 7870(129) 7980(97) 7550(123) 7355(78) 8510(188) 3.74(8) 3.84(11) 4.10(12) 3.91(21) 3.26(10) 3.23(10) 4.35(9) 3.81(16) 4.00(20) 4.20(1) 3.90(14) 3.83(10) 4.21(11) 3.77(11) 3.82(28) 3.76(20) 3.67(18) 3.73(17) 3.43(10) 3.73(10) 3.78(10) 3.90(18) 4.22(16) 3.77(30) 3.65(10) 3.49(18) 4.03(10) 3.90(13) 3.97(14) 4.14(14) 3.72(10) 3.61(25) 3.86(10) 3.76(10) 4.07(9) 3.88(9) 4.10(8) 3.93(12) 3.19(4) 3.49(5) 3.70(24) 3.56(6) 4.04(13) 3.88(14) 3.67(10) 4.15(6) 3.96(10) 4.13(4) 3.63(21) 3.95(18) 3.69(17) 3.54(4) 3.99(9) 3.97(8) 3.84(15) 3.47(14) 3.90(3) 1.27(19) 2.29(42) 2.60(18) 1.49(10) 0.93(10) 1.12(16) 1.43(30) 2.14(11) 2.35(1) 1.66(22) 1.75(22) 2.33(30) 1.82(1) 1.83(12) 1.85(30) 1.70(30) 1.49(29) 1.58(15) 1.02(8) 0.74(30) 1.36(26) 1.33(30) 2.34(21) 0.86(30) 1.35(21) 1.83(45) 2.64(13) 1.85(30) 1.96(11) 1.90(28) 1.76(53) 0.48(7) 0.85(34) 1.90(30) 1.71(23) 2.32(30) 1.27(1) 2.41(12) 0.80(30) 1.51(30) 1.24(30) 1.86(29) 2.81(33) 1.82(2) 1.28(30) 0.83(23) 1.08(27) 1.64(42) 1.26(30) 1.84(1) 1.08(30) 0.41(30) 1.47(41) 1.75(18) 1.76(15) 0.93(24) 1.16(16) 1.45(8) 1.02(17) 0.90(7) 1.35(4) 1.57(4) 1.50(6) 1.45(12) 1.09(5) 1.01(1) 1.32(9) 1.26(9) 1.01(12) 1.24(1) 1.20(5) 1.20(12) 1.26(12) 1.35(12) 1.31(6) 1.53(3) 1.66(12) 1.42(11) 1.42(12) 1.01(9) 1.60(12) 1.40(8) 1.22(18) 0.88(5) 1.20(12) 1.16(5) 1.22(11) 1.24(21) 1.75(3) 1.62(13) 1.18(12) 1.28(9) 1.01(12) 1.48(1) 0.97(5) 1.62(12) 1.34(12) 1.46(12) 1.19(11) 0.82(13) 1.21(1) 1.43(12) 1.70(9) 1.53(11) 1.32(17) 1.43(12) 1.22(1) 1.51(12) 1.79(12) 1.36(16) 1.25(7) 1.23(6) 1.56(10) 1.50(6) 4 E. Paunzen, I.Kh. Iliev, I. Kamp, I.S. Barzova Table 2. Calibrated values for the program stars using the models of Claret (1995); trel is the relative age ranging from zero for an object just arriving at the ZAMS to unity for a star at the TAMS; ∆log g = log g(evol) − log g(phot); log t1 and log t2 are the minimum and maximum ages derived from the error boxes. HD M [M⊙] ∆log g log t [dex] log t1 [dex] log t2 [dex] trel 319 6870 7908 11413 13755 15165 23392 24472 30422 31295 35242 54272 74873 75654 81290 83041 83277 84123 87271 90821 91130 101108 102541 105058 105759 106223 107233 109738 110377 110411 111005 111604 120500 120896 125162 125889 130767 142703 142944 149130 153747 154153 156954 168740 168947 170680 175445 183324 184779 192640 193256 193281 198160 204041 210111 216847 221756 2.12(9) +0.15 1.70(12) +0.22 +0.02 1.61(5) 2.02(4) +0.03 +0.31 2.21(4) 2.14(6) +0.38 2.30(11) −0.09 1.74(4) +0.10 1.76(2) +0.21 2.08(10) +0.00 1.96(6) +0.19 1.69(10) +0.16 2.00(3) +0.00 1.85(5) +0.15 1.85(11) −0.01 1.90(11) +0.06 1.98(12) +0.05 1.95(5) +0.03 2.20(3) +0.28 2.38(16) +0.03 2.10(10) +0.16 2.08(12) −0.03 1.75(6) −0.06 2.26(11) −0.06 2.05(8) +0.15 1.85(15) +0.30 1.62(4) +0.14 1.87(12) +0.09 1.84(4) +0.08 2.02(10) +0.14 1.88(20) +0.05 2.42(3) −0.02 2.30(12) −0.06 1.83(11) +0.16 2.04(7) +0.12 1.70(10) +0.19 +0.02 2.28(4) 1.67(2) +0.15 2.25(11) +0.31 1.97(12) +0.23 2.15(12) +0.21 1.84(10) +0.30 1.56(8) +0.14 1.88(2) +0.10 2.09(12) +0.13 2.50(10) −0.08 2.25(12) −0.03 2.09(17) +0.07 2.08(12) +0.08 1.90(2) +0.10 2.18(12) +0.09 2.50(12) +0.08 2.02(18) −0.06 1.93(7) +0.06 1.90(5) +0.11 2.21(10) +0.17 2.20(8) +0.06 8.88 9.02 9.02 8.91 9.01 9.00 8.00 9.13 8.69 8.51 8.81 9.12 8.52 9.03 9.10 9.05 9.03 9.04 8.90 8.79 8.86 8.92 8.85 8.85 8.96 9.10 8.93 8.98 8.94 8.00 9.09 8.79 8.81 9.05 8.61 9.03 8.60 9.02 8.99 9.04 8.84 9.08 8.92 8.97 8.94 8.48 8.78 8.50 8.97 8.90 8.89 8.75 8.92 8.90 8.98 8.90 8.78 8.84 8.70 8.88 8.88 8.90 8.93 7.00 9.09 8.52 7.60 8.69 9.00 8.20 9.01 9.03 8.99 8.96 9.01 8.80 8.74 8.83 8.87 8.42 8.79 8.92 9.00 8.60 8.90 8.90 7.00 8.99 8.77 8.75 9.01 8.00 8.99 8.48 8.83 8.90 8.98 8.81 9.03 7.00 8.94 8.88 8.30 8.73 7.00 8.91 8.85 8.83 8.70 8.86 8.82 8.95 8.87 8.73 8.91 9.05 9.11 8.93 9.05 9.10 8.50 9.17 8.80 8.71 8.86 9.20 8.70 9.05 9.17 9.11 9.10 9.10 9.00 8.90 8.89 8.92 8.99 9.00 9.10 9.20 9.06 9.02 8.97 8.66 9.18 8.90 8.90 9.06 8.75 9.10 8.68 9.10 9.06 9.10 8.86 9.13 9.01 8.99 9.01 8.57 8.82 8.73 9.03 8.91 9.00 8.81 8.95 8.92 9.02 9.00 8.82 0.86(5) 0.69(13) 0.70(8) 0.83(3) 1.03(8) 1.01(9) 0.35(23) 0.88(4) 0.54(7) 0.47(23) 0.70(6) 0.83(10) 0.51(13) 0.85(3) 0.92(9) 0.90(8) 0.93(9) 0.93(5) 0.93(9) 0.93(11) 0.83(5) 0.86(6) 0.60(16) 0.94(11) 0.94(10) 0.92(12) 0.61(13) 0.81(7) 0.76(3) 0.33(24) 0.93(14) 0.96(7) 0.90(9) 0.85(6) 0.53(19) 0.77(7) 0.68(7) 0.71(9) 1.05(10) 0.94(8) 0.83(6) 0.89(7) 0.44(29) 0.81(2) 0.91(8) 0.66(10) 0.83(6) 0.45(29) 0.93(8) 0.75(2) 0.92(10) 0.93(7) 0.84(9) 0.76(5) 0.84(4) 0.96(8) 0.80(5) ) n u s ( L / L g o l 2.0 1.9 1.8 1.7 1.6 1.5 1.4 1.3 1.2 1.1 1.0 0.9 0.8 0.7 0.6 Bootis normal stars l 7.4 7.8 8.4 8.6 8.8 9.0 9.2 4.00 3.98 3.96 3.94 3.92 3.90 3.88 3.86 3.84 3.82 3.80 log Teff Figure 1. The sample of λ Bootis (filled circles) and normal-type (open circles) stars within a log L∗/L⊙ versus log Teff diagram. The isochrones (the different log t values are listed in the legend) were taken from Claret (1995). Table 3. Normal-type stars selected as comparison. HD HIP HD HIP HD HIP 2262 3003 5382 9919 11636 13041 16555 17943 19107 39060 40136 45320 49434 50241 50277 50506 56405 59037 70574 2072 2578 4366 7535 8903 9977 12225 13421 14293 27321 28103 30666 32617 32607 33024 31897 35180 36393 41036 71297 79439 85364 87696 88824 96113 98058 98353 101107 102124 105211 111968 116706 118232 122405 135379 135559 145631 159561 41375 45493 48341 49593 50070 54137 55084 55266 56770 57328 59072 62896 65466 66234 68478 74824 74689 79439 86032 160613 170642 175638 181296 186689 187642 192425 195050 196078 201184 205835 205852 205924 207235 207958 210300 210739 211356 220061 86565 90887 92946 95261 97229 97649 99742 100907 101608 104365 106711 106787 106856 107596 108036 109412 109667 109984 115250 The status of Galactic field λ Bootis stars in the post-Hipparcos era 5 26 24 22 20 18 16 14 12 10 8 6 4 2 0 -0.05 0.00 0.05 0.10 0.20 0.25 0.30 0.35 0.15 (B - V)0 Bootis l normal stars 7.0 7.2 7.4 7.6 7.8 8.0 8.2 log age 8.4 8.6 8.8 9.0 9.2 e g a t n e c r e P e g a t n e c r e P 30 28 26 24 22 20 18 16 14 12 10 8 6 4 2 0 24 22 20 18 16 14 12 10 8 6 4 2 e g a t n e c r e P e g a t n e c r e P 0 0.0 0.1 0.2 0.3 0.4 0.5 0.7 0.8 0.9 1.0 1.1 1.2 0.6 trel 36 34 32 30 28 26 24 22 20 18 16 14 12 10 8 6 4 2 0 10000 9500 9000 8500 8000 7500 7000 6500 Teff Figure 2. Histogram of the calibrated ages (upper panel) and relative ages for both samples (lower panel). Figure 4. Histogram of (B − V )0 values (upper panel) and the calibrated effective temperature (lower panel) for the sample of λ Bootis stars. log g = +0.10(11) 4.50 4.35 4.20 4.05 3.90 3.75 3.60 3.45 3.30 3.15 ) t o h p ( g g o l 3.15 3.30 3.45 3.60 3.75 3.90 log g (evol) 4.05 4.20 4.35 4.50 Figure 3. Surface gravities calibrated photometrically and cal- culated via the mass, luminosity and effective temperature; the mean difference ∆log g = +0.10(11) is not significant. 3 A TEST SAMPLE OF NORMAL STARS results obtained for comparison with the For the λ Bootis stars we generated a test sample of apparently nor- mal dwarfs with the same effective temperatures. We limited the sample to luminosity class V objects for which photo- metric and Hipparcos measurements are available. Known spectroscopic binary systems were excluded. Then, only ob- jects for which the spectral classifications published by Gray & Garrison (1987; 1989a,b) and Garrison & Gray (1994) agree within two temperature sub-classes with those by Abt & Morell (1995) have been considered. The final sample (see Table 3) has the same number of stars as the λ Bootis list. Special attention has been paid to the (B − V )0 distribu- tion of the test sample to ensure that it similar to that of the λ Bootis objects (Fig. 4, upper panel). Thus, the test sample represents luminosity class V stars in the so- lar neighbourhood, covering the same spectral range as the λ Bootis group. All basic parameters were derived in exactly the same way as for the λ Bootis objects. All statistical significance levels presented in the following sections are based on several hypothesis tests (e.g. t-test; Rees 1987). We have not explicitly listed the parameters of this sam- ple but it is available in electronic form via anonymous ftp 130.79.128.5 or http://cdsweb.u-strasbg.fr/Abstract.html. D 6 E. Paunzen, I.Kh. Iliev, I. Kamp, I.S. Barzova 4 THE HERTZSPRUNG -- RUSSELL DIAGRAM BASED ON HIPPARCOS AND PHOTOMETRIC DATA Earlier investigations already concluded that the group of λ Bootis stars comprises the whole area between the ZAMS and TAMS. Only the interpretation of the evolutionary stage (PMS or post-MS) varied. In this work we have not inves- tigated a possible PMS hypothesis as proposed by Paun- zen (1997). Beside the arguments given in Bohlender et al. (1999) and Faraggiana & Bonifacio (1999) we would like to add a few more points. If we speculate that some of these stars are in a PMS phase then they would have ages of one Myr or even younger since they are in the 'right upper cor- ner' of the Hertzsprung -- Russell diagram (HRD hereafter) very close to the birth-line according to the models of Palla & Stahler (1993). So they should be very bright in the IR region and lie within the boundaries of star-forming regions or molecular clouds; neither implication has been verified so far (King 1994). Figure 1 shows the HRD with the data taken from Ta- ble 1. The post-MS evolutionary isochrones are from Claret (1995). It is encouraging to see that no object lies below the ZAMS; that gives further confidence in the calibrations used. To estimate the most probable age distribution of our program stars, we have used a moving average (Fig. 2). Such a method takes into account the errors of individual data as it counts all points (plus and minus the standard deviation) which lie in a certain bin (Rees 1987). From the histogram it is clear that the group of λ Bootis stars comprises evo- lutionary stages of the entire MS with a peak at trel ≈ 0.85 (about 1 Gyr). Taking the few candidates in the Orion OB1 association and NGC 2264, the percentage for very young objects (log t < 7.0) would be approximately 5% which fits very well into the overall picture. Since the distribution of the test sample (Fig. 2) agrees at a 99.9% significance level, the age distribution of the λ Bootis objects is not distinct from that of other luminos- ity class V objects in the same spectral domain within the solar neighbourhood. That is already a first hint that the mechanism behind the λ Bootis phenomenon operates only within very tight constraints, since only 2% of all stars are objects of the λ Bootis type. Gray & Corbally (1998, 1999) made an extensive survey for λ Bootis stars in 24 open clusters and associations and found not a single candidate. The cluster ages range from 15 to 700 Myr. Therefore, only a few candidates in the Orion OB1 association and NGC 2264 remain. It seems that the star-formation process within open clusters does not favour the manifestation of the λ Bootis phenomenon. Table 4 lists the 16 stars that we have in common with the papers of Iliev & Barzova (1995) and Paunzen (1997). For the calibration of effective temperatures they both used the method of Moon & Dworetsky (1985) within the Stromgren uvbyβ photometric system; Iliev & Barzova (1995) also derived the luminosities with this calibration whereas Paunzen (1997) took advantage of the Hipparcos data. For the calibration of mass and age, Iliev & Barzova (1995) interpolated between the evolutionary tracks given by Schaller et al. (1992), whereas Paunzen (1997) used the CESAM models by Morel (1997). As expected, there are star-to-star variations, but overall the parameters fit well. The effective temperatures from this work are in very good agreement with those of both references. The luminosities and thus the calibrated ages are in better agreement with those of Iliev & Barzova (1995). As a further test of the calibrated values, we have cal- culated the surface gravity via the effective temperature, mass and luminosity. Figure 3 shows the correlation of the photometrically calibrated surface gravity log g(phot) and the calculated one log g(evol). The mean errors for both parameters are typically ±0.1 dex. Although there is some indication of a systematic offset, the mean difference ∆log g = log g(evol) − log g(phot) = +0.10(11) is not signifi- cant. A similar effect was already noticed by Iliev & Barzova (1995) whereas Faraggiana & Bonifacio (1999) reported an inconsistency between the position in the HRD and the log g values derived from Stromgren uvbyβ photometry using the calibration of Moon & Dworetsky (1985). We conclude that the photometric calibrations for the group of λ Bootis stars are valid, in contradiction to the results of Faraggiana & Bonifacio (1999). Before making a detailed statistical analysis, we looked to see whether our sample exhibits an apparent bias. Fig- ure 4 shows that the sample includes significantly more cooler objects (70%) with effective temperatures lower than 8000 K. However, there is no observational bias from classification-resolution spectroscopy since the spectroscopic survey for new members included many more hotter- than cooler-type objects (Gray & Corbally 1999, Paunzen 2001). That fact could be interpreted as a manifestation of the working mechanism itself, or it could be due to the method used in this very limited spectral range. It is well known that at cooler temperatures (spectral domain A5 to F2), even a moderate metal-weakness can be detected at classification resolution since the overall metallic-line spectrum is much richer than for an A0-type object. However, we are not able to decide whether the distribution of the λ Bootis sample is due to a bias within the observational technique or due to the phenomenon itself. Figure 5 shows the averaged effective temperatures, masses and projected rotational velocities for each age bin. The mean effective temperature is rather constant up to trel ≈ 0.5 with values between 8300 K to 8800 K. It then de- creases linearly almost to 7000 K for the most evolved ob- jects. The mean masses are constant within the error bars at about 1.9 M⊙ for trel < 0.75 and then increase to almost 2.2 M⊙. Most interesting is the non-existence of a correla- tion of the projected rotational velocity with age. The mean value for the whole range is about 120 km s−1. If we compare these results with those of our test sample, we find several differences as well as similarities: • The trends for the effective temperature and mass are identical. However, the λ Bootis objects seem to have lower temperatures as well as masses for trel > 0.8 • The v sin i distributions are with a slightly higher (σLB = 19 km s−1 and σNor = 15 km s−1) scatter identical within 1σ the λ Bootis group for Overall, there is no obvious distinction between the two sam- ples. Gray (1988) reported that several λ Bootis stars exhibit peculiar hydrogen-line profiles with weak cores and broad The status of Galactic field λ Bootis stars in the post-Hipparcos era 7 Table 4. Comparison of the calibrated stellar parameters from this work (columns TW), Iliev & Barzova (1995; columns IL95) and Paunzen (1997; columns PA97). HD 319 11413 30422 31295 107233 110411 125162 142703 183324 192640 193256 193281 198160 204041 210111 221756 TW 3.904(7) 3.899(4) 3.896(6) 3.950(9) 3.861(9) 3.951(10) 3.935(5) 3.868(9) 3.950(5) 3.903(2) 3.894(9) 3.907(8) 3.907(13) 3.906(6) 3.888(3) 3.931(9) log Teff IL95 ±0.02 PA97 ±0.02 TW log L∗ /L⊙ IL95 ±0.01 PA97 TW M IL95 ±0.2 PA97 ±0.1 3.92 3.91 3.91 3.96 3.86 3.96 3.95 3.87 3.97 3.91 3.92 3.92 3.90 3.91 3.89 3.96 3.91 3.91 3.90 3.95 3.87 3.94 3.94 3.86 3.96 3.90 3.90 3.91 3.90 3.91 3.88 3.96 1.45(8) 1.35(4) 1.01(1) 1.32(9) 0.88(5) 1.22(11) 1.28(9) 0.97(5) 1.32(17) 1.22(1) 1.51(12) 1.79(12) 1.36(16) 1.25(7) 1.23(6) 1.50(6) 1.53 1.41 1.03 1.40 0.94 1.30 1.38 1.00 1.51 1.23 1.48 1.81 1.24 1.21 1.30 1.67 1.34(5) 1.28(4) 0.96(3) 1.20(3) 0.80(5) 1.09(3) 1.20(1) 0.93(3) 1.20(4) 1.16(2) 1.50(27) 1.96(27) 1.37(7) 1.20(7) 1.16(5) 1.41(4) 2.12(9) 2.02(4) 1.76(2) 2.08(10) 1.62(4) 2.02(10) 2.04(7) 1.67(2) 2.09(17) 1.90(2) 2.18(12) 2.50(12) 2.02(18) 1.93(7) 1.90(5) 2.20(8) 2.2 2.1 1.8 2.2 1.6 2.1 2.1 1.7 2.3 1.9 2.1 2.5 1.9 1.9 2.0 2.4 2.0 1.9 1.8 2.0 1.6 1.9 2.0 1.7 2.1 1.9 2.2 2.5 2.0 1.9 1.8 2.2 TW 8.88 8.91 8.69 8.51 8.93 8.00 8.61 9.02 8.50 8.90 8.89 8.75 8.92 8.90 8.98 8.78 log t IL95 ±0.05 PA97 ±0.05 trel TW IL95 8.81 8.87 8.58 8.54 9.04 8.30 8.58 9.02 8.52 8.85 8.86 8.72 8.86 8.81 8.94 8.65 8.76 8.82 7.00 7.00 7.00 7.00 7.20 8.50 7.00 8.65 8.79 8.84 8.77 8.57 8.84 8.10 0.86(5) 0.83(3) 0.54(7) 0.47(23) 0.61(13) 0.33(24) 0.53(19) 0.71(9) 0.45(29) 0.75(2) 0.92(10) 0.93(7) 0.84(9) 0.76(5) 0.84(4) 0.80(5) 0.77 0.71 0.23 0.38 0.52 0.19 0.40 0.56 0.42 0.54 0.78 0.91 0.58 0.48 0.69 0.68 but often shallow wings. Iliev & Barzova (1993) examined the peculiar profiles of four objects. They were able to fit those profiles with two models having different tempera- tures (hotter for the wings by approximately 400 K) and concluded that it is a sign of circumstellar material around the objects. Faraggiana & Bonifacio (1999) interpret the pe- culiar profiles as an indication of undetected spectroscopic- binary systems in which two stars with solar abundances but different stellar parameters mimic one apparently metal- weak object. The classification of the hydrogen-line profiles for 19 stars were taken from Gray & Corbally (1993) and Paunzen & Gray (1997). For our sample we find only two objects (HD 30422 and HD 107233) with peculiar profiles among the younger objects (trel < 0.66) but seven with nor- mal ones. The picture for the older objects is just the oppo- site: there are only two objects (HD 90821 and HD 120500) with normal profiles but nine with peculiar ones. Several attempts were made to detect signs of circum- stellar lines in the optical domain (Andrillat, Jaschek & Jaschek 1995; Hauck, Ballereau & Chauville 1995, 1998; Hol- weger, Hempel & Kamp 1999). From our list, sixteen objects were investigated but only three objects have positive detec- tions: HD 11413, HD 193256 and HD 198160 (Holweger et al. 1999). These stars are rather evolved (trel = 0.83, 0.92 and 0.84, respectively) but owing to the poor number statistics, any conclusion about the significance has to be treated with caution. 9750 9500 9250 9000 8750 8500 8250 8000 7750 7500 7250 7000 2.7 2.6 2.5 2.4 2.3 2.2 2.1 2.0 1.9 1.8 280 260 240 220 200 180 160 140 120 100 80 60 ] K [ f f e T ] ) n u s ( M [ M ] 1 - s m k [ i n i s v normal stars l Bootis 5 THE ABUNDANCE PATTERN OF λ BO OTIS STARS Since the first detailed abundance analysis in the late eight- ies, there has been a question as to whether a unique abun- dance pattern exists for the λ Bootis group. Most of the proposed candidates for membership have been found by classification-resolution spectroscopy (typically 40 A mm−1 to 120 A mm−1). In the classical photographic domain (3800 to 4600 A) spectral lines of Ca, Mg, Fe and Ti are the main contributors to the overall appearance of the metallic-line spectrum. Gray (1988) defined the membership of the class in terms of a weak Ca K and metallic-line spectrum com- pared to the temperature classification of the hydrogen lines. Paunzen et al. (1999a) and Kamp et al. (2001) already 0.0 0.1 0.2 0.3 0.4 0.5 0.6 trel 0.7 0.8 0.9 1.0 1.1 1.2 1.3 Figure 5. Correlation found for the effective temperature (lower panel) and mass (middle panel) with the relative age; no corre- lation is detected with the projected rotational velocity (upper panel). The λ Bootis (filled circles) and normal-type (open cir- cles) stars are binned with respect to their relative age. investigated the behaviour of the lights elements C, N, O and S for a statistically significant number of members: • The star-to-star scatter for the abundances of C, N, O and S is much smaller than that for heavier elements • The abundances of C, N, O and S are not strictly solar but range from −0.8 dex to +0.2 dex compared to the Sun • Fe-peak elements are always more underabundant 8 E. Paunzen, I.Kh. Iliev, I. Kamp, I.S. Barzova Table 5. Abundances from the literature for our program stars. Different values for a single element were added and a mean was calculated. All values are given with respect to the Sun as: [X] = log X -- log X⊙. In parentheses are the estimated errors of the means or the individual error taken from the relevant reference. HD C N O Na Mg Al Si S Ca -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- −0.10(30) +0.50(30) −0.90(10) −1.10(30) −0.68(20) +0.22(10) +1.00(10) −0.73 (5) −0.58(20) −0.45(10) 319 −0.16(15) −0.20(20) −0.10(10) −0.74(30) −1.18(20) 11413 −0.15(15) −0.90(30) +0.00(30) −1.00(30) −0.30(30) +0.00(30) −1.50(30) 15165 +0.00(30) −0.95(20) −0.16(20) −0.90(20) 31295 −0.17(15) −0.24(15) −0.29(15) −0.52(20) −1.15(20) 74873 +0.60(10) −1.00(30) 75654 −0.44(10) +0.00(15) +0.14(10) −0.10(15) −0.84(15) −0.52(18) −0.75(20) −0.50(10) −0.58(20) 84123 −0.01(24) +0.00(30) −0.30(20) −0.55(40) −0.76(20) 91130 +0.20(30) +0.30(15) +0.11(30) −0.63 (6) −0.45 (7) +0.18(15) −3.00(15) −0.60(10) −0.30(20) −0.90(30) −0.50(30) 101108 +0.10(10) −1.59(10) −1.40(30) −2.10(30) −1.70(30) 106223 +0.30(10) −1.40(10) −0.80(20) −2.20(30) 107233 −0.30(30) −0.50(10) −0.84(20) 110411 +0.10(20) +0.30(30) +0.47(15) −0.98 (5) 111604 −0.25 (5) 120500 +0.20(20) +0.00(15) −0.09(15) +0.32(30) −0.23 (7) 125162 −0.34(20) −0.12(15) −0.22(20) −1.30(15) −2.00(20) −1.20(20) 142703 −0.28(20) −0.50(15) −0.19(10) −0.90(10) 168740 −0.42(10) −0.03(10) 170680 −0.06(10) −0.07(10) −0.20(20) 183324 +0.11(20) −0.03(30) −0.10(30) +0.10(30) −1.73(20) −1.50(30) −1.10(20) −0.13(15) −1.32(30) −1.00(30) −0.31(15) −1.35(20) 192640 −0.19(20) −0.22(15) −0.32(30) −1.10(20) −1.62(20) 193256 −0.45(20) −0.23(10) +0.90(30) −0.15(20) +0.00(30) −0.51(30) +0.14(15) −0.68(20) 193281 −0.49(20) +0.30(15) −0.05(10) +1.00(20) −0.08(10) 198160 −0.18(30) −0.18(10) +0.30(20) −0.13(10) −0.67(30) 204041 −0.58(20) −0.05(15) −0.42(20) +0.36(20) −1.04(20) −0.93(20) −0.63(20) +0.07(20) −0.98(20) 210111 −0.23(10) −0.96(20) −1.00(20) +0.06(15) −0.49(20) 221756 +0.00(20) +0.20(15) +0.17(20) +1.17(20) −0.70(20) −0.30(30) −0.85(30) −0.95 (5) −0.13(15) −0.30(15) −1.00(20) −0.45(15) −1.98(15) −0.53(15) −1.20(10) −0.60(20) −0.20(10) +0.60(10) −1.00(12) −0.52(20) −0.65(10) −0.20(20) -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- Sc Ti Mn Cr Fe Ni Sr Y Ba 319 −1.10(20) −0.75(20) 11413 −1.10(30) −1.40(20) 15165 −1.40(30) −0.50(30) 31295 −0.90(30) −0.96(20) 74873 −0.40(30) −0.90(10) 75654 84123 −0.90(31) −0.86(10) −0.84(20) −0.85(30) −0.72(20) −0.74(30) −0.50(30) −0.55(13) −0.60(20) 91130 −0.35(30) −0.40(20) −0.64(20) −1.20(20) −1.43(20) −1.50(20) −0.90(30) −1.60(30) −1.00(30) −0.80(30) −0.75(20) −1.07(20) −0.50(20) −1.51(20) −0.30(20) −0.70(10) −0.20(30) −0.90(40) −1.01(29) −1.00(42) −1.10(30) −1.08(14) −0.48 (3) −1.16 (8) −0.25(20) −0.90(20) −0.64 (8) −0.45(20) -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- −1.30(30) −0.90(30) −1.10(20) 101108 −0.60(40) −0.20(20) −0.40(20) −0.60(10) −0.70(10) −0.30(10) −0.70(40) +0.00(40) −0.70(30) 106223 −1.10(30) −1.02(10) −1.70(10) −1.34(20) −1.60(10) −0.97(20) −1.70(30) −0.50(30) −1.60(20) 107233 −1.70(30) −1.34(30) −1.30(20) −1.28(30) −1.38(21) 110411 −1.10(30) −0.90(10) −1.07(20) −1.07(30) −1.08 (4) 111604 120500 −0.67 (6) 125162 −0.70(20) −2.01(20) 142703 −1.50(30) −1.27(10) 168740 −1.10(30) −0.88(10) 170680 −0.40(10) 183324 −1.46(40) −1.42(20) −2.00(30) 192640 −1.15(20) −1.33(20) −2.00(20) −1.65(20) −1.63(20) −0.60(20) −1.49(30) −0.50(30) 193256 −0.90(40) −0.30(30) −0.10(20) −0.36(20) 193281 198160 −0.50(20) −1.30(30) 204041 −1.03(40) −1.21(20) 210111 −1.30(20) −1.05(20) 221756 −0.60(20) −0.50(20) −1.00(20) −1.98(20) −1.20(20) −2.10(15) −1.40(10) −1.32(12) −0.90(20) −1.20(30) −1.10(10) −0.88(10) −1.00(30) −0.40(30) −0.50(10) −1.38(20) −1.50(30) −1.00(30) −0.90(30) −0.20(20) −0.93(20) −0.90(20) −0.78(30) −0.72(30) −0.87(20) −0.40(20) −1.18(20) −0.89(20) −0.50(20) −0.59(30) −0.67(20) −1.03(20) −0.88(50) −0.54(20) −0.98(30) −0.09(20) −0.20(20) −0.15(30) −0.90(20) −1.50(30) −0.50(30) -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- +0.45(20) -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- Heiter (2002) and Heiter et al. (2002) tried to shed more light on the abundance pattern in the context of the theo- ries proposed. They used mean values for a sample of 34 stars (not all elements' abundances were determined for all objects) and concluded: • The iron-peak elements from Sc to Fe as well as Mg, Si, Ca, Sr and Ba are underabundant by 1 dex compared to the Sun • Al is slightly more depleted whereas Ni, Y and Zr are slightly less depleted The status of Galactic field λ Bootis stars in the post-Hipparcos era 9 Table 6. Kinematic data from the Hipparcos database for the program stars. In parentheses are the errors in the final digits of the corresponding quantity. HD π [mas] 319 6870 7908 11413 15165 30422 31295 35242 74873 75654 84123 87271 91130 105058 106223 110377 110411 111005 111604 120500 120896 125162 130767 142703 153747 154153 168740 170680 183324 192640 193281 198160 204041 210111 216847 221756 12.45(0.74) 10.30(0.61) 11.41(0.98) 13.37(0.64) 8.51(0.89) 17.40(0.68) 27.04(0.94) 13.32(0.98) 16.38(1.16) 12.82(0.58) 9.09(0.90) 6.80(0.88) 13.33(0.76) 5.32(1.04) 9.10(0.86) 14.60(0.80) 27.10(0.82) 5.75(1.01) 8.43(0.73) 6.97(0.90) 4.87(1.15) 33.58(0.61) 7.82(0.78) 18.89(0.78) 5.32(0.98) 15.16(1.11) 14.03(0.69) 15.30(0.81) 16.95(0.87) 24.37(0.55) 4.58(1.59) 13.67(1.16) 11.46(0.99) 12.70(0.89) 6.76(0.67) 13.97(0.63) d [pc] 80(5) 97(6) 88(8) 75(4) 118(12) 57(2) 37(1) 75(6) 61(4) 78(4) 110(11) 147(19) 75(4) 188(37) 110(10) 68(4) 37(1) 174(31) 119(10) 143(19) 205(48) 30(1) 128(13) 53(2) 188(35) 66(5) 71(4) 65(3) 59(3) 41(1) 218(76) 73(6) 87(8) 79(6) 148(15) 72(3) µα [mas yr−1] µδ [mas yr−1] +70.01(0.94) +120.99(0.51) −5.54(1.04) −47.97(0.51) +35.50(1.10) −4.04(0.48) −34.71(0.41) −12.65(0.59) −37.51(0.50) −4.30(0.57) −11.70(0.74) +18.19(0.60) +40.08(1.04) −128.37(0.65) +16.63(0.63) −30.42(0.99) −51.06(0.70) −65.20(1.25) +41.13(0.49) −63.91(0.42) −86.15(0.43) −20.74(1.11) +3.33(0.86) +5.01(0.49) +6.89(0.42) +17.10(0.60) −9.80(0.84) +0.66(0.81) +19.29(0.55) −67.32(1.04) −106.08(0.78) +0.71(0.48) −89.51(0.48) +82.62(0.80) −43.12(0.86) −6.92(0.58) +21.83(0.56) −88.59(0.67) +10.97(0.68) −26.63(0.70) −18.65(1.14) +3.00(0.78) −187.42(0.52) +159.01(0.43) −46.63(0.69) −6.87(0.62) +71.62(0.83) −305.20(0.67) +14.08(0.68) +25.04(0.61) +0.57(0.59) −101.87(0.49) −6.98(0.84) −21.43(0.64) −33.35(0.43) −0.51(0.72) +68.67(0.45) +69.22(0.46) −4.32(1.54) −0.75(1.07) −49.38(0.92) +83.04(0.86) +50.00(1.18) +13.84(0.77) +24.90(0.48) +13.66(0.86) +12.33(0.40) +16.85(0.43) −17.71(0.45) −45.77(0.42) +6.55(0.99) +11.40(0.90) RV [km s−1] −10.7(9.9) +12.4(3.0) +13.2(3.0) +7.2(5.6) +33.5(2.1) +16.5(3.5) +12.9(1.7) +9.0(3.0) +23.3(3.0) +9.4(1.1) +16.3(3.0) +0.1(3.0) −14.5(3.0) −7.2(3.0) −15.4(3.7) +3.0(8.5) −12.7(11.5) +0.6(3.0) −17.1(4.7) +7.1(3.3) −24.4(3.0) −0.3(4.0) −14.0(3.0) +15.9(5.4) −6.3(0.5) −32.6(3.0) −21.1(3.0) −34.6(3.5) +12.0(3.0) −18.1(1.1) +2.0(2.7) −16.0(3.0) −15.1(26.7) −4.4(1.1) −3.0(1.4) +13.4(0.5) U [km s−1] V [km s−1] −18.1(1.8) −39.2(2.6) +9.2(1.0) +14.3(0.8) −32.4(1.8) −12.4(1.8) −6.0(1.6) −9.7(2.7) −23.0(2.4) −28.0(0.8) −16.0(2.3) +0.4(1.5) +11.6(1.5) −5.5(1.9) −32.8(2.9) −29.0(1.9) +18.7(1.5) −27.1(5.3) −46.7(3.7) −15.0(2.4) −25.0(3.7) −34.6(0.5) −20.5(2.0) +23.5(4.8) −3.2(1.1) −28.6(2.9) −36.3(2.6) −31.8(3.4) +14.1(2.3) −22.7(0.4) +4.2(2.5) −34.9(2.4) −25.1(13.3) −8.5(0.7) −12.6(1.2) +7.7(0.5) −25.5(1.8) −41.7(2.8) −11.3(1.1) +6.5(2.4) −8.6(1.4) −6.4(2.1) −23.9(0.7) +8.0(1.1) −22.9(1.4) −5.3(1.2) −45.5(4.4) +3.3(1.5) +5.6(0.4) −5.3(1.3) −7.3(1.9) −18.7(2.5) − 1.6(3.1) −23.6(3.1) −19.4(2.4) −6.4(1.9) −5.5(3.0) −5.1(1.7) −24.4(2.1) +4.0(0.7) +14.6(1.8) +17.5(1.0) −17.5(1.9) −15.2(0.9) +0.7(1.9) −12.4(1.1) −1.0(1.3) −8.6(2.3) −6.7(17.9) +7.7(0.7) +5.3(0.7) +10.9(0.6) W [km s−1] +4.9(9.7) −0.3(2.3) −14.5(3.0) −9.8(5.1) −20.8(2.1) −10.2(2.2) −10.8(0.8) −8.7(1.2) −7.9(2.1) −8.2(0.9) +8.4(2.3) +2.5(2.3) −9.4(2.6) −8.3(2.8) −20.9(3.7) +1.6(8.1) −16.2(11.0) −2.6(3.5) −19.3(4.8) +13.0(2.9) −16.7(2.6) −2.6(3.6) −0.7(3.0) −9.2(2.6) +2.8(1.9) +3.0(0.5) −2.8(1.2) +1.3(0.4) −5.7(0.5) −4.2(0.3) +2.5(2.3) −9.3(2.5) −3.4(14.8) +0.7(0.9) −6.1(1.5) −16.7(0.6) • The mean abundance of Na is solar, but the star-to-star scatter is about ±1 dex • The star-to-star scatter is twice as large as for a com- parable sample of normal stars Otherwise they find no, or only a poor, correlation of individ- ual abundances with astrophysical parameters such as the effective temperature, surface gravity, projected rotational velocity, age and pulsational period. We have used the individual abundances published by Venn & Lambert (1990), Sturenburg (1993), Holweger & Rentzsch-Holm (1995), Chernyshova et al. (1998), Heiter et al. (1998), Paunzen et al. (1999a,b), Kamp et al. (2001), Solano et al. (2001), Heiter (2002) and Andrievsky et al. (2002) for members of the λ Bootis group. The individ- ual values were weighted according to the errors listed in the references. For our further analysis we have used only objects for which an abundance of carbon or oxygen is avail- able, since those are key elements for the definition of the λ Bootis group. Values for stars of superficially normal type were taken from Adelman (1991, 1994, 1996), Adelman et al. (1991, 1997), Hill & Landstreet (1993), Hill (1995), Caliskan & Adelman (1997) and Varenne & Monier (1999). Let us recall that the membership of an object in the λ Bootis group is mainly based on spectroscopy at classifi- cation resolution. The only other approach is the definition of membership criteria in the UV region (Solano & Paun- zen 1999). No reference in the literature was found which describes membership criteria based on detailed abundance analysis in the optical region. If we compare the status of other chemically peculiar stars of the upper main sequence then a similar situation is evident (Wolff 1983; Cowley 1995). Objects have been classified as being chemically peculiar but their individual elemental abundances differ widely. No at- tempt has so far been made to define the membership of the classical CP stars to a sub-class by detailed abundances alone (Preston 1974). The λ Bootis group is unusual in this respect since it shows strong underabundances, not found for any other group, of most heavier elements. The only exceptions are intermediate and true Population II-type objects and 10 E. Paunzen, I.Kh. Iliev, I. Kamp, I.S. Barzova 2.0 1.5 1.0 0.5 0.0 -0.5 -1.0 -1.5 3.5 3.0 2.5 2.0 1.5 1.0 0.5 0.0 -0.5 -1.0 ] e F / C [ ] a C / C [ 1.5 1.0 0.5 0.0 -0.5 -1.0 -1.5 -2.0 2.0 1.5 1.0 0.5 -1.5 -1.0 -0.5 0.5 1.0 1.5 0.0 [Na] ] a N / C [ -2.5 -2.0 -1.5 -1.0 -0.5 0.0 0.5 1.0 [Fe] ] g M C / [ 0.0 -0.5 -1.0 -1.5 -4.0 -3.5 -3.0 -2.5 -2.0 -1.5 -1.0 -0.5 0.0 0.5 1.0 [Ca] -2.5 -2.0 -1.5 -0.5 0.0 0.5 -1.0 [Mg] Figure 6. Abundance ratios for the λ Bootis stars (filled circles); open circles denote normal-type stars from the literature. field blue stragglers, post-AGB and F-weak stars (Gray 1988, 1989; Jaschek, Andrillat & Jaschek 1989; Andrievsky, Chernyshova & Ivashchenko 1995; van Winckel, Waelkens & Waters 1995). Objects with very low surface gravities (post-AGB and Population II-type objects) are easily distin- guished even at classification resolution and will not be con- sidered in the following discussion. For the other groups the underabundances of the Fe-peak elements are rather mod- erate. But, more importantly, the abundances of the light elements C, N, O and S scale just like those of the heavier el- ements. We have therefore chosen to use [C/Z] versus [Z] dia- grams ([O/Z], [N/Z] and [S/Z] behave analogously) in order to investigate the behaviour of the group of λ Bootis stars. The field blue stragglers, intermediate Population II and F- weak types of stars all fall in the area around [C/Z] ≈ 0 in these diagrams. Figure 6 shows such diagrams for the elements Fe, Na, Ca and Mg for the stars on our program as well as for those of superficially normal types. From this Figure we are able to conclude: • All program stars are Population I-type objects with [C/Fe] ≫ 0 • The λ Bootis stars exhibit iron abundances that are sig- nificantly lower than those found for the superficially normal stars • There is a large overlap for all other heavier elements Cowley et al. (1982) have proposed that λ Bootis -type and other weak-line stars may arise from small (≈ 0.3 dex) abun- dance fluctuations in the interstellar medium. That might be true for a small fraction of the objects, but underabun- dances up to a factor of 100 can not be explained without The status of Galactic field λ Bootis stars in the post-Hipparcos era 11 1.5 1.0 0.5 0.0 -0.5 -1.0 -1.5 ) s i t o o B ( ] a N [ 1.0 1.5 2.0 2.5 3.0 log [Na] (ISM, arbitrary units) Figure 7. Correlation between the sodium abundances for 13 λ Bootis stars and the values, taken from Welsh et al. (1998), for the surrounding environment. The latter values are scaled in three steps (a value of three denotes the highest density). some other mechanism such as diffusion, accretion or mass- loss. A very intriguing fact is that eight program stars ex- hibit magnesium abundances which range from −0.52 to −0.13 dex, which seems to contradict the classification based on moderate-resolution spectroscopy. One of the most im- portant classification criterion is the moderate to extreme weakness of the Mg II 4481 A line (Gray 1988). What can ac- tually cause that discrepancy? Slettebak, Kuzma & Collins (1980) have shown that the equivalent width of the line decreases with increasing rotation for models later than A0. The same fact was also described by Abt & Mor- rell (1995). A much stronger effect was found for Hγ. That means that a rapidly rotating star will be classi- fied much later according to Hγ than on the basis of Mg II 4481 A. Taking a rapidly rotating A5 type star one would classify it as hF1mA7 or metal-weak. However, the three objects with the highest magnesium abundances of our sample are indeed the fastest rotators: HD 193256 (250 kms−1; −0.15 dex), HD 170680 (205 kms−1; −0.20 dex) and HD 198160 (200 kms−1; −0.13 dex). But there is no overall correlation for the whole sample, fast rota- tors do also exhibit rather strong underabundances (e.g. HD 111604) and vice versa. Otherwise the fast rotators are not outstanding in any respect. i.e. One observational fact is the wide range of abundances (−1.3 to +1.2 dex) for sodium (Table 5 and Fig. 6) first indicated in the work of Sturenburg (1993). In the recent literature no explanation of the variability has been given. Sodium is the only element for which such a behaviour has been detected so far. Furthermore, predictions for that el- ement have not been discussed so far within any proposed theory (Turcotte & Charbonneau 1993). In an effort to shed some light on the subject, we have investigated whether a correlation can be found between the individual sodium abundances and the density of sodium in the surrounding interstellar medium. Welsh, Crifo & Lallement (1998) pub- lished the local distribution of interstellar Na I within 250 pc of the Sun. They used all published absorption densities and the distances derived from the Hipparcos satellite. The den- sities were scaled in three steps (a value of three denotes the highest density) and plotted for different Galactic coor- dinates and distances from the Sun. We have selected mem- bers of the λ Bootis group whose sodium abundances are known and for which nearby data points are available in the maps of Welsh et al. (1998). We have also checked more recent references such as Sfeir et al. (1999) and Vergely et al. (2001) which give essentially the same results. In total, thirteen objects fulfill the requirements: HD 319, HD 31295, HD 74873, HD 84123, HD 125162, HD 183324, HD 192640, HD 193256, HD 193281, HD 198160, HD 204041, HD 210111 and HD 221756. We have to emphasize that not a single sodium abun- dance for the line of sight to any λ Bootis star has yet been measured. Bohlender et al. (1999) reported a few detections of such features, e.g. for HD 319, HD 192640 and HD 221756, but they did not derive quantitative densities or abundances. Figure 7 shows the correlation of the individual sodium abundances with the absorption densities for the local in- terstellar medium (ISM hereafter). Since there is a linear correlation visible it suggests that there is an interaction (e.g. accretion) between the stars and their environments at some stage of stellar evolution. The correlation does not reflect any age dependency, since objects are included with 0.33 < trel < 1.01. This rather small sample is reasonably rep- resentative of the whole sample of program stars in terms of its distribution in trel and effective temperature. Unfortu- nately it is not possible to include any data for superficially normal objects in Fig. 7 since no sodium abundances for bright field stars are available. The only data are those pub- lished by Varenne & Monier (1999; Fig. 6) for members of the Hyades. From our current analysis two main questions arise for the λ Bootis phenomenon: How many mechanisms are in- volved? What are the observational constraints? It is clear that there is one mechanism which produces the observed pattern throughout the whole MS lifetime for stars between late B and early F types. We are not able to decide if it is 'internal' or 'external' but we have some hints about it. It is highly improbable that one mechanism works for early evolutionary stages and an independent second one at very late stages producing the same absolute abundance pattern. That seems to be supported by the non-existence of a cor- relation between the iron abundance and age (Fig. 8). The abundance of sodium for the stellar atmosphere is corre- lated with that of the local ISM. There are two possible explanations for that: 1) the atmospheric abundance resem- bles the one from the cloud in which the star was born, or 2) it currently interacts with the local ISM. If we believe in the first interpretation then all other abundances ought also l 12 E. Paunzen, I.Kh. Iliev, I. Kamp, I.S. Barzova ] e F [ ] e F [ -0.4 -0.6 -0.8 -1.0 -1.2 -1.4 -1.6 -1.8 -2.0 -2.2 -2.4 -0.4 -0.6 -0.8 -1.0 -1.2 -1.4 -1.6 -1.8 -2.0 -2.2 -2.4 8.0 8.2 8.4 8.6 log age 8.8 9.0 9.2 0.3 0.4 0.5 0.6 0.8 0.9 1.0 1.1 0.7 trel 55 50 45 40 35 30 25 20 15 10 5 0 55 50 45 40 35 30 25 20 15 10 5 0 55 50 45 40 35 30 25 20 15 10 5 0 e g a t n e c r e P e g a t n e c r e P e g a t n e c r e P Bootis l normal stars U V W -60 -50 -40 -30 -20 -10 0 10 20 30 Velocity [kms-1] Figure 8. No correlation is found between iron abundance and age. Figure 9. The histograms of the Galactic space-velocity com- ponents (U, V and W ) for the λ Bootis and normal-type samples. The distributions agree at significance levels of 72% (U ), 74% (V ) and 97% (W ). to resemble those in the local ISM. That seems not to be the case, since many other 'normal' stars located within the same ISM clouds show no significant elemental underabun- dances at all. So we are left with the picture of an interaction between the star and its environment. 6 THE SPACE MOTIONS OF λ BO OTIS STARS In order to study the kinematics of nearby stars, one needs to calculate the Galactic space-velocity components (U , V and W ), given the star's proper motion, parallax (both listed in the Hipparcos catalogue) and radial velocity. Here, the formulae and error propagation are taken from the Hippar- cos documentation as well as from Johnson & Soderblom (1987). It was chosen to calculate heliocentric Galactic veloc- ity components, which can easily be corrected for the solar motion. A right-handed coordinate system for U , V and W was used, so they are positive in the directions of the Galac- tic centre, Galactic rotation and the North Galactic Pole. The 'standard solar motion' in this system would be (+10.4, +14.8, +7.3) taken from Mihalas & Routly (1968). The authors followed the approach of the Hipparcos consortium who discussed the calculation, transformation matrix and error propagation. We would like to give only a short description of the error estimation. The general equa- tion for the variance of a function of several variables is used for this purpose. That formula is true only if the covariances are zero (the errors are uncorrelated). That assumption is fulfilled since the two proper-motion components are mea- sured independently. Furthermore it is assumed that only the radial velocity, proper motions and parallax contribute to the error distribution. For nearby stars (about 25 pc) an error of the proper motions of about ±3 mas yr−1 (most of the Hipparcos measurements are a factor three better than that) corresponds to an uncertainty in the transverse mo- tion of only about 0.5 km s−1; it is clear that the errors of the radial velocities are much more significant. The radial velocities for our program stars are listed in Table 6. The following references were used for that Table: Wilson (1953), Evans (1967), Batten, Fletcher & MacCarthy (1989), Barbier-Brossat, Petit & Figon (1994), Duflot et al. (1995), Hauck et al. (1995, 1998), Fehrenbach et al. (1996, 1997), Nordstrom et al. (1997) and Grenier et al. (1999a,b). The mean and the error of the mean were calculated with- out weighting the individual measurements. From Table 6 only three stars (HD 319, HD 110411 and HD 204041) show evidence for variable radial velocities but no prominent pho- tometric or spectroscopic variability has been detected so far. Moreover, Table 6 lists the kinematic data for the pro- gram λ Bootis stars from the Hipparcos database. The in- vestigated sample is limited in distance (≈ 220 pc) because of the lack of Hipparcos data and/or radial velocities for more distant objects. No star exceeds V = 50 km s−1 and the sample is very homogeneously distributed. A similar conclusion is given by G´omez et al. (1998a) for classical CP The status of Galactic field λ Bootis stars in the post-Hipparcos era 13 stars as well as by G´omez et al. (1998b) for a small sample of λ Bootis stars. We have compared the velocities of stars in our λ Bootis sample with the results of Caloi et al. (1999). They studied the relationship between the kinematics, age and heavy-element content for the solar neighbourhood, using the Hipparcos data. For V < −40 km s−1, U > +60 km s−1 and W > +30 km s−1, the minimum stellar age is about 2 Gyr. Only two of the λ Bootis stars would meet the V criterion, HD 6870 (V = −41.7(2.8) km s−1) and HD 84123 (V = −45.4(4.4) km s−1), but their U and W velocities (like those of all other stars in the sample) are well below the limits given. Overall, the velocities are typical of true Pop- ulation I objects. The comparison with the sample of normal stars (Fig. 9) shows very good agreement between the two distributions. They agree at significance levels of 72% (U ), 74% (V ) and 97% (W ). Notice that Faraggiana & Bonifacio (1999) made a simi- lar analysis for a different sample of candidate λ Bootis stars (see Section 4 therein). They only investigated the parame- ter (U 2 + W 2)1/2 which is a measure of the kinematic energy not associated with Galactic rotation. They come to the con- clusion that all of their program stars are qualified as disk members. Since they give no individual values for the space motions and radial velocities, we are not able to compare our results with theirs. 7 CONCLUSIONS We have used all currently available photometric data as well as Hipparcos data to determine astrophysical parame- ters such as the effective temperatures, surface gravities and luminosities. As a next step, masses and ages were calibrated within appropriate post-MS evolutionary models. Further- more, Galactic space motions were calculated with the help of radial velocities from the literature. The comparison with already published results shows good agreement of the de- rived parameters. All results were compared with those of a test sample of normal-type objects in the same spectral range chosen in order to match the (B − V )0 distribution of the λ Bootis group. From a comprehensive statistical anal- ysis we conclude: • The standard photometric calibrations within the John- son UBV, Stromgren uvbyβ and Geneva 7-colour systems are valid for this group of chemically peculiar stars. • The group of λ Bootis stars consists of true Population I objects which can be found over the whole area of the MS with a peak at a rather evolved stage (≈ 1 Gyr). That is in line with the distribution of the test sample. • The λ Bootis type group is not significantly distinct from normal stars except, possibly, by having slightly lower temperatures and masses for trel > 0.8. The v sin i range is rather narrow throughout the MS with a mean value of about 120 kms−1. • There seems to exist a non-uniform distribution of ef- fective temperatures for group members with a large pro- portion of objects (more than 70%) cooler than 8000 K. • It seems that objects with peculiar hydrogen-line pro- files are preferentially found among later stages of stellar evolution. • No correlation of age with elemental abundance or pro- jected rotational velocity has been detected. • A comparison of the stellar Na abundances with nearby IS sight lines hints at an interaction between the λ Bootis stars and the ISM. • There is one single mechanism responsible for the ob- served phenomenon which produces moderate to strong underabundances working continuously from very early (10 Myr) to very late evolutionary stages (2.5 Gyr). It pro- duces the same absolute abundances throughout the MS life- time for 2% of all luminosity class V objects with effective temperatures from 10500 K to 6500 K. • The current list of stars seem to define a very homo- geneous group, validating the proposed membership criteria in the optical and UV region. These rather strict observational results for a significant number of λ Bootis stars will need to be taken into account in future work on theories and models trying to explain the phenomenon. The constraints presented here will help considerably to reduce the number of free parameters in the models and finally to provide a critical test for them. Acknowledgements. This work was partly supported by the Fonds zur Forderung der wissenschaftlichen Forschung, project P14984 and it is dedicated to G. Derka who died during its preparation. Use was made of the SIMBAD database, operated at CDS, Strasbourg, France and the GCPD database, operated at the Institute of Astronomy of the University of Lausanne. REFERENCES Abt H.A., Morrell N.I., 1995, ApJS, 99, 135 Adelman S.J., 1991, MNRAS, 252, 116 Adelman S.J., 1994, MNRAS, 271, 355 Adelman S.J., 1996, MNRAS, 280, 130 Adelman S.J., Bolcal C., Hill G., Kocer D., 1991, MNRAS, 252, 329 Adelman S.J., Caliskan H., Kocer D., Bolcal C., 1997, MNRAS, 288, 470 Andrievsky S.M., Chernyshova I.V., Ivashchenko O.V., 1995, A&A, 297, 356 Andrievsky S.M., Chernyshova I.V., Korotin S.A., et al., 2002, A&A (in preparation) Andrillat Y., Jaschek C., Jaschek M., 1995, A&A, 299, 493 Arenou F., Grenon M., G´omez A., 1992, A&A 258, 104 Barbier-Brossat M., Petit M., Figon P., 1994, A&AS, 108, 603 Batten A.H., Fletcher J.M., MacCarthy D.G., 1989, Publ. Do- minion Astrophys. Obs., 17 Bohlender D.A., Gonzalez J.F., Matthews J.M., 1999, A&A, 350, 553 Caliskan H., Adelman S.J., 1997, MNRAS, 288, 501 Caloi V., Cardini D., D'Antona F., Badiali M., Emanuele A., Mazzitelli I., 1999, A&A, 351, 925 Cayrel de Strobel G., 1996, A&AR, 7, 243 Chen B., Vergely J.L., Valette B., Carraro G., 1998, A&A, 336, 137 Chernyshova I.V., Andrievsky S.M., Kovtyukh V.V., Mkrtichian D.E., 1998, Contributions of the Astronomical Observatory Skalnat´e Pleso, Vol. 27, No. 3, p. 332 Claret A., 1995, A&AS, 109, 441 Cowley C.R., 1995, PASPC, 81, 467 Cowley C.R., Sears R.L., Aikman G.C.L., Sadakane K., 1982, ApJ, 254, 191 14 E. Paunzen, I.Kh. Iliev, I. Kamp, I.S. Barzova Drilling J.S., Landolt A.U., 2000, in A.N. Cox, ed, Allen's Astro- Paunzen E., Andrievsky S.M., Chernyshova I.V., Klochkova V.G., physical Quantities, 4th edition, Springer Verlag, p. 388 Panchuk V.E., Handler G., 1999b, A&A, 351, 981 Duflot M., Fehrenbach Ch., Mannone C., Burnage R., Genty V., 1995, A&AS, 120, 177 Evans D.S., 1967, Catalogue of stellar radial velocities, IAU Symp. 30, 5 Faraggiana R., Bonifacio P., 1999, A&A, 349, 521 Faraggiana R., Gerbaldi M., 1998, Contributions of the Astro- nomical Observatory Skalnat´e Pleso, Vol. 27, No. 3, p. 413 Perryman M.A.C., Høg E., Kovalevsky J., Lindegren L., Turon C., 1997, The Hipparcos and Tycho Catalogues, ESA SP-1200 Preston G.W., 1974, ARA&A, 12, 257 Rees D.G., 1987, Foundations of Statistics, Chapman & Hall, Lon- don Sandage A., 1972, ApJ, 178, 1 Schaller G., Schaerer D., Meynet G., Maeder A., 1992, A&AS, Fehrenbach Ch., Duflot M., Genty V., Amieux G., 1996, Bull. 96, 269 Inform. CDS 48, 11 Sfeir D.M., Lallement R., Crifo F., Welsh B.Y., 1999, A&A, 346, Fehrenbach Ch., Duflot M., Mannone C., Burnage R., Genty V., 785 1997, A&AS, 124, 255 Garrison R.F., Gray R.O., 1994, AJ, 107, 1556 G´omez A.E., Luri X, Grenier S., et al., 1998a, A&A, 336, 953 G´omez A.E., Luri X, Sabas V., et al., 1998b, Contributions of the Astronomical Observatory Skalnat´e Pleso, Vol. 27, No. 3, p. 171 Gray R.O., 1988, AJ, 95, 220 Gray R.O., 1989, AJ, 98, 1049 Gray R.O., Corbally C.J., 1993, AJ, 106, 632 Gray R.O., Corbally C.J., 1998, AJ, 116, 2530 Gray R.O., Corbally C.J., 1999, AAS, 195, 4702 Gray R.O., Garrison R.F., 1987, ApJS, 65, 581 Gray R.O., Garrison R.F., 1989a, ApJS, 69, 301 Gray R.O., Garrison R.F., 1989b, ApJS, 70, 623 Grenier S., Burnage R., Faraggiana R., et al., 1999a, A&AS, 135, 503 Grenier S., Baylac M.O., Rolland L., et al., 1999b, A&AS, 137, 451 Slettebak A., Kuzma T.J., Collins G.W., 1980, ApJ, 242, 171 Solano E., Paunzen E., 1999, A&A, 348, 825 Solano E., Paunzen E., Pintado O.I., Varela J., 2001, A&A, 374, 957 Sturenburg S., 1993, A&A, 277, 139 Turcotte S., Charbonneau P., 1993, ApJ, 486, 536 Uesugi A., Fukuda I., 1982, Revised Catalogue of Stellar Rota- tional Velocities, Department of Astronomy, Kyoto Univ., Ky- oto van Winckel H., Waelkens C., Waters L.B.F.M., 1995, A&A, 297, 25 Varenne O., Monier R., 1999, A&A, 351, 247 Venn K.A., Lambert D.L., 1990, ApJ, 363, 234 Vergely J.-L., Freire Ferrero R., Siebert A., Valette B., 2001, A&A, 366, 1016 Welsh B.Y., Crifo F., Lallement R., 1998, A&A, 333, 101 Wilson R.F., 1953, General Catalogue of stellar radial velocity, Carnegie Inst. of Washington, Publ. 601 Hakkila J., Myers J.M., Stidham B.J., Hartmann D.H., 1997, AJ, Wolff S.C., 1983, The A-type Stars: Problems and Prospectives, NASA-SP 463 114, 2043 Hauck B., Ballereau D., Chauville J., 1995, A&AS, 109, 505 Hauck B., Ballereau D., Chauville J., 1998, A&AS, 128, 429 Heiter U., 2000, PASP, 112, 1509 Heiter U., 2002, A&A, 381, 959 Heiter U., Kupka F., Paunzen E., Weiss W.W., Gelbmann M., 1998, A&A, 335, 1009 Heiter U., Weiss W.W., Paunzen E., 2002, A&A, 381, 971 Hill G.M., 1995, A&A, 294, 536 Hill G.M., Landstreet J.D., 1993, A&A, 276, 142 Holweger H., Rentzsch-Holm I., 1995, A&A, 303, 819 Holweger H., Hempel M., Kamp I., 1999, A&A, 350, 603 Iliev I.K., Barzova I.S., 1993, Ap&SS, 208, 277 Iliev I.K., Barzova I.S., 1995, A&A, 302, 735 Iliev I.K., Paunzen E., Barzova I.S., et al., 2002, A&A, 381, 914 Jaschek M., Andrillat Y., Jaschek C, 1989, A&A, 218, 180 Johnson D.R.H., Soderblom D.R., 1987, AJ, 93, 864 Kamp I., Iliev I.Kh., Paunzen E., et al., 2001, A&A, 375, 899 King J.R., 1994, MNRAS, 269, 209 Kobi D., North P., 1990, A&AS, 85, 999 Koen C., 1992, MNRAS, 256, 65 Kunzli M., North P., Kurucz R.L., Nicolet B., 1997, A&AS, 122, 51 Mihalas D., Routly P.M., 1968, Galactic Astronomy, Freeman & Co., San Francisco, p. 101 Moon T.T., Dworetsky M.M., 1985, MNRAS, 217, 305 Morel P., 1997, A&AS, 124, 597 Napiwotzki R., Schonberner D., Wenske V., 1993, A&A, 268, 653 Nordstrom B., Stefanik R.P., Latham D.W., Andersen J., 1997, A&AS, 126, 21 Oudmaijer R.D., Groenewegen M.A.T., Schrijver H., 1998, MN- RAS, 294, L41 Palla F., Stahler S.W., 1993, ApJ, 418, 414 Paunzen E., 1997, A&A, 326, L29 Paunzen E., 2001, A&A, 373, 633 Paunzen E., Gray R.O., 1997, A&AS, 126, 407 Paunzen E., Kamp I., Iliev I.Kh., et al., 1999a, A&A, 351, 981
astro-ph/0608470
2
0608
2007-05-08T18:50:48
A three stage model for the inner engine of Gamma Ray Burst: Prompt emission and early afterglow
[ "astro-ph" ]
We propose a new model within the ``Quark-nova'' scenario to interpret the recent observations of early afterglows of long Gamma-Ray Bursts (GRB) with the Swift satellite. This is a three-stage model within the context of a core-collapse supernova. Stage 1 is an accreting (proto-) neutron star leading to a possible delay between the core collapse and the GRB. Stage 2 is an accreting quark-star, generating the prompt GRB. Stage 3, which occurs only if the quark-star collapses to form a black-hole, consists of an accreting black-hole. The jet launched in this accretion process interacts with the ejecta from stage 2, and could generate the flaring activity frequently seen in X-ray afterglows. This model may be able to account for both the energies and the timescales of GRBs, in addition to the newly discovered early X-ray afterglow features.
astro-ph
astro-ph
A three stage model for the inner engine of Gamma Ray Burst: Prompt emission and early afterglow Department of Physics and Astronomy, University of Calgary, 2500 University Drive NW, Jan Staff and Rachid Ouyed Calgary, AB T2N 1N4, Canada and Manjari Bagchi 1 Tata Institute of Fundamental Research, Homi Bhaba Road, Colaba, Mumbai 400005, India ABSTRACT We propose a new model within the "Quark-nova" scenario to interpret the recent observations of early afterglows of long Gamma-Ray Bursts (GRB) with the Swift satellite. This is a three-stage model within the context of a core- collapse supernova. Stage 1 is an accreting (proto-) neutron star leading to a possible delay between the core collapse and the GRB. Stage 2 is an accreting quark-star, generating the prompt GRB. Stage 3, which occurs only if the quark- star collapses to form a black-hole, consists of an accreting black-hole. The jet launched in this accretion process interacts with the ejecta from stage 2, and could generate the flaring activity frequently seen in X-ray afterglows. This model may be able to account for both the energies and the timescales of GRBs, in addition to the newly discovered early X-ray afterglow features. Subject headings: gamma rays: bursts, stars: evolution, (stars:) supernovae: general 1. Introduction With the launch of the Swift satellite (Gehrels et al. 2004), observations of early X-ray afterglows from gamma ray bursts (GRBs) have become possible (sometimes as early as 80 1Department of Physics and Astronomy, University of Calgary, Canada AB T2N 1N4 -- 2 -- seconds after the GRB trigger). This led to some surprising observations, most notably, the existence of one or more (sometimes giant) flares (first discussed by Burrows et al. 2005) in the early X-ray afterglow (after a few hundred seconds to several thousand seconds) whose rapid variability has been interpreted as the inner engine being active much longer than the duration of the GRB itself (see for instance Zhang et al. 2006). Figure 1 shows a canonical X-ray afterglow light curve. The early X-ray afterglow often has a very steep power-law decay, lasting up to about a thousand seconds, followed by a flattening of the light curve, which lasts for 104 − 105 seconds. This is often overlayed with flares and bumps in the light curve. Thereafter, a "classical" decaying afterglow is seen, as it was known from the pre-Swift era. It should be noted that not every bursts exhibits all of these features. The flare(s) and the flattening of the light curve are likely due to extended inner engine activity, and, this extended inner engine activity is what we focus on in this paper. The inner engine for the GRB in our model is an accreting quark-star formed shortly (hours) after the core collapse in a massive star. This process operates in three steps: i) the quark nova (Ouyed et al. 2002), where the core is converted to quark matter resulting in mass ejection, ii) fall back material from the supernova together with some of the matter ejected during the quark nova can form an accretion disk around the quark-star. Accretion onto the quark-star launches a jet that will overtake the shell ejected by the quark nova and is a possible location for the prompt X-ray emission. Internal shocks within the jet produce the GRB, iii) if enough matter is accreted onto the quark-star, it will collapse and form a black-hole. Continued accretion onto the black-hole can lead to an ultra-relativistic jet. Interactions between this jet and the jet from the quark-star create shocks which lead to the flaring activity frequently seen in X-ray afterglows. In this paper we assume that the gamma-ray emission is produced by internal shocks. The afterglow is produced when the merged shell creating the gamma-ray emission interacts with the external medium in an external shock. It should be emphasized that quark stars are hypothetical objects, that have not been observed. They have been discussed theoretically for more than thirty years (see for instance Itoh 1970). Two requirements are needed for quark stars to form. First the strange quark matter (SQM) hypothesis must hold true. This hypothesis states that SQM has a lower energy state than 56F e, thus being the ground state of hadronic matter. Quark stars are usually thought to form from neutron stars whose central density increases above a critical density. Only the more massive neutron stars are thought to be able to reach this density (Staff et al. 2006). Second, the SQM EOS must support a rather massive quark star (the mass will probably not change much in converting from a neutron star to a quark star). Optionally, a quark star can form directly in the collapse of the iron core in massive stars, -- 3 -- in which case the second requirement might be relaxed. In § 2 we present our model for an inner engine for GRBs in three stages, followed by a discussion of the energies and timescales of the inner engine in § 3. Then in § 4 we discuss what happens when the jet from the black-hole and the quark-star interacts followed by a discussion of the formation of the different features in the X-ray afterglow in § 5. We devote § 6 to correlations of features seen in the early afterglow light curve, and then in § 7 we apply our model to different observed GRBs. Then in § 8 we discuss our model and finally we summarize our model in § 9. 2. The three stages of activity In this section, we discuss the three stages in our model for the inner engine for long GRBs (see Fig. 2). These stages involve a neutron star phase, a quark star phase, and a black hole phase. The goal is to explain features (e.g. flares and flattening in early X-ray afterglow) observed in long GRBs as a result of interactions between the neutron star jet, quark star jet, and the black hole jet. In essence the interaction between these jets is more generic and could be applicable to other multi-stage models involving other engines. 2.1. Stage 1: Neutron star as the inner engine The first stage in our model is the formation of a (proto-) neutron star from the collapse of the core in the supernova. Several suggestions on how to generate GRBs from a newly formed neutron star exist (e.g. Usov 1992; Kluzniak & Ruderman 1998; Wheeler et al. 2000; Ruderman et al. 2000; Dai et al. 2006). More elaborate models involve in addition a transi- tion to a black hole. Here we bring in an intermediate stage where the neutron star collapses to a quark star (i.e. a quark nova) first. Our model is effectively a two-stage model since we will be mostly focusing on the interaction between the quark star ejecta and the black hole ejecta. We note however that the neutron star phase can lead to a short delay between the collapse of the iron core and the GRB, which in our model occurs during the last two stages which we describe next. 2.2. Stage 2: Quark star as the inner engine The neutron star will accrete fall-back material from the exploding star until its den- sity reaches deconfinement density, whereby the star is converted to a strange quark-star. -- 4 -- Alternatively, if the density of the compact object left behind in the core collapse is above deconfinement density, the quark-star can be formed directly in the collapse. The accreting quark-star will produce the GRB (Ouyed et al. 2005, hereafter ORV), which is discussed later. The lifetime of the neutron star will determine whether the GRB and supernova will be seen as simultaneous (or nearly simultaneous) events, or as temporally separated events. The timescale for the conversion of the neutron star can vary dramatically, depending on the accretion rate, the spin-down rate, and how close the state of the neutron star is to a deconfined state. Observations indicate that the core collapse and the GRB are normally almost simultaneous events (Della Valle 2006), which means that stage 1 in our model is usually short or not present at all. The quark-star will be surrounded by an accretion disk, due to fallback material from the supernova, as well as quark nova material. As the star accretes matter, it will be heated up. The decay of Goldstone bosons producing photons is the main cooling mechanism in this phase (Vogt et al. 2004). For temperatures above about 7.7 MeV these photons can escape the star, whereas for lower temperature they will be absorbed by the star (see ORV). If the star heats up above 7.7 MeV, the emitted photons will interact with accreting material and eject it. This halts accretion until the star has cooled down below 7.7 MeV again. This way, episodes of accretion and ejection will occur. The accreting material will follow the star's magnetic field lines towards the magnetic pole of the star. Hence, most of the ejected material will be ejected from the polar regions, and it will be collimated by the magnetic field, i.e. it is a jet (ORV). We assume an accretion m ∼ 10−4M⊙ s−1 onto the quark star (similar to what is expected for neutron stars, rate of Fryer et al. 1996). Explaining the physical process that limits the accretion rate is beyond the scope of this work and is left for future work. If the accretion rate is higher, we suggest a black hole is formed quickly. This gives us a one stage model, as described in § 8.2.1. We find that for typical values of the QS magnetic field, B ∼ 1015 G1, the maximum accretion rate that can be channeled towards the polar cap2 to be m ≈ 10−3M⊙ s−1 (for higher accretion rates the Alfv´en radius lies inside the star). For B ∼ 1014 G, the accretion rate must be smaller than m ≈ 10−5M⊙ s−1 for the accretion to be channeled to the polar 1Recent work shows that 1015 G magnetic fields can be obtained during QS formation due to the response of quarks to the spontaneous magnetization of the gluons (e.g. Iwazaki 2005, and references therein). 2The lack of observational evidence of a precessing jet (Beloborodov et al. 2000) is naturally explained in our model since the magnetic field immediately after birth will align with the rotational axis of the star (Ouyed et al. 2006). -- 5 -- cap by the magnetic field, too low for the more energetic burst but suitable for lower energy bursts and very long duration bursts. Most photons will be emitted from the polar regions (where accretion heats up the star), and they will then interact with some of the accreting material (see Fig 3). The Lorentz factor of the ejected matter is Γ = ηQSmaccr/mejec, with ηQS ∼ 0.1 (Frank et al. 1992), maccr and mejec are the accreted and ejected mass respectively (ORV). In order to achieve Lorentz factors of the order 100, only a small fraction (10−3) of the accreting material can therefore be ejected. Such a small fraction can be realized since the cooling time (and therefore the period in which photons are emitted) is of the order microseconds, whereas the heating (accretion) time is of the order milliseconds (ORV). If we assume a steady accretion, and that the photons eject an amount of mass mejec ∼ m∆tem for each episode we find: ∆taccr ∆tem ∼ meject maccr ∼ 10−3, (1) ∆taccr and ∆tem being the time interval for accretion and ejection during one episode respec- tively. However, it is likely that only a fraction of this material is ejected, leading to higher Lorentz factors. This fraction might vary from episode to episode, giving rise to varying Lorentz factors in the outflow. Internal shocks created by colliding shells in the jet acceler- ate electrons that emit synchrotron radiation observed as gamma rays, as in Narayan et al. (1992). In order to produce merged shells with an internal energy of about 1050 erg (which is generally required to explain the energies involved in GRBs) and Lorentz factors of more than a hundred (to overcome the compactness problem; see e.g. Piran 1999), the shells must be about 10−6M⊙ or 1027 g each. If the Lorentz factors are too large (a few thousands), internal shocks will occur too late and external shocks occur before the internal shocks can take place. This sets the upper and lower limits on the Lorentz factors required for GRBs. For ηQS = 0.1, this means that maccr = 1030 − 1031g ∼ 10−3 − 10−2M⊙. A more likely scenario is that in each episode less mass is ejected, and several of these ejecta quickly merge to produce shells with mass ∼ 1027 g relaxing the constraints on maccr to lower values. These shell mergers will not be observed, as they occur while the jet is still inside the exploding star. The most efficient conversion of kinetic energy to internal energy is achieved when the masses of the shells are similar, and the ratio in Lorentz factor of the colliding shells are big, at least a factor two. The duration of the accretion process depends on the mass of the disk and the maximum temperature that the star can be heated to. ORV found that this process can last hundreds of seconds, and possibly thousands. We will assume that the accretion eventually settles into a steady state. If this accretion is not high enough to heat the star much above 7.7 -- 6 -- MeV, the ejection will be halted or very limited and the GRB comes to an end. However, the inner engine is still actively accreting, until the disk has been depleted; this is unless the star turns into a BH before disk depletion (see next subsection for more discussion). ORV found that 10% of the rest mass energy of the accretion disk being accreted can be used to power a jet. A quark-star can probably accrete up to 0.1M⊙ without collapsing to a black-hole. Hence 0.1M⊙c2 ≈ 2 × 1053 erg is the maximum jet energy powered by accretion onto a quark-star. We assume a radiative efficiency of 10% in shell collisions during the QS phase. This means that in our model only about 1% (or about 2 × 1051 erg) of the rest mass energy of the accretion disk surrounding the quark-star can be released as gamma rays in a GRB. Using a collimation angle of a few degrees (ORV), this corresponds to Eγ,iso ∼ 1054 erg. The ejected shells will collide with each other as explained, and be decelerated by the interstellar medium forming an external shock. Acceleration of electrons in this shock creates the afterglow (as in the internal -- external shocks model). If the outflow creating the afterglow is beamed with jet angle θjet, a jet break will be observed as Γ−1 > θjet (Rhoads 1999). This usually happens after about a day. 2.3. Stage 3: Black-hole as the inner engine For a given temperature and equation of state (EOS), a mass-radius curve for quark- stars has a maximum, i.e. there is a maximum mass for quark-stars. As is the case for the maximum mass of neutron stars, this maximum mass depends on the EOS which is still being studied. Harko & Cheng (2002) shows that quark stars whose EOS can be approximated by a linear function of the density has a Chandrasekhar limit based on degenerate quarks. This point denotes an instability, and so if there is accretion onto a star having this maximum mass, it collapses into a black hole. We note however, that thermally induced instabilities can drive the collapse to a BH before reaching the Chandrasekhar limit (for details see Bagchi et al. 2006). When the quark-star collapses into a black-hole, the accretion process changes dramat- ically. The accretion rate is of the order 0.01 to 10 M⊙ s−1 in a hyperaccretion disk around black holes (Popham et al. 1999), much higher than around quark stars where surface radi- ation and magnetospheric effects should in principle reduce the accretion rate compared to the accretion rate onto black holes. An ultrarelativistic jet is launched from the accretion process onto the black hole as in De Villiers et al. (2005). Interaction between this jet and the slower parts of the jet from the quark-star can generate flares seen in the early X-ray -- 7 -- afterglow and when this jet collides with the external shock, late time bumps will be seen. Internal shocks within the black hole jet itself can also occur, which likely would lead to flares in either X-ray or gamma-ray wavelengths. Since the duration of this jet is likely short, the width of these flares would also be short. This may add to the complexity of the observed light curve. The duration of the accretion/ejection process depends on the mass of the accretion disk after the quark-star to black-hole conversion and the accretion rate. The disk is unlikely to be more than a few solar masses, giving a maximum accretion time of about a hundred seconds. However as found by De Villiers et al. (2005), accretion onto a black-hole is a much faster process (about 1M⊙ s−1), which limits the duration of the accretion process to a few seconds. This we take as a typical timescale for the accretion rate onto black-holes for the rest of this paper. 3. Inner engines timescales and energies We can estimate the ratio between the duration of the inner engine in the black-hole stage and in the quark-star stage as: tBH tQS = mdisk,BH mBH mdisk,QS mQS =(cid:16) mQS mBH(cid:17)mdisk − mdisk,QS mdisk,QS = ζ m(cid:16) mdisk mdisk,QS − 1(cid:17) (2) where mdisk = mdisk,QS + mdisk,BH and ζ m parameterizes the ratio between the accretion rate onto a quark-star and a black-hole. As typical values we use the same accretion rate mQS ≈ 10−4M⊙ s−1 (Fryer et al. 1996, in fact the onto quark stars as for neutron stars, accretion rate depends on the maximum temperature the star is heated to as in ORV) and mBH ≈ 1M⊙ s−1 (De Villiers et al. 2005) giving ζ m ≈ 10−4, and find that the duration of the BH era is much shorter than the QS era. The ratio between the energy produced by the inner engine in the two stages can be found by: EBH EQS = ηBHmdisk,BH ηQSmdisk,QS = ζm mdisk − mdisk,QS mdisk,QS = ζm(cid:16) mdisk mdisk,QS For typical values we use ηBH ≈ 10−3 (mejec/macc = 10−5 was found in De Villiers et al. 2005, assuming Γ = 100 this gives ηBH = Γmejec/macc = 10−3), and ηQS = 10−1 (similar to the efficiency in neutron stars, Frank et al. 1992). This gives ζm = 10−2. Hence, most of the energy will be output during the QS era, unless almost all of the disk (> 99%) is accreted during the black-hole era. − 1(cid:17). (3) -- 8 -- We can now combine the two above expressions, to get a simple ratio: tBH tQS = ζ m ζm EBH EQS . (4) From Eq. 3 we find that if 10% of a disk is accreted onto a quark-star (which then collapses into a black-hole) and the rest is accreted into the black-hole, the energy output in a jet from the quark-star is ten times larger than the energy output in a jet from the black-hole. In this case we see from Eq. 4 that the duration of the accretion process onto the quark-star is a thousand times longer than the duration of the accretion into the black-hole. The above discussion is valid for the inner engine. We will now try to relate that to observations. We assume a direct relation between the energy output from the inner engine in the black-hole and quark-star phases and the observed energies from the two phases: EBH,obs ≃ hζsh,BHiEBH EQS,obs ≃ hζsh,QSiEQS, (5) where hζsh,BHi and hζsh,QSi is the energy conversion efficiency in the shocks averaged over the entire duration of the shock activity (e.g. Kobayashi et al. 1997). For multiple shocks, we assume this efficiency to be the same for the jet from the black-hole and the quark-star, hence hζBHi ≃ hζQSi. Using this, we obtain the following ratios between the observed energies and the accreted masses: EBH,obs EQS,obs ≃ EBH EQS ≃ ζm(cid:16) mdisk mdisk,QS − 1(cid:17) (6) To a first approximation we can assume that EQS,obs is the observed GRB energy, and EBH,obs is the observed energy in flares and bumps in the X-ray afterglow. The observed energies in X-ray flares and bumps compared to the observed energies released in gamma rays varies a lot in different bursts (from no flares or bumps to flares with fluence equal to the GRB fluence). If we take EBH,obs EQS,obs ≃ 10−2 then, in our model, it would imply that mdisk,BH ≃ mdisk,QS. (7) (8) In other words, an equal amount of mass is accreted onto the quark-star and the black-hole. This holds true if we neglect the energy carried by the slow shells from the quark-star, causing -- 9 -- the flattening of the early X-ray afterglow light curve. This is a reasonable assumption, since the energy in the flattening is less than in the GRB itself. As for a direct comparison of the time scale of the inner engine to the observed time scale, this is not possible since this is rather dictated by the complex interaction between the black hole ejecta and the quark star ejecta as we show next. 4. Interaction between black-hole ejecta and quark-star ejecta Figure 4 illustrates the process of flattening the X-ray light curve and generating X-ray flares and bumps. If the quark-star emits low-Lorentz factor shells at late stages of shell emission, the shell from the black-hole can catch up with the last shell and produce an X-ray flare via internal shocks. This merged shell may be capable of colliding with other slow shells from the quark-star, whereby more X-ray flares will be seen, an idea bearing resemblance to what is discussed in Zou et al. (2006). The energy output from the collisions depends on the difference in Lorentz factor of the colliding shells. This is likely highest in the first collision. However, the energy also depends on the masses of the colliding shells, so the energy output is not necessarily highest in the first X-ray flare (see eqs. A3 and A4). If the mass and Lorentz factor of the merged shell producing the X-ray flare(s) are high enough, a bump can be seen in the X-ray afterglow light curve as this shell collides with the external shock from the GRB. The pulse width of the X-ray flares are generally longer than the width of the flares in the GRB. There are also indications that later occurring flares are wider than flares occurring early (Zhang et al. 2006). Qualitatively this can be understood from Eq. A1 in the Appendix. The width of the burst is proportional to the separation of the colliding shells. The shells creating the GRB were all produced by the quark-star in a fairly short time, whereas the black-hole shell was created long thereafter. Our model seems capable of explaining why later flares have a longer duration. From Eq. A1 it is seen that the duration (for equal mass shells) depends on the initial distance between two shells. Since the flares in our model are all due to the jet from the black-hole colliding with slower parts of the quark-star jet, later flares are created by shells that were farther away from the black-hole jet initially. This leads to longer duration. The flattening of the X-ray light curve is due to lower Lorentz-factor shells emitted in the later stages of the quark-star jet. These shells will catch up with the external shock from the GRB at later stages, slowly re-energizing the external shock. This is reminiscent -- 10 -- of the refreshed shocks scenario in Sari & M´esz´aros (2000). Hence the re-energization is independent of the black-hole. The flare does depend on the black-hole formation, and also on the fact that the quark-star jet emitted slower shells. A flare does not necessarily lead to flattening, as only one slow shell from the quark-star is needed to produce a flare, whereas a more continuous sequence of low-Lorentz factor shells is needed to produce the flattening. The external shock will be decelerated by the surrounding medium. A low Lorentz factor shell will catch up with the external shock when this has decelerated to a comparable Lorentz factor. The range in Lorentz factors of the slow part of the quark-star jet is necessary to explain a given flattening therefore depends on the deceleration of the external shock, which is not well known. Following Falcone et al. (2006), a slow shell with Lorentz factor of about 20 will catch up with the external shock after t = 104 seconds, assuming a uniform external medium with density n = 10 cm−3. A shell with a Lorentz factor Γ ∼ 9 will catch up with the external shock after t = 105 seconds. 5. Light curve features: timescales and energies In this section we first summarize the light curve features that our model can produce (see Fig. 4), and then we discuss the different possible lightcurves (Fig. 5). Steep decay: The steep decay commonly seen in the early afterglow (t ≈ 102 − 103 seconds) is due to the curvature effect (Kumar & Panaitescu 2000). This is prompt gamma radiation from parts of the jet directed at an angle relative to the line of sight. Because of the large Lorentz factors involved, most of the radiation is beamed with a beaming angle Γ−1. Some of it will be directed to angles outside of Γ−1, but then softened by a Doppler factor. This is why what is seen as gamma rays along the line of sight is seen as X-rays at an angle from the line of sight. The steep decay continues until the luminosity from the external shock is dominant. This effect is dominant only because the central engine does not produce any visible activity, and the external shock is too weak to be seen. Flattening: The flattening of the light curve is due to the external shock being re- energized by low Lorentz factor shells emitted from the quark-star. These shells catch up with the decellerating external shock. These shells were among the last ejected from the quark star. It is important to note that there is not a one-to-one ratio for the duration of the inner engine and the observed radiation in this case. The inner-engine ejection of shells could last for some hundred seconds, whereas the re-energization can last for 104 − 105 seconds. Flares: The flares are caused by interaction between a jet launched by a black-hole and the slower parts of the ejecta from the quark-star. In order for flares to occur, the quark-star -- 11 -- must have accreted enough matter that it collapses to a black-hole. The shell ejected by the black hole can be very massive, mBH ∼ 1028 g and have a very large Lorentz factor. The shells ejected in the quark-star phase are likely less massive, however some of these might merge into massive ones before the black-hole shell interacts with them. In this case more than 1051 erg can be converted in one flare. The pulse width depends on the initial separation of the shells and on the width of the rapid shell (Kobayashi et al. 1997). The black hole shell can be fairly wide initially, and it is not unlikely that it spreads more before it collides with the quark-star shell. Finally, we mention recent work by (Liang et al. 2006) where a zero time point (right before the beginning of the flare) was taken as a signature of the reactivation of the engine. If true, this observation suggests that the engine reactivates more than once which cannot be reconciled with our model. We will tackle this issue elsewhere. Bump: The bumps seen at late times (t ∼ 104 − 105 seconds as in GRB 050502B; Falcone et al. 2006) are the result of the black-hole jet or the merged jet colliding with the external shock. The black-hole jet may or may not have interacted with shells from the quark-star to create flares before colliding with the external shock. If there are no, or only a few, slow shells from the quark-star that the black-hole jet collides with, the black-hole jet might not be slowed down significantly and it will catch up with the external shock at an earlier stage. If this happens, a flare with not too steep rise will result. However, after this flare, there will be no flattening, and no bumps. The black-hole jet's collision with the external shock marks the end of the inner engine's contribution to the observed afterglow light curve. Thereafter, a "classical" afterglow decay is the result, with a possible jet break due to non-isotropic ejecta creating the afterglow. There is a correlation between the energy emitted in the flares and in the bump (see § 6), so that bursts with more significant flares will have less pronounced bumps, and vice versa. This is assuming that a black-hole was formed and an equal sized disk formed around the black-hole. A small disk around the black-hole will lead to weak or no flaring and bumps. In the case when no black-hole formed, no flares and no bump will be seen. 5.1. Generic light curves in our model Here we discuss the eight different types of light curves that our model can generate (Fig. 5). Case 1. This case has all the properties discussed in the previous section i.e. flares, flattening and a late time bump. The flares are produced when a black-hole jet collides with slower shells from the quark-star. The flattening is due to slower shells from the quark-star -- 12 -- re-energizing the external shock, whereas the bump occurs when the black-hole jet collides with the external shock. This case is also shown in Fig. 4. Case 2. This case shows a light curve with one or more flares and a flattening of the X-ray afterglow light curve, but no bump. This will happen if the later stages of the ejection from the quark-star produces slower shells. These slower shells will re-energize the external shock, whereby the light curve is flattened. The accretion onto the black-hole launches an ultrarelativistic shell, and when this interacts with a slower shell from the quark-star, an X-ray flare is seen. However, the merged shell from the black-hole and the quark-star jet is not fast enough to catch up with the external shock, so no bump is seen. Case 3. This case shows flattening and a bump, but no flares. This is because the shell from the black-hole did not interact with any of the slower shells from the quark-star. This indicates that the black-hole shell was fairly slow. When the black-hole shell collides with the external shock, a bump is seen. As before, slower shells in the quark-star jet re-energizes the external shock, which flattens the light curve. Case 4. This case has a bump, but no flare and no flattening. This occurs when the quark-star jet does not produce any late time shells that can flatten the light curve, and the shell from the black-hole can not interact with any shells to create a flare. A bump is seen when the shell from the black-hole jet collides with the external shock. Case 5. This case has a flare and a bump, but no flattening. This is rather similar to case 4, however the quark-star jet emitted one or a few low Lorentz factor shells that the black-hole jet interacted with. This happened before the late shells from the quark-star collided with the external shock, so no flattening was created. GRB 050502B is an example of a burst in this category. Case 6. This case shows flattening, but no flare and no bump. This occurs when the quark-star emits low Lorentz factor shells that can flatten the light curve, but no black-hole was formed, or the jet from the black hole was to weak too make any observable signatures. Case 7. This case is similar to case 6, but the quark-star did not emit late time shells, so no flattening is seen. Case 8. This case shows flaring, but no flattening and no bump. This occurs when the quark-star generates some slow shells and a black-hole is formed. The jet from the black-hole interacts with the slow shells and produces flaring. However, in this process the jet is slowed down enough that it cannot make any significant impact on the external shock. Case 1a. In many observed afterglows the X-ray afterglow decays very steeply just after the end of the gamma ray emission. We have explained this as being due to the -- 13 -- curvature effect, that is an off-axis gamma ray emission, that in other directions is seen as the prompt emission. However, in some bursts this steep decay is not observed. Instead the X-ray afterglow declines gradually, with a power law similar to the decline at late times (t > 105 seconds). We suggest that in this case the external shock sets in earlier. There is still the possibility of having flares and flattening in the same way as before, when the jet from the black-hole collides with slower shells from the quark-star or the slower quark-star shells re-energize the external shock. 6. Light curve features: Correlations 6.1. Anti-correlation between X-ray flares and bumps We have proposed that the jet launched by the black-hole is responsible for the X-ray flares and bumps. The flares are produced when the black hole jet collides with slower parts of the jet from the quark-star. The bumps seen at later times in the X-ray afterglow are due to the black-hole jet colliding with the external shock. In Fig. 6 we explore the amount of energy converted and the resulting Lorentz factor of the merged shell when the black-hole jet with Lorentz factor Γ = 100 and mass m = 3 × 1028 g collides successively with N quark-star shells. We show four different cases, when the quark-star shells have Lorentz factors Γ = 10, 20, 30, or 40 and mass m = 3 × 1027 g (see Eqs A3 and A4 in the Appendix). In the first couple of collisions, the difference in Lorentz factor is fairly large, so a lot of energy is converted. However, after about 15 to 20 collisions, not much more energy is converted. This is because the Lorentz factors of the colliding shells are not very different. Several consecutive collisions will lead to several flares. We also see that if the Lorentz factor of the slow shells are high, the merged shell will get a higher Lorentz factor but less energy will be converted. The energy released in the bump depends on the kinetic energy of the black-hole jet (or the merged jet that produced the flares) when it collides with the external shock. If the jet is not very energetic, i.e. a lot of energy was lost due to flaring, a weak bump will be seen. If not much energy was lost due to flaring, because there were no slower shells from the quark-star, the bump can be more pronounced. To summarize: Since both the flares and the bump are generated by the black-hole jet, the maximum energy available to do so is the initial kinetic energy of the jet. If a large fraction of this energy is used to produce flares, less is available for the bump, and vice versa. This is in agreement with O'Brien et al. (2006a,b), who find that there is a correlation between the energy output in flares and in what they call the hump in observed GRB afterglows. The hump is the flattening starting from about t = 103 seconds and includes possible bumps. -- 14 -- 6.2. Correlation between prompt gamma energy and early X-ray afterglow energy The total energy output from the quark-star inner engine is ηQSmdisk,QSc2 ≈ 2 × 1052erg for mdisk,QS ∼ 0.1M⊙. This energy is shared between the prompt GRB3, the external shock creating the decaying afterglow, and the re-energization of the external shock creating the flattening of the early X-ray afterglow light curve. A GRB with low energy in a prompt gamma-ray phase has more energy available for the early X-ray afterglow. Note however, that in the early X-ray afterglow more energy can be emitted by the black hole jet interacting with the quark star ejecta, as explained in the previous section. 7. Case study In this section we will apply our model to a few observed X-ray afterglows that illustrate the different properties of our model: flares, flattening and bumps. 7.1. GRB 050219A (case 2) The observed light curve with the X-ray telescope (XRT) and the Burst Alert telescope (BAT) do not look like they agree in this burst. This apparent discontinuity was first reported in Tagliaferri et al. (2005) and may be due to an early X-ray flare (as proposed by O'Brien et al. 2006a) after about 90 seconds. Since this flare is so early, it may either be a late internal shock produced purely by the quark-star, or possibly the result of an early black-hole formation. The early steep decay after this proposed flare is either the end of the flare or more likely due to the curvature effect (M´esz´aros 2006). At later stages (from about 800 seconds) the light curve flattens and there are some small flares. These are the late and slow parts of the quark-star jet that re-energize the external shock. If we are right in our assumption that there is an early flare after about 90 seconds, this would correspond to case 2 in Fig. 5. 3 For a 10% radiative efficiency during shell collisions (as in the internal-external shocks model) in the quark-star phase this would imply that a maximum of about 2×1051 erg is released as synchrotron radiation. -- 15 -- 7.2. GRB 050502B (case 5) GRB 050502B shows a very strong X-ray flare, starting after about 350 seconds and peaking around 700 seconds (Falcone et al. 2006). In addition, there is one (or possibly two) bumps seen at about 40000 to 50000 seconds after the GRB trigger. The observed fluence of the flare is comparable or even bigger than the GRB fluence. There are indications of substructures within the flare itself. The total energy of the flare (or the GRB) is not known. We find by using Eqs. A3 and A4 that a black-hole jet of mr = 3 × 1028 g, moving with Lorentz factor of Γr = 200 and colliding with a shell from the quark-star with Lorentz factor Γs = 10 and mass ms = 5 × 1027 gives an internal energy Eint = 2.4 × 1051 ergs and a Lorentz factor of the merged shell of Γm = 96. The mass of the slow shell from the quark-star may seem large, but this can be the result of collisions at earlier times. These collisions may not have left observable flares if the difference in Lorentz factors in the colliding shells were not large or these collisions could have been part of the GRB itself. The black-hole must have ejected a fairly large amount of mass (3 × 1028 g), indicating that the accretion disk must have been fairly large. The bump(s) at t > 104 seconds occur when the merged shell that created the flare catches up with the external shock (as suggested in Falcone et al. 2006). The underlying decay of the afterglow is due to the external shock. This burst corresponds to case 5 in Fig. 5. 7.3. GRB 050421 (case 1, 2, 5 or 8) GRB 050421 shows two flares in the early X-ray afterglow, the first flare peaking after 111 seconds and the second after 154 seconds (Godet et al. 2006). The duration of the gamma ray emission was about 10 seconds. We remind the reader that these flares are due to a jet from a black-hole colliding with slower parts of the jet from the quark-star. The width of the first peak is about 10 seconds. This can be achieved if the quark-star emits a slow shell about 90 seconds following the first shell (c.f. eq. A1): δt = L ac = c∆t ac = 20 2 = 10 s (9) (∆t being the time interval between the emission of the quark-star shell and the black-hole shell). Alternatively, if the slower shell is emitted at the end of this burst (after about 10 seconds) but the jet from the black-hole has a significantly higher Lorentz factor, about 10 times higher than the quark-star shell : δt = L ac = 100c 10c = 10 s. (10) -- 16 -- The second flare is created when the merged shell creating the first flare collides with another slow shell emitted from the quark-star. The later stages of this afterglow (> 103 seconds) are not observed. Therefore we cannot say if there is any flattening or bump. This corresponds to any one of cases 1, 2, 5 or 8 in Fig. 5. Alternatively, these two flares could both be produced by internal shocks from shells created in the late stages of the accretion onto the quark star. 7.4. GRB050401 (case 6 without steep decay) An example of a GRB without the early steep decline is GRB050401 (De Pasquale et al. 2006). The afterglows declines gradually with a decay index α ≈ 0.65 (Panaitescu et al. 2006) until about 4300 seconds. After this, the decay is steeper with a decay slope α ≈ 1.39. We ascribe the behavior until about 4300 seconds as flattening due to slower quark star shells re-energizing the external shock. There is no clear evidence of any flaring in this afterglow, in which case no black-hole jet interacted with the quark star jet or the external shock. This corresponds to case 6 without the initial steep decay. The isotropic equivalent energy of this burst is about Eγ,iso ∼ 3 × 1053 erg. Assuming a beaming angle of 5◦, this gives a total energy in gamma rays of about 4 × 1050 erg (for a bipolar jet). Zhang et al. (2006) found that almost five times as much energy was injected in this burst during the afterglow than during the prompt gamma radiation, meaning about 2 × 1051 ergs was released in the early X-ray afterglow. In our model, this corresponds to 0.1M⊙ disk being accreted onto a quark star with 10% of the accreted energy ejected into a bipolar jet, and only about 2% of the jet energy goes into radiation. Most of the jet energy (the remaining 98%) is used to re-energize the external shock. 8. Discussion 8.1. Hypernova-GRB association in our model The formation of the quark-star through a quark nova (between stage1 and 2) shortly after the core collapse or directly in the core collapse releases an extra amount of energy that can re-energize the SN ejecta, reminiscent of a hypernova (Ouyed et al. 2002; Keranen et al. 2005). GRBs seem to be associated with hypernovae (i.e. GRB 030329; Hjorth et al. 2003), (GRB 980425; Galama et al. 1998), however, the opposite is not always true (i.e. SN 2002ap; Mazzali et al. 2002, although a lack of GRB observation does not necessarily mean that no -- 17 -- GRB occurred). In order to get a hypernova, the energy from the conversion to quark matter is necessary. In our model, the GRB is produced by the jet launched by accretion onto the quark star. If a black hole is formed directly in the collapse (no quark star stage), there will be no hypernova. Accretion onto the black-hole can launch a jet, and internal shocks in the jet can produce a GRB (see § 8.2.1). However, since there is only one stage, there can be no late activity. 8.2. SN-less GRBs in our model GRB060614 and GRB060505 were apparent nearby, long duration GRBs that showed no sign of a supernova, leading to the term SN-less GRBs (Fynbo et al. 2006). In the litterature possible explanation for this have been that not enough 56Ni is formed in the explosion (Tominaga et al. 2007), or that this is a very long and energetic "short" burst (Zhang et al. 2007). Here we give a possible interpretation of this phenomena within our model. We suggested above that the energy released by the quark nova (formation of a quark-star) could potentially lead to a hypernova. However, if the star is too massive to explode even with the extra energy released from the quark nova it will instead collapse and no supernova/hypernova will be seen. We assume that a hyperaccretion disk can form around the quark star while most of the star is trying to explode4, and as before accretion onto the quark-star powers an ultrarelativistic jet giving rise to a GRB. In this scenario a black-hole is the most likely final outcome: either the envelope falls back directly onto the compact core, or it feeds to the accretion disk. In the first case a black hole is formed but no ultrarelativistic jet is launched from the accretion onto it and hence no flaring or bump activity is seen in the early X-ray afterglow. In the second case the quark-star collapses to a black hole when it has accreted enough mass and continued accretion onto the black hole will launch an ultrarelativistic jet that can give rise to flaring and bump activity seen in the early X-ray afterglow. 8.2.1. Black-hole formation without a quark-star stage In the event that a black-hole is formed immediately following the core collapse, a collapsar scenario can occur if there is sufficient angular momentum present. Simulations of this scenario by De Villiers et al. (2005) show that an ultrarelativistic jet is launched 4This scenario is initially similar to that leading to a GRB with an associated hypernova, only that in this case the ejected material does not have enough energy to escape the star. -- 18 -- that can give rise to internal shocks and a GRB. In addition to the jet, a coronal wind was launched that could explode the remainder of the star. However, the energy of this wind was of the order 1048 − 1049 erg, which is not enough to explain a supernova. Hence this could be another possibility for explaining supernova-less GRBs, as discussed above. The duration of the accretion process is short (accretion rate of the order 1M⊙ s−1 leading to a rather short duration GRB). An alternative occurs when the coronal wind is not strong enough to explode the star. The star will then collapse, and assuming that it has high angular momentum, it will form a massive accretion disk around the black-hole. The accretion rate onto the black-hole is probably of the same order as with a smaller disk, however with much more mass to accrete this could lead to a somewhat longer duration jet and GRB. This resembles the original failed supernova discussed by Woosley (1993). Both cases discussed above lead to a one stage model resulting in the absence of late time activity. Therefore, there will be no flares or bumps in the afterglow light curve, but flattening is still possible. 8.3. Collapsar vs. accreting quark-star as inner engine for GRBs In this section we will briefly discuss the differences and similarities between our model and the collapsar model (Woosley 1993) for the inner engine for GRBs. Similarities: Both the collapsar and our model assume the death of a massive, rapidly rotating star as the triggering mechanism. Both models launch an ultrarelativistic jet, in which internal shocks accelerate electrons that produce synchrotron radiation. As this jet is decelerated by the ISM, an external shock forms in which electrons again are accelerated and emit synchrotron emission that is responsible for the afterglow. Differences: The collapsar model assumes that the compact core left behind in the collapse of the core of the progenitor star quickly collapses to a black-hole, whereas in our model this compact core collapses to a quark star instead. This opens up some interesting possibilities. The accretion rate into a black-hole and a quark-star are totally different. Accretion into a black-hole is a very rapid process, with accretion rates of the order of one solar mass per second. Accretion onto a quark star on the other hand is much slower (about 104 times slower). This can therefore easier explain the long duration (∼ 1000 seconds) seen in some bursts. A main focus of this paper is the flaring often seen in the X-ray afterglow, which is indicative of the engine being restarted. With a quark-star the possibility of it collapsing to a black-hole exists, which explains how the inner engine can be restarted leading -- 19 -- to the observed features in the early X-ray afterglow. 9. Summary and conclusion To summarize, the properties of the three stages of the inner engine in our model are: • The process is initiated by the collapse of the iron core of a massive, rapidly rotating star. • The collapsed core will either leave a neutron star or a quark-star behind. • If a neutron star is left behind, it can collapse to a quark-star at a later stage, creating a delay between the supernova and the GRB. • Fallback material from the supernova and the quark nova forms a disk around the quark-star. Accretion onto the quark-star generates the GRB by powering an ultrarel- ativistic jet. Internal shocks in this jet create the GRB. • Accretion continues, but at some point it cannot heat up the star sufficiently. This halts the emission of shells, ending the GRB. • When the quark-star has accreted too much material, it collapses into a black-hole. Further accretion onto the black-hole launches an ultrarelativistic jet. The emission features in our model can be summarized as follows: • Early, steep decay in the X-ray afterglow is due to the curvature effect (this is not specific to our model, but rather a generic feature in many models). • Flares are created by interaction between the jet from the accretion onto a black-hole and slower parts from the jet from the quark-star. • Re-energization of the external shock (seen as flattening of the X-ray afterglow light curve) is due only to the jet from the quark-star. Slower parts of this jet re-energizes the external shock. • When the jet from the black-hole collides with the external shock from the GRB, a bump is seen in the afterglow light curve. -- 20 -- In conclusion, we have presented a three stage (effectively two stage) model for the inner engine for GRBs involving a neutron star phase, followed by a quark star phase then by a black hole phase. This model seems to account for the observed prompt gamma ray emission, as well as the features of the early X-ray afterglow and as such warrants further study. We would like to thank J. Hjorth, M. Lyutikov M. Cummings, W. Dobler, D. Leahy and B. Niebergal as well as the anonymous referee for helpful remarks. A. Internal-external shocks model Piran (2005) shows that the pulse width is proportional to the separation between two shells: δt ≈ Rs/2aΓ2c = L/ac equal mass shells, (A1) where Rs is the distance at which the collision takes place, a is the ratio between the fast and slow shells Lorentz factor, Γ the Lorentz factor of the slow shell, L the separation of the shells and c the speed of light. Assuming that the colliding shells have equal mass, and the faster has a Lorentz factor two times the slower shell, the separation between the two shells must be 3 × 1013 cm in order to have a pulse width of 500 seconds. This corresponds to a thousand seconds separation between the emission of the two shells. The shells will collide at a distance Rs ≈ 2Γ2L (A2) where Γ refers to the Lorentz factor of the slower shell. For Γ = 20 and a separation of 3 × 1013 cm as before, the collisions occur at Rs = 2 × 1016 cm. If the Lorentz factors are too big, the external shocks will occur before internal shocks could occur. If the photon energy is large enough to produce e+e− pairs, the Lorentz factors of the shells have to be above a hundred to overcome the compactness problem. However, if the shocks creates X-rays, the photons cannot create e+e− pairs, and there is no lower bound on the Lorentz factor. The internal energy of the merged shell is given by: Eint = mrc2(Γr − Γm) + msc2(Γs − Γm), (A3) where mr is the mass of the rapid shell, ms of the slow shell, Γr is the Lorentz factor of the rapid shell, Γs is the Lorentz factor of the slower shell and Γm of the merged shell: Γm =s mrΓr + msΓs mr/Γr + ms/Γs . (A4) -- 21 -- A certain fraction of this energy will be emitted as radiation. In order to have an efficient collision (a collision in which a lot of energy is converted to internal energy), the masses must be similar, and the Lorentz factor of the fast shell must be at least twice that of the slower shell. For a mass of the rapid shell of about 3 × 1028 g, which is not unlikely in a black-hole jet ( M ∼ 10−5Mdisk/s in De Villiers et al. 2005), and a mass of the slow shell of 3 × 1027 g, Lorentz factor of the rapid shell of Γr = 65, and Lorentz factor of the slow shell of Γs = 20, which gives Γm = 57 and an internal energy of about 1050 erg in a collision. Alcock, C., Farhi, E., & Olinto, A. 1986, ApJ, 310, 261 REFERENCES Bagchi, M., Ouyed, R., & Staff, J. E., Ray, S., Dey M., Dey J., 2006, astro-ph/0607509. Beloborodov, A. M., Stern, B. E., & Svensson, R., 2000, ApJ, 535, 158. Burrows, D. N., et al., 2005, Science, 309, 1833 Dai, Z. G., Wang, X. Y., Wu, X. F., Zhang, B., 2006, Science, 311, 1127 De Villiers, J. P., Staff, J. E., & Ouyed, R., 2005, astro-ph/0502225. Della Valle, M. "16th Annual October Astrophysics Conference in Maryland", eds. S. Holt, N. Gehrels and J. Nousek, AIP Conf. Procs. astro-ph/0604110 De Pasquale, M., et al., 2006, MNRAS, 365, 1031 Falcone, A. D., et al., 2006, ApJ, 641, 1010. Frank, J., King, R., Raine, & D. J. 1992, Accretion Power in Astrophysics (Cambridge: Cambridge Univ. Press) Fryer, Ch. L., Benz, W., Herant, M. 1996, ApJ, 460, 801 Fynbo, J., et al. 2006, Nature, 444, 1047 Galama, T. J., et al. 1998, Nature, 395, 670 Gehrels, N., et al. 2004, ApJ, 611, 1005 Godet, O., et al., 2006, A&A, 452, 819 Harko, T., Cheng, K. S., 2002, A&A, 385, 947 -- 22 -- Hjorth, J. et al., 2003, Nature, 423, 847 Itoh, N, 1970, PThPh, 44, 291 Iwazaki, A. 2005, PhRvD, 72, 114003 Keranen, P., Ouyed, R., & Jaikumar, P., 2005, ApJ, 618, 485 Klu´zniak, W., Ruderman, M., 1998, ApJL, 505, 113 Kobayashi, S., Piran, T., & Sari, R., 1997, ApJ, 490, 92 Kumar, P. & Panaitescu, A. 2000, ApJ, 541,L51 Liang, E. W., et al., 2006, ApJ, 646, 351 Mazzali, P., et al., 2002, ApJ, 572, L61 M´esz´aros, P. , 2006, "16th Annual October Astrophysics Conference in Maryland", eds. S. Holt, N. Gehrels and J. Nousek, AIP Conf. Procs. astro-ph/0601661 Narayan, R., Paczynski, B., & Piran, T., 1992, ApJ, 392, L83 O'Brien, et al., 2006a, ApJ, 647, 1213 O'Brien, P. T., Willingale, R., Osborne, J. P., & Goad, M. R., 2006b, NJPh, 8, 121. Ouyed, R., Dey, J., & Dey, M. 2002, A&A, 390, 39 Ouyed, R., Rapp, R., & Vogt, C., 2005, ApJ, 632, 1001 (ORV). Ouyed, R., Niebergal, B., Dobler, W., & Leahy, D., 2006, ApJ, 653, 558. Panaitescu, A., M´esz´aros, P., Burrows, D., Nousek, J., Gehrels, N., O'Brien, P., & Willingale, R. 2006, MNRAS, 369, 2059 Piran, T., 1999, PhR, 314, 575 Piran, T., 2005, Rev. Mod. Phys., 76, 1143. Popham, R., Woosley, S.E., & Fryer, Ch. 1999, ApJ, 518, 356. Rhoads, J. E. 1999, ApJ, 525, 737. Ruderman, M. A., Tao, L., Klu´zniak, W., 2000, ApJ, 542, 243 Sari, R. & M´esz´aros, P. 2000, ApJ, 535, L33. -- 23 -- Staff, J., Ouyed, R., & Jaikumar, P. 2006, ApJ, 645, L145 Tagliaferri, G. et al., 2005, Nature, 436, 985 Tominaga, N., Maeda, K., Umeda, H., Nomoto, K., Tanaka, M., Iwamoto, N., Suzuki, T. & Mazzali, P. A., 2007, ApJ, 657, L77. Usov, V. V. 1992, Nature, 357, 472 Vogt, C., Rapp, R., & Ouyed, R. 2004, Nuc. Phys. A, 735, 543. Wheeler, J. C., Yi, I., Hoflich, P., Wang, L., 2000, ApJ, 537, 810. Woosley, S.E. 1993, ApJ, 405, 273. Zhang, B., Fan, Y.Z., Dyks, J. Kobayashi, S., M´esz´aros, P., Burrows, D.N., Nousek, J.A., & Gehrels, N. 2006, ApJ, 642, 354. Zhang, B, Zhang, B. B., Liang, E. W., Gehrels, N., Burrows, D. N. & M´esz´aros, P. 2007, ApJ, 655, L25. Zou, Y. C., Dai, Z. G., & Xu, D. 2006, ApJ, 646, 1098. This preprint was prepared with the AAS LATEX macros v5.2. -- 24 -- "Early/New afterglow" "Late/Classical afterglow" Steep decay flare 2 10 − 10 4 s flattening e t a r t n u o C bump jet break 6 10 − 10 5 s 3 10 s 4 10 − 10 5 s Time Fig. 1. -- Generic X-ray afterglow (e.g. Zhang et al. 2006). At early times (t < 1000 seconds), a steep decay is often seen, followed by a flattening starting at t = 103 and lasting for about 104 − 105 seconds. Thereafter a steeper decay, and after about a day or so a new break is seen. The last break is termed "jet break". On top of this one or more flares are often observed between t = 102 −104 seconds, and one or more bumps can be seen between t = 104 seconds and t = 105 seconds. NOTE: Not all features are seen in every bursts. Before (the launch of the) Swift satellite, the part of the figure including the steep decay, the flares and the flattening was essentially unobserved. -- 25 -- SN ejecta External shock Ultrarelativistic jet SN ejecta External shock Ultrarelativistic jet SN ejecta Accretion disk QN NS Accretion disk BH QS Hours to days 2−2000 s Accretion disk Black hole Stage 1 Stage 2 GRB light curve e t a r t n u o C Stage 3 X−ray afterglow e t a r t n u o C Time Time Fig. 2. -- The three stages in a GRB inner engine. Stage 1: A neutron star is formed in the core collapse of a massive star. The neutron star can later collapse into a quark-star due to spin-down or accretion. If this takes a long time, it will cause a delay between the supernova and the GRB. Stage 2: A quark-star in the CFL phase is formed, either directly in the supernova, or from the neutron star in Stage 1. Accretion onto this star heats the star, leading to emission of photons as the main cooling mechanism. If the star is heated above the plasma frequency of about 7.7 MeV (ORV), the photons can escape the star and will interact with the accreting matter. The infalling matter is ejected, leading to a halt in the accretion process. When the star cools down below 7.7 MeV, accretion can be resumed. This creates episodes of accretion and ejection. Internal shocks created from colliding shells in the ejected material create the GRB. Stage 3: At later stages in the accretion process the accretion may not be strong enough to heat the star above 7.7 MeV. This prevents ejection of material, and hence terminates the GRB. However, the accretion will continue. If the star accretes too much matter, it will collapse to a black-hole. This dramatically changes the accretion, and launches an ultrarelativistic jet. This jet will interact with slower parts of the jet from the quark-star, leading to internal shocks. This creates the flares often seen in the early X-ray afterglow. When the jet from the black-hole collides with the external shock, a bump is observed on the light curve. -- 26 -- Accretion/Heating Ejection/Cooling accretion ω magnetic field line heating ejected material magnetic field line photon cooling QS ~1ms QS ∼1µ s heating magnetic field line accretion photon cooling magnetic field line ejected material Fig. 3. -- Cartoon illustrating the jet launching mechanism in our model. Infalling material follows magnetic field lines to the polar cap (magnetic pole is at the same location as the geographic pole in a CFL star, see text), where it heats the star. The timescale for heating the star is of the order milliseonds. The star then cools by emitting photons on a timescale of the order of microseconds. These photons interact with the infalling material, ejecting some of that with large Lorentz factors. -- 27 -- Fast shell from BH phase Slower (late) shells from QS phase External shock (ES) from QS phase Γ >Γ s,BH s,QS s,QSΓ >ΓES,QS ΓES,QS Re−energization of QS ext. shock X−ray flare Bump (BH jet +QS ES collision) x u l f time Fig. 4. -- Illustration of the mechanism leading to X-ray flares and flattening of the X-ray afterglow light curve. In both cases, the quark-star needs to emit low-Lorentz factor shells in the late stages of shell emission. When these shells collide with the external shock from the GRB, that shock is re-energized and a flattening of the light curve results. If the quark star accretes too much mass, it will collapse to a black-hole. The jet emitted in the accretion process onto the black-hole can be massive and have a high Lorentz factor. When this outflow collides with a slow shell from the quark-star, an X-ray flare results. If this merged shell collides with more low-Lorentz factor shells from the quark-star, more flares will result. A bump can be seen when the black-hole ejecta collide with the external shock. Case 1 Case 2 -- 28 -- Case 3 Case 4 Case 5 x u l f x u l f time Case 6 x u l f x u l f x u l f x u l f x u l f time Case 7 time Case 8 time Case 1a time x u l f x u l f time time time time Fig. 5. -- The differeant kinds of light curves derived from our model. Case 1 is also illustrated in Fig. 4. There are flares, flattening and a bump. This occurs when the quark-star jet produces slow shells that can re-energize the external shock. The black hole jet interacts with some of these shells and creates flares. When the merged black-hole jet collides with the external shock, the bump is formed. Case 2: We see one or more flares and a flattening of the light curve, but no bump. This is because the shell from the black-hole is slowed down so much by the other shells from the quark-star that it reaches the external shock at very late times. Case 3: Flattening and a bump, but no flare. The shell from the black- hole did not collide with any shell and therefore did not create a flare. It creates a bump when it collides with the external shock. Case 4: No flare, no flattening, but a bump. The quark-star jet did not contain any late time shells, so there were no flattening and no flare created. As in Case 3, the black-hole jet creates a bump when it collides with the external shock. Case 5: Flare and bump, but no flattening. There are only a few late time shells emitted from the quark-star. The black hole jet collides with these and creates flares. When this merged shell collides with the external shock a bump is seen. No late time shells can flatten the light curve. Case 6: No flares, no bump but flattening. The quark-star emits late time shells that flattens the light curve. No black-hole was formed. Case 7: No flares, no bump, no flattening. No black-hole was formed, and the quark-star did not emit any late time shells. Case 8: Flares, but no bump and no flattening. This indicates that a black-hole which launched an ultrarelativistic jet was formed, but in the process of forming the flares this jet was slowed down enough so that it cannot energize the external shock to generate a bump. Case 1a: Same as case 1, but no early steep decay. All cases from 1 to 8 can exist without the early steep decay as illustrated in case 1a where the decay is absent for case 1. Note that a possible break in the light curve due to collimation of the outflow producing the GRB and the afterglow will occur at later times than what is shown in this figure. 2.0⋅1051 2.0⋅1051 2.0⋅1051 2.0⋅1051 1.8⋅1051 1.8⋅1051 1.8⋅1051 1.8⋅1051 1.6⋅1051 1.6⋅1051 1.6⋅1051 1.6⋅1051 1.4⋅1051 1.4⋅1051 1.4⋅1051 1.4⋅1051 1.2⋅1051 1.2⋅1051 1.2⋅1051 1.2⋅1051 t t t t n n n n E E E E i i i i 1.0⋅1051 1.0⋅1051 1.0⋅1051 1.0⋅1051 8.0⋅1050 8.0⋅1050 8.0⋅1050 8.0⋅1050 6.0⋅1050 6.0⋅1050 6.0⋅1050 6.0⋅1050 4.0⋅1050 4.0⋅1050 4.0⋅1050 4.0⋅1050 2.0⋅1050 2.0⋅1050 2.0⋅1050 2.0⋅1050 0.0⋅100 0.0⋅100 0.0⋅100 0.0⋅100 0 0 0 0 100 100 100 100 m m m m Γ Γ Γ Γ 80 80 80 80 60 60 60 60 40 40 40 40 20 20 20 20 0 0 0 0 0 0 0 0 -- 29 -- Γ Γ Γ Γ s s s s =10 =10 =10 =10 Γ Γ Γ Γ s s s s =20 =20 =20 =20 Γ Γ Γ Γ s s s s =30 =30 =30 =30 Γ Γ Γ Γ s s s s =40 =40 =40 =40 10 10 10 10 20 20 20 20 30 30 30 30 40 40 40 40 50 50 50 50 60 60 60 60 70 70 70 70 N N N N Γ Γ Γ Γ s s s s =40 =40 =40 =40 Γ Γ Γ Γ s s s s =30 =30 =30 =30 Γ Γ Γ Γ s s s s =20 =20 =20 =20 Γ Γ Γ Γ s s s s =10 =10 =10 =10 10 10 10 10 20 20 20 20 30 30 30 30 40 40 40 40 50 50 50 50 60 60 60 60 70 70 70 70 N N N N Fig. 6. -- Figure showing the total energy converted to internal energy versus number of collisions (upper panel) and the Lorentz factor of the merged shell versus number of collisions (lower panel). In all cases the rapid shell has a Lorentz factor of 100 initially, and all the slow shells have Lorentz factors of 10, 20, 30 or 40 respectively. The mass of the rapid shell is 3 × 1028 g, and the mass of the slow shells are 3 × 1027 g. The figures show that after 10 to 20 collisions a plateau is reached and no more energy is converted. The Lorentz factor reaches a plateau. In this plateau, collisions can still occur, so the merged shell can gain mass. When the slower shells have higher Lorentz factors, the merged shell ends up with a higher Lorentz factor after the collisions, and the energy conversion is reduced. This leaves more energy to be converted when this merged shell collides with the preceding quark-star external shock.
astro-ph/9811445
1
9811
1998-11-27T18:22:38
Proceedings of the Bonn/Bochum-Graduiertenkolleg Workshop 'The Magellanic Clouds and Other Dwarf Galaxies'
[ "astro-ph" ]
The Workshop 'The Magellanic Clouds and Other Dwarf Galaxies' was held at the Physikzentrum Bad Honnef in January 1998. The proceedings comprise 79 contributions. About 1/3 of the 352 pages contain the following Reviews: The Violent Interstellar Medium in Dwarf Galaxies: Atomic Gas (Elias Brinks and Fabian Walter), Hot Gas in the Large Magellanic Cloud (You-Hua Chu), Astrophysics of Dwarf Galaxies: Structures and Stellar Populations (John S. Gallagher), Star-forming regions and ionized gas in irregular galaxies (Deidre A. Hunter), The Law of Star Formation in Disk Galaxies (Joachim Koeppen), Strange Dark Matters in Nearby Dwarf Galaxies (Mario Mateo), Holes and Shells in Galaxies: Observations versus Theoretical Concepts (Jan Palous), Detailed Recent Star Formation Histories of Dwarf Irregular Galaxies and Their Many Uses (Evan D. Skillman et al.), and Nearby Young Dwarf Galaxies (Trinh X. Thuan and Yuri I. Izotov). See the complete electronic version for further details.
astro-ph
astro-ph
Bonn/Bochum-Graduiertenkolleg Workshop The Magellanic Clouds and Other Dwarf Galaxies Physikzentrum Bad Honnef, Germany 19th – 22nd January 1998 Edited by Tom Richtler & Jochen M. Braun Sternwarte der Universitat Bonn, Auf dem Hugel 71, D–53121 Bonn, Federal Republic of Germany 8 9 9 1 v o N 7 2 1 v 5 4 4 1 1 8 9 / h p - o r t s a : v i X r a Shaker Verlag October 1998 Foreword by the Spokesman of the Graduiertenkolleg It is a great pleasure that the 25th meeting of our Graduate Working Group, which is being supported now in its 6th year by the Deutsche Forschungsgemeinschaft, could become a significant international event. It is also gratifying to see this workshop as a harvest after 5 years of intensive work. The Graduate Working Group has seen a host of guest scientists who have visited in the framework of close collaboration, or gave lectures and seminars during the many previous meetings. This gathering of experts on galaxies, and particularly dwarf galaxies, represents a small milestone in the recent past of galaxy research at the Universities of Bonn and Bochum. Until the late 70's, the investigation of dwarf galaxies was restricted to a few small groups. Although the Magellanic Clouds had been studied extensively due to their proximity, it took a while until the term dwarf galaxy became fashionable among astrophysicists. There are several reasons why dwarf galaxy research is one of today's fastest growing branches of modern astrophysics. First, the Magellanic Clouds have always been appreciated as the ideal astrophysical laboratory in which one can study all kinds of processes without being hampered by distance and line-of-sight problems. Second, since it was known early on that dwarf galaxies are so little evolved, their relevance to cosmology was recognized. Third, by virtue of their low mass, the evolution of their ISM and stellar populations can be studied in the absence of internal large-scale triggers like stellar bars or density waves. It is therefore not surprising that the past two decades have seen a rapidly increasing rate of related research projects and publications, accompanied by an ever increasing number of conferences dedicated to this subject, one of which is being reported in these proceedings. While to the non-expert it might at first glance look as if dwarf galaxy research is a rather specialized field, a quick look at contemporary reviews on galaxy research demonstrates that understanding low-mass galaxies is of utter importance to also comprehend what is going on in more massive and, in particular, more distant ones. A prime example is the understanding of the faint blue galaxy population at redshift 0.4 to 1.0, which has experienced a further boom with the scrutiny of distant objects in the Hubble Deep Field. It is obvious that in order to understand stellar systems at large distances, it is all the more important (also for the cosmologists) to understand those just outside our door step! Research on the Magellanic Clouds and other dwarf galaxies had been going on at the Universities at Bonn and Bochum for some time prior to the Bonn/Bochum Graduate Working Group. It was therefore quite natural to combine the experience gained in two small research groups in order to intensify this research, while at the same time establishing an educational programme in which graduate students could benefit for their own career, and contribute to the progress in a very proliferous field of research. We are grateful to the Deutsche Forschungsgemeinschaft for financing this Graduate Working Group, thus rendering possible this intensive educational and research programme in a very prolific and fascinating field of research. I would like to take this opportunity to thank Christian Bruns for his invaluable help in organizing this workshop. It has been (and still is) a great pleasure to work with the students in the Graduate Working Group. It is their enthusiasm and diligence that, besides my natural endeavour in this field of research, ensures permanent motivation! Uli Klein Bonn, April 1998 Foreword of the Editors After having completed the present edition of the proceedings of our workshop on "The Magellanic Clouds and Other Dwarf Galaxies" we wish to express our gratitude to all contributors. The concept of our "Graduiertenkolleg" is meant to provide an efficient framework for graduate education and research and to foster not only the relations between students and supervisors but also between students and distinguished representatives of the community of our field of research. In this sense the graduate school can be seen as an attempt to realize the old (and often forgotten) idea of unifying teaching and research at universities. This edition is a handy manifestation. We remember quite well that, when we tried to expand our scientific ambitions centered on the Magellanic Clouds, the extension "... and Other Dwarf Galaxies" sounded meaningless to many ears. Looking at the entire collection of reviews, talks and posters addressing such a large variety of astrophysical issues, it now seems to us that no better headline for merging the various interests within our graduate school could have been found. We hope that it is of interest for many researchers to have a look at the extensive key word and object lists (see the apendices "Subject Index" and "List of Objects" on page 337 and 331 respectively) to quickly find specific information among the wealth of topics and objects, which appear in contributed talks and posters. We also hope that the attractive cover (designed by Michael Hilker and Sven Kohle) strengthens the wish of a potential reader to know what is in the book. We wish to thank the Shaker Verlag, namely Dr. Shaker, for giving us the permission to make the proceedings available in free electronic form on our webserver (see also the next preface about "Electronic Publishing") and all the participants (see the "List of Participants" on page 313) for their help in creating a fascinating workshop, for their contributions, and of course for further supporting the electronic version with updates and new links. We very much appreciate the contribution by Joachim Koppen, who fell sick shortly before the workshop and could not participate. Considering the statistics of incoming contributions (see electronic version, at the URL: http://www.astro.uni-bonn.de/webgk/ws98/techinf a.html#subm), it is noteworthy that the first contribution was a review, and so was the last. The distribution of time of the arrival of the reviews was nearly homogeneous, whereas the one for the talks and especially for the posters was peaking at our deadline. Many thanks go to the Deutsche Forschungsgemeinschaft as the funding institution of the Graduiertenkolleg and also for supporting this workshop. Tom Richtler & Jochen M. Braun Bonn, October 1998 Electronic Publishing In the last few years sciences have improved by new electronic means, not only to create huge data bases and providing tools for data reduction, but also in electronic publishing (EP) and information exchange. One of the most powerful aids apart from E-Mail and FTP (File Transfer Protocol) is the World Wide Web (WWW or W3), which has developed from the "toy" status to a very fast publication means, useful for announcing meetings, presenting institutes and working groups, providing the latest data about equipment, and nowadays also for publishing. Right from the start WWW was intended as a scientific tool developed at CERN, but soon was adopted by the whole society, at least in some countries with approriate and cheap telecommunication infrastructure. Thus in addition to the tools (like browsers) developed by scientists, commerial products arose and the language of the Web, HyperText Markup Language - HTML, came into Babylonian trouble. To create future standards and secure the general accessibility, the WWW Consortium (W3C) formed, which has created the current (but still unsupported) standards HTML 4.0 and CSS 2.0, and is developing further scientificaly relevant techniques, namely XML, XSL, and DOM. To focus on astrophysics, literature search in the good old libraries was supported by electronic search engines, litearature data bases, and preprint servers, e.g. ADS, CDS, and XXX, and publishing in printed journals including the well-tried referee process was extended by electronic versions, in different quality and accessibility. That EP is of enhanced importance today can be seen e.g. by the controversial panel discussion at the "AG Tagung" in Insbruck 1997 after the talk "Electronic publishing in its context and in a professional perspective" by Andr´e Heck (1998, Reviews in Modern Astronomy 11, 337). Let me stress a few aspects which, from my personal point of view are import in future: • We need an international referee process decoupled from journals or publishers and revised international copyright laws doing justice to sciences. • Scientists have to stay involved in modern techniques, i.e. knowing the basics of TEX, HTML etc., and every institution needs scientists especially trained in these subjects. • Continent based mirror servers should be erected providing certified scientific information easily and freely accessible for the whole international community. As a small step in this respect we present these proceedings in two different versions. One is the classical book (constructed with the typesetting system TEX & LATEX), which is published and distributed by the Shaker Verlag, the second is the electronic version (written in HTML) on the WWW server of the Astronomical Institutes of the University of Bonn. The latter can be visited via the following Uniform Resource Locator (URL): http://www.astro.uni-bonn.de/webgk/ws98/cover.html and should provide the full information including updates after the day of printing (noted in the "Erratum", an appendix of the electronic version) and related URLs. As technical modifications are possible - and even probable, people using the HTML version may also visit the appendix "Technical Information" (e.g. giving recommended browsers). I hope that the electronic version will be a fruitful and frequently visited supplement to the printed version. Jochen M. Braun Bonn, October 1998 Table of Contents Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Foreword by the Spokesman of the Graduiertenkolleg . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Foreword of the Editors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Electronic Publishing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Table of Contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Review Talks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . The Violent Interstellar Medium in Dwarf Galaxies: Atomic Gas Elias Brinks and Fabian Walter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Hot Gas in the Large Magellanic Cloud You-Hua Chu . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Astrophysics of Dwarf Galaxies: Structures and Stellar Populations John S. Gallagher . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Star-Forming Regions and Ionized Gas in Irregular Galaxies Deidre A. Hunter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . The Law of Star Formation in Disk Galaxies Joachim Koppen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Strange Dark Matters in Nearby Dwarf Galaxies Mario Mateo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Holes and Shells in Galaxies: Observations versus Theoretical Concepts Jan Palous . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Detailed Recent Star Formation Histories of Dwarf Irregular Galaxies and Their Many Uses Evan D. Skillman, Robbie C. Dohm-Palmer, and Henry A. Kobulnicky . . . . . . . . . . Nearby Young Dwarf Galaxies Trinh X. Thuan and Yuri I. Izotov . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iii iii v vii ix 1 1 11 25 37 45 53 67 77 91 Talks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107 The Star Formation History of Nearby Dwarf Galaxies: problems, methods and results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Antonio Aparicio 107 Diffuse Hot Gas in Dwarf Galaxies Dominik J. Bomans . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111 The Young Large-Scale Features in the Large Magellanic Cloud Jochen M. Braun . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115 A detailed view of the Magellanic Clouds in the FIR M. Braun, R. Assendorp, Tj.R. Bontekoe, D.J.M. Kester, and G. Richter . . . . . . . . 121 Bow Shock Induced Star Formation in the LMC: a Large Scale View Klaas S. de Boer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125 6 Preface III The binary star cluster SL 538/ NGC 2006 (SL 537) Andrea Dieball and Eva K. Grebel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129 Tidal Dwarf Galaxies P.-A. Duc, U. Fritze-v. Alvensleben, and P. Weilbacher . . . . . . . . . . . . . . . . . . . . . . . . . . 133 The ATCA Radio-continuum Investigation of the Magellanic Clouds Miroslav D. Filipovi´c and Lister Staveley-Smith . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137 Tidal Dwarf Galaxies: Their Present State and Future Evolution Uta Fritze - v. Alvensleben and Pierre-Alain Duc . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141 The star formation history of Leo I Carme Gallart, Antonio Aparicio, Wendy Freedman, Giampaolo Bertelli, and Cesare Chiosi . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147 The Recent Star Formation History of the Large Magellanic Cloud Eva K. Grebel and Wolfgang Brandner . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151 Cold dust in NGC 205 Martin Haas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155 Is the metallicity and the luminosity related for dwarf irregular galaxies? A.M. Hidalgo-G´amez and K. Olofsson . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157 The stellar census of the nearby BCDG VII ZW 403 with HST Ulrich Hopp, Regina E. Schulte-Ladbeck, and Mary M. Crone . . . . . . . . . . . . . . . . . . . 161 H i Properties of new nearby Dwarf Galaxies from the Karachentsev Catalog W.K. Huchtmeier . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165 An H i Aperture Mosaic Survey of the Large Magellanic Cloud Sungeun Kim, Lister Staveley-Smith, and Michael A. Dopita . . . . . . . . . . . . . . . . . . . . 169 dSph Satellite Galaxies without Dark Matter: a Study of Parameter Space Pavel Kroupa . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173 Starbursts in dwarf galaxies: Blown out or blown away? Mordecai-Mark Mac Low and Andrea Ferrara . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177 Recent Star Formation in the Wolf-Rayet BCDG MRK 1094 David I. M´endez, L.M. Cair´os, C. Esteban, and J.M. V´ılchez . . . . . . . . . . . . . . . . . . . . 181 Interpreting the H ii Region Luminosity Function M.S. Oey and C.J. Clarke . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185 The first detection of H2 in absorption in the LMC Philipp Richter, Klaas S. de Boer, Dominik J. Bomans, Andreas Heithausen, and Jan Koornneef . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189 Chemodynamical Evolution of Dwarf Galaxies Andreas Rieschick and Gerhard Hensler . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193 HST Study of the Stellar Populations around SN 1987A Martino Romaniello, Nino Panagia, Salvo Scuderi, and the SINS collaboration . . 197 The Role of Galaxy Interactions in H ii Galaxies Christopher L. Taylor, Elias Brinks, and Evan D. Skillman . . . . . . . . . . . . . . . . . . . . . . 201 CO Emission in Low Luminosity Star Forming Galaxies Christopher L. Taylor, Henry A. Kobulnicky, and Evan D. Skillman . . . . . . . . . . . . . 205 Table of Contents 7 The Outer Halo of NGC 4449 Christian Theis and Sven Kohle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209 WFPC2 Observations of Leo A: A Young Galaxy? Eline Tolstoy, J.S. Gallagher, A.A. Cole, J.G. Hoessel, A. Saha, R. Dohm-Palmer, E. Skillman, and M. Mateo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213 The recent star formation history in NGC 1569 M. Tosi, M. Clampin, G. De Marchi, L. Greggio, C. Leitherer, and A. Nota . . . . . 217 The Violent Interstellar Medium of IC 2574 Fabian Walter, Elias Brinks, Neb Duric, Jurgen Kerp, and Uli Klein . . . . . . . . . . . . 221 Posters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 225 First results of a CO survey of dwarfs Marcus Albrecht, Roland Lemke, and Rolf Chini . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 225 Preliminary Results on the Resolved Stellar Population of I Zw 18 A. Aloisi, L. Greggio, M. Tosi, M. Clampin, A. Nota, and M. Sirianni . . . . . . . . . . . 227 Hot subdwarfs, their kinematics and their galactic distribution Martin Altmann, Yolanda Aguilar-S´anchez, Klaas S. de Boer, Michael Geffert, Michael Odenkirchen, and Jacques Colin . . . . . . . . . . . . . . . . . . . . . . . 229 Structure and stellar content of dwarf galaxies: Surface photometry of dwarf galaxies in the M81 & M101 groups Torbjørn Bremnes, Bruno Binggeli, and Philippe Prugniel . . . . . . . . . . . . . . . . . . . . . . . 231 H i-Velocity-Bridges in the Magellanic Stream Christian Bruns, Peter Kalberla, and Ulrich Mebold . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233 Multiband Analysis of the Brightness Distribution of a Sample of Blue Compact Dwarf Galaxies L.M. Cair´os and J.M. V´ılchez . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235 Recent radio and polarization observations of dwarf irregulars Krzysztof T. Chyzy, Sven Kohle, Rainer Beck, Uli Klein, and Marek Urbanik . . . 237 DeNIS Observations on the Magellanic Clouds Maria Rosa Cioni, Cecile Loup, Harm J. Habing, and Erik R. Deul . . . . . . . . . . . . . . 239 New H i shells in the Milky Way Sona Ehlerov´a and Jan Palous . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 241 The Formation and Evolution of Rich Star Clusters in the LMC: NGC 1818 Rebecca A.W. Elson, Gerard Gilmore, Sverre Aarseth, Melvyn Davies, Steinn Sigurdsson, Basilio Santiago, and Jarrod Hurley . . . . . . . . . . . . . . . . . . . . . . . . . 243 The Local Group at High Redshift R. Elson, R. Abraham, T. Kodama, B. Poggianti, and M.S. Oey . . . . . . . . . . . . . . . . 245 Dwarf Galaxies in the Centaurus A Group: Interacting with their Environment? T. Fritz, U. Mebold, U. Klein, and S. Cot´e . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247 Observational discovery of the AGB-bump in the color-magnitude diagram of nearby galaxies Carme Gallart and Giampaolo Bertelli . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249 8 Preface III Be Stars and the IMF of Young Clusters Eva K. Grebel, Wolfgang Brandner, and Andrea Dieball . . . . . . . . . . . . . . . . . . . . . . . . 251 First Results from the Magellanic Cloud Photometric Survey Eva K. Grebel, Dennis Zaritsky, Jason Harris, and Ian Thompson . . . . . . . . . . . . . . . 253 The Center of the Fornax Cluster: Dwarf Galaxies, cD Halo and Globular Clusters Michael Hilker, Leopoldo Infante, Markus Kissler–Patig, and Tom Richtler . . . . . . 255 Identification of Be stars in the Magellanic Clouds Dirk Hoffmann . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 257 Neutral Hydrogen in a sample of selected Blue Compact Dwarf Galaxies W.K. Huchtmeier, Gopal Krishna, and A. Petrosian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 259 NIR-Imaging Of Dwarf Galaxies Marcus Jutte and Rolf Chini . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 261 Comparison of H i and magnetic field structures in our Galaxy H.-R. Klockner, P.M.W. Kalberla, and U. Klein . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 263 A CO map of NGC 4449 S. Kohle, U. Klein, C. Henkel, and D.A. Hunter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 265 A Relation between Box/Peanut Bulges and their Satellite Systems Rainer Lutticke and Ralf-Jurgen Dettmar . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267 ROSAT observations of the giant H ii complex N 11 in the LMC Mordecai-Mark Mac Low, Thomas H. Chang, You-Hua Chu, Sean D. Points, R. Chris Smith, and Bart P. Wakker . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 269 The young population of the Local Group dwarf galaxy Phoenix D. Mart´ınez-Delgado and A. Aparicio . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 271 The cool atomic gas in the Large Magellanic Cloud M. Marx-Zimmer, U. Herbstmeier, F. Zimmer, J.M. Dickey, Staveley-Smith, and U. Mebold . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 273 CO emission toward H i absorption sources in the Large Magellanic Cloud M. Marx-Zimmer, F. Zimmer, U. Herbstmeier, Y.-N. Chin, J.M. Dickey, and U. Mebold . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 275 ISO-[C ii]-investigation of cool H i clouds in the Large Magellanic Cloud M. Marx-Zimmer, U. Herbstmeier, F. Zimmer, J.M. Dickey, and U. Mebold . . . . . 277 Multi-Color Surface Photometry of Blue Compact Dwarf Galaxies K.G. Noeske, P. Papaderos, K.J. Fricke, and T.X. Thuan . . . . . . . . . . . . . . . . . . . . . . . . 279 The Near-IR Tully-Fisher Relation for Giant and Dwarf Late-type Galaxies Daniele Pierini and Richard Tuffs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281 The Kinematic Structure of the Supergiant Shell LMC 2 Sean D. Points . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 283 Mass Segregation in Young LMC Clusters: NGC 2004 and NGC 2031 T. Richtler, P. Fischer, M. Mateo, C. Pryor, and S. Murray . . . . . . . . . . . . . . . . . . . . . 285 A photometric method to study the Wolf-Rayet content of compact H ii regions in nearby galaxies P. Royer, J.-M. Vreux, and J. Manfroid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 287 Table of Contents 9 Does the low-mass galaxy IC 2574 obey MOND? F.J. S´anchez-Salcedo and A.M. Hidalgo-G´amez . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 289 U BV RI Photometry of Supergiants in Multiple Star Systems and the OB Association LH 10 Th. Schmidt-Kaler, J. Gochermann, J. Middendorf, and W.F. Wargau . . . . . . . . . . 291 Red Supergiants as a Tool for the Study of Galaxies Th. Schmidt-Kaler . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 293 Photoelectric U BV Photometry of Galactic Foreground and LMC Member Stars: A New Database Th. Schmidt-Kaler, J. Gochermann, M.O. Oestreicher, H.-G. Grothues, C. Tappert, A. Zaum, Th. Berghofer, and H.R. Brugger . . . . . . . . . . . . . . . . . . . . . . . . . 295 Interacting and merging processes between spirals and dwarf galaxies Uwe Schwarzkopf and Ralf-Jurgen Dettmar . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 297 Holes and Shells in IC 2574 Fabian Walter, Elias Brinks, and Uli Klein . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 299 H i study of the dwarf galaxy DDO 47 Fabian Walter, Elias Brinks, and Uli Klein . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 301 Bubbles around massive stars in LMC H ii regions Kerstin Weis and Wolfgang J. Duschl . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303 Spatial distribution of halo high velocity clouds towards the LMC Bernhard Wierig and Klaas S. de Boer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 305 Appendices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 307 Conference Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Photograph of Participants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . List of Participants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Final Programme . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Glossary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Author Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . List of Objects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Subject Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Sample Name Label . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Technical Information . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Erratum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 307 311 313 319 323 327 331 337 EV EV EV EV – only part of the electronic version available at the WWW URL: http://www.astro.uni-bonn.de/webgk/ws98/cover.html page numbers refer to the printed version published by Shaker Verlag, Aachen, ISBN 3-8265-4457-9 Conference Summary Evan D. Skillman University of Minnesota This is not formally a conference summary. Foremost, the hosts are thanked and congratulated. I then share a few impressions and a gedanken experiment is proposed. 1. Mandatory Laudatory Comments It has been my great pleasure to be a participant in this workshop. The Bad Honnef conference center is certainly an exceptional facility which is conducive to interaction. The conference organizers have done a superb job in arrangements so that we haven't had to worry about anything (with the possible exception of the participants who still haven't made it back from Koln yet), and the result has been a wealth of exchange of ideas. I especially appreciated the poster review sessions, which helped to act as an introduction to the many exciting results presented as posters. This meeting represents a solid endorsement of the Graduiertenkolleg (GK), in both its choice of science themes and its organization. When I was a first year graduate student, Bart Bok visited the University of Washington, and repeated over and over, "If you want to study star formation, study the Magellanic Clouds". At the time he was bemoaning the lack of a millimeter observing facility in the southern hemisphere - something which has been rectified. The expansion of the theme from the Magellanic Clouds to dwarf galaxies in general has allowed the work at Bonn and Bochum to reach an even larger audience. Having attended an earlier meeting of the GK, I should state that my impressions are all very positive. The benefits of forming a concentration of studies shared between two university departments are obvious and impressive. The students have access to numerous first class facilities, faculty, and many other resources. In fact, one can say that the students have enhanced access to numerous facilities since the resources available through the GK help to make them more competitive for observing time. When I attended the GK meeting in the Fall of 1995, I was impressed with the many ideas for new projects that students were then forming. Many of those projects have been completed and were presented at this meeting. Good science is being done. The GK experience has become an excellent platform from which to launch a career in astrophysics. 2. Some Impressions Some of the most successful conference summaries that I have witnessed were, in fact, not summaries at all. Thus, I will not attempt to summarize the whole conference, but rather to present some of my impressions gained during the workshop. Of course, the overall impression is that by concentrating on the nearest galaxies one can gain insights unafforded in any other context. Sometimes we may feel that this is lost in the rush to high redshift, but it certainly is not lost on the participants of this workshop. Again and again we were treated to stunning views of the Magellanic Clouds in wavelength after wavelength. There is a lot still to be learned from the Magellanic Clouds and even more to be learned by pushing for comparable studies of other nearby galaxies. Conference Summary 11 Simply because of my recent research activities, many of my impressions are grouped under the category of stellar populations. First among them is that the HST is an instrument ideally suited for stellar population studies. The leaps forward afforded by the vast improvement over ground based resolution has revolutionized our view of galaxies. Nowhere is this more true than for the dwarf spheroidal galaxies. In the last decade, our views of these "simple" systems has changed enormously. The variety of star formation histories is still begging for a simple explanation. The kinematic studies of literally hundreds of stars (many now with multiple epochs) in most of the dSphs have continued to bear out Aaronson's bold claim that these are dark matter dominated systems. As spectra of individual stars provide detailed abundance analyses, we may still be in for more surprises. I was very impressed by the degree of sophistication in treatment of color-magnitude-diagrams (CMDs) and the general agreement between independent groups. For example, the presentations by Aparicio and Gallart showed how the different features in the CMDs could be used in concert to place strong constraints on star formation histories reaching back to the earliests epochs. Tolstoy and Hopp independently studied HST data on VII Zw 403 and (1) came to very similar conclusions about basic properties and (2) agreed on the features in the CMD which were most difficult to fit with a range of reasonable models. There was a great deal of attention paid to the detailed structure in the ISM and its origins. I will never again confuse the terms "bubbles", "super-bubbles", and "supergiant shells". In the detailed H i imaging of the Magellanic Clouds, we are confronted with the problem of how best to deal with the high degree of complexity. Does it make sense to divide everything into categories of "holes" and "concentrations"? Perhaps it is pertinent to reflect back on the work that Hodge conducted for many decades, providing H ii region luminosity functions (which also required making decisions about whether to divide features into components or combine features into single entities). We have seen that Oey has taken these H ii region luminosity functions and used them to distinguish between "saturated" and "unsaturated" IMFs. While the analogy is not sound on a physical basis, I wonder that we may not see a similar understanding of the H i holes and concentrations in the future. Concerning ISM structure, it is evident that x-ray observations are rapidly playing an increasing role in our understanding of the phase structure of the ISM. The x-ray images available today remind me of the H i images that were available a few decades ago. As the spatial and spectral resolution and sensitivity increase, important new insights are bound to come from this waveband. Also, concerning the bubbles, I was disappointed that after presenting a detailed star formation history of the Sextans A dwarf, no one asked about the connection between the stellar population and the H i hole. I tried to check on this in preparing for this summary talk, but the limited field of view of the HST WFPC2 observations didn't allow for a clear answer. Recently, van Dyk has shown evidence from ground based data for a radial gradient in stellar ages that supports a wind blown bubble model for the central H i deficit. Moving to galaxy scales, the numerical calculations presented by Mac Low were very impressive. There has been endless speculation concerning the effects of star formation on the ISM of dwarf galaxies, and, in particular, the ability of a galaxy to blow away all of its ISM. These new calculations appear to present a more realistic impression of what is possible and what is not. This leads to the connected question of what is the H i that is frequently found in the vicinity of dwarf galaxies. I refer to this H i which is usually at similar velocity but disconnected from the normal velocity fields, as "floppy disks". Two alternative origins immediately come to mind. It could be that this material is tidal debris, left over from an interaction or a merger. Alternatively, this could be material which is primordial in nature and has not yet been incorporated into the galaxy. Within the limitations of available observing facilities, it is difficult to design an observational test which distinguishes these two possibilities. It is, however, a very important question. 12 E.D. Skillman - Appendix A Table 1. A Summary of Impressions 1) DH TR DE ES 2) MM P BO BIP 3) ET UH FSP 4) NO AA SA 5) XR HY 6) HST BF SPS IBVC BSS 7) 8) DS EBKS 9) A H&C LHR 10) NN RFP 11) FD TR PG Finally, I have to admit that one of the most remarkable things that I heard during the entire workshop was the story of Hunter's graduate school foreign language exam. To pass the exam, she was required to translate a scientific article written in German. In fact, she translated the article into Spanish. I was impressed not only with the demonstration of language skills, but the display of stubborn independence found in astronomers everywhere. This presentation of my impressions was accompanied by a visual aid. Uli Klein asked me to reproduce it in my conference summary, so it appears here as Table 1. 3. A Gedanken Experiment Finally, knowing that most of the workshop participants would be spending one more night at Bad Honnef, I proposed an exercise to be carried out after dinner. At the time I called it an exercise, but since this is Germany, it is probably better referred to as a gedanken experiment. The conference participants were treated to a trip to Koln, the highlight of which was a tour of the Dom. The Dom had recently experienced some refurbishing work, as evidenced by the many ton scaffold that was still hanging in the top of the church. The following exercise occurred to me. What if that scaffold had fallen and crushed some of the workshop participants? For this exercise, I assume the following cosmology: (1) there is a heaven, (2) there is a single deity, (3) when good people go to heaven the deity answers their questions, and (4) all astronomers are good. The purpose of the exercise is not to quibble with my assumed cosmology, but, rather, to imagine the questions of the deceased workshop participants as they enter heaven. To kick off the proposed exercise, I tackled the question what if all of the invited review speakers were done in? I started with Jay Gallagher. (The main reason for this was because I wasn't sure how this would go over and Jay had already left for Munich.) Jay's talk highlighted the degree of the interconnections between different studies and how observations of dwarf galaxies had impact on literally every major question in astronomy today. Actually, I think that Jay's question was very easy. He would ask "What are the answers to everyone else's questions?" Mario Mateo's talk emphasized the sophistication of the measurements of the dark matter associated with dSphs and their complicated star formation histories. I imagine that he would be able to negotiate a second question and the two together would be "What is dark matter, really?" and "Concerning dwarf spheroidals, where did the gas go?" Elias Brinks has always been fascinated with detailed images of the distribution of neutral hydrogen in galaxies. On one hand, the energetics of the holes can be explained by the presence of stellar clusters. On the other hand, few of the known holes have identifiable stellar populations in their interiors. Elias would ask "How are these holes formed?" Conference Summary 13 Deidre Hunter presented detailed observations of star formation on different scales. In one case, she pointed to a sequence of two generations of star formation and proposed that a third might be ready to occur. A theory of star formation should have predictive power, and thus, Deidre might ask "Is there going to be a third generation of star formation?" You-Hua Chu presented detailed comparisons of the x-ray distributions and the gas kinematics in the Magellanic Clouds. She questioned the assumptions of pressure equilibrium. You-Hua would simply demand a complete picture of the multi-phase ISM in the Large Magellanic Cloud. Jan Palous showed us, through numerical simulations, how ISM bubbles evolve. The calculations are simplified by taking advantage of certain symmetries. He would like to know if 2-dimensional simulations give the correct insight into a three dimensional world. Trinh Thuan presented observations of SBS 0335 − 052, which he proposes is a young galaxy which is just forming now. He would like to know if it is possible for galaxies to be forming in the current epoch. In summary, I would like to thank the organizers one last time for a truly enjoyable workshop. I congratulate them on their 25th GK meeting and I look forward to the 50th! List of Participants Marcus Albrecht (13) Astronomisches Institut der Ruhr-Universitat Bochum [email protected] Alessandra Aloisi (67) Dipartimento di Astronomia, Universit´a di Bologna [email protected] Martin Altmann (11) Sternwarte der Universitat Bonn [email protected] Christian Bruns (50) Radioastronomisches Institut der Universitat Bonn [email protected] Luz Marina Cair´os Barreto (57) Instituto de Astrof´ısica de Canarias (IAC) [email protected] Rolf Chini (62) Astronomisches Institut der Ruhr-Universitat Bochum [email protected] Ralph-P. Andersen (5) Max-Planck-Institut fur Astronomie (MPIA), Heidelberg [email protected] You-Hua Chu (-) Astronomy Department, University of Illinois [email protected] Antonio Aparicio (40) Instituto de Astrof´ısica de Canarias (IAC) [email protected] Dominik J. Bomans (12) Astronomisches Institut der Ruhr-Universitat Bochum [email protected] Jochen M. Braun (32) Sternwarte der Universitat Bonn [email protected] Michael Braun (63) ESA-VILSPA Villafranca del Castillo Satellite Tracking Station [email protected] Torbjørn Bremnes (29) Astronomisches Institut der Universitat Basel [email protected] Elias Brinks (39) Departamento de Astronomia, Universidad de Guanajuato [email protected] Krzysztof Tadeusz Chy zy (23) Astronomical Observatory, Jagiellonian University [email protected] Maria Rosa Cioni (20) Sterrewacht Leiden [email protected] Oliver-Mark Cordes (60) Sternwarte der Universitat Bonn [email protected] Klaas S. de Boer (43) Sternwarte der Universitat Bonn [email protected] Ralf-Jurgen Dettmar (71) Astronomisches Institut der Ruhr-Universitat Bochum [email protected] Andrea Dieball (54) Sternwarte der Universitat Bonn [email protected] Participants 15 Pierre-Alain Duc (44) European Southern Observatory (ESO), Garching [email protected] Christian Dusterberg (70) Radioastronomisches Institut der Universitat Bonn [email protected] Harm Jan Habing (48) Sterrewacht Leiden, Huygens Laboratorium [email protected] Andreas Heithausen Radioastronomisches Institut der Universitat Bonn [email protected] Sona Ehlerov´a (1) Astronomical Institute, Academy of Sciences of the Czech Republic [email protected] Fabian Heitsch (72) Sternwarte der Universitat Bonn [email protected] Rebecca A.W. Elson (38) Institute of Astronomy, Cambridge [email protected] Ana Maria Hidalgo-G´amez (58) Astronomiska Observatoriet, Uppsala [email protected] Miroslav D. Filipovi´c (66) University of Western Sydney Nepean / MPE, Garching [email protected] Michael Hilker (31) Departamento de Astronom´ıa y Astrof´ısica, P. Universidad Cat´olica, Santiago, Chile [email protected] Thomas Fritz (49) Radioastronomisches Institut der Universitat Bonn [email protected] Uta Fritze - v. Alvensleben (45) Universitats-Sternwarte Gottingen [email protected] John S. Gallagher (24) Department of Astronomy, University of Wisconsin at Madison [email protected] Carme Gallart (42) Carnegie Observatories [email protected] Eva K. Grebel (68) UCO/Lick Observatory, University of California at Santa Cruz [email protected] Martin Haas (4) Max-Planck-Institut fur Astronomie (MPIA), Heidelberg [email protected] Dirk Hoffmann (15) Astronomisches Institut der Ruhr-Universitat Bochum [email protected] Ulrich Hopp (14) Universitats-Sternwarte Munchen [email protected] Walter Huchtmeier (17) Max-Planck-Institut fur Radioastronomie (MPIfR), Bonn [email protected] Deidre A. Hunter (46) Lowell Observatory, Flagstaff [email protected] Susanne Huttemeister Radioastronomisches Institut der Universitat Bonn [email protected] Norbert Junkes (61) Max-Planck-Institut fur Radioastronomie (MPIfR), Bonn [email protected] 16 Appendix C Marcus Jutte (34) Astronomisches Institut der Ruhr-Universitat Bochum [email protected] Sungeun Kim (41) Mount Stromlo and Siding Spring Observatories [email protected] Uli Klein (36) Radioastronomisches Institut der Universitat Bonn [email protected] Hans-Rainer Klockner Radioastronomisches Institut der Universitat Bonn [email protected] Joachim Koppen (-) Equipe 'Evolution Galacticque', Observatoire de Strasbourg [email protected] Sven Kohle (53) Radioastronomisches Institut der Universitat Bonn [email protected] Pavel Kroupa (33) Institut fur Theoretische Astrophysik der Universitat Heidelberg [email protected] Harald Lesch (19) Universitats-Sternwarte Munchen [email protected] Rainer Lutticke (51) Astronomisches Institut der Ruhr-Universitat Bochum [email protected] Mordecai-Mark Mac Low (3) Max-Planck-Institut fur Astronomie (MPIA), Heidelberg [email protected] David Mart´ınez-Delgado (55) Instituto de Astrof´ısica de Canarias (IAC) [email protected] Monika Marx-Zimmer Radioastronomisches Institut der Universitat Bonn [email protected] Mario L. Mateo (8) Department of Astronomy, University of Michigan [email protected] Ulrich Mebold Radioastronomisches Institut der Universitat Bonn [email protected] David Israel M´endez Alcaraz (56) Instituto de Astrof´ısica de Canarias (IAC) [email protected] Andrea Moneti (30) ESA-VILSPA Villafranca del Castillo Satellite Tracking Station [email protected] Kai Noeske (28) Universitats-Sternwarte Gottingen [email protected] Marion Siang-Li Oey (74) Institute of Astronomy, Cambridge [email protected] Jan Palous (10) Astronomical Institute, Academy of Sciences of the Czech Republic [email protected] Nino Panagia (7) Space Telescope Science Institute (STScI), Baltimore [email protected] Daniele Pierini (18) Max-Planck-Institut fur Kernphysik, Heidelberg [email protected] Sean Points (52) Astronomy Department, University of Illinois [email protected] Participants 17 Gotthard Richter Astrophysikalisches Institut Potsdam (AIP) [email protected] Philipp Richter Sternwarte der Universitat Bonn [email protected] Tom Richtler (16) Sternwarte der Universitat Bonn [email protected] Andreas Rieschick (64) Institut fur Theoretische Physik und Astrophysik, Kiel [email protected] Martino Romaniello (69) STScI, Baltimore, USA and Scuola Normale Superiore, Pisa, Italy [email protected] Pierre Royer (47) Institut d'astrophysique de Li`ege [email protected] Theodor Schmidt-Kaler Astronomisches Institut der Ruhr-Universitat Bochum Chris Taylor (21) Astronomisches Institut der Ruhr-Universitat Bochum [email protected] Christian Theis (65) Institut fur Theoretische Physik und Astrophysik, Kiel [email protected] Trinh X. Thuan (26) Department of Astronomy, University of Virginia [email protected] Eline Tolstoy (25) European Southern Observatory (ESO), Garching [email protected] Monica Tosi (22) Osservatorio Astronomico di Bologna [email protected] Hugo van Woerden (59) Kapteyn Institute, Groningen [email protected] Fabian Walter (35) Radioastronomisches Institut der Universitat Bonn [email protected] Uwe Schwarzkopf (6) Astronomisches Institut der Ruhr-Universitat Bochum [email protected] Kerstin Weis (2) Institut fur Theoretische Astrophysik, Heidelberg [email protected] Wilhelm Seggewiss (73) Sternwarte der Universitat Bonn / Observatorium Hoher List [email protected] Evan D. Skillman (27) Astronomy Department, University of Minnesota [email protected] Axel Weiss (37) Radioastronomisches Institut der Universitat Bonn [email protected] Bernhard Wierig (9) Sternwarte der Universitat Bonn [email protected] 18 Appendix C Hans Zinnecker Astrophysikalisches Institut Potsdam (AIP) [email protected] The list above shows the names of the participants, the numbers belonging to the photograph on p. 311 (in brackets, ciphers for invited speakers additionally set in boldface), the names of the institutes and the E-Mail addresses. The full list of participants including all links is available at the WWW URL: http://www.astro.uni-bonn.de/ webgk/ws98/part a tbl.html.
astro-ph/0212055
1
0212
2002-12-03T04:18:52
Relativistic Effects on the Appearance of a Clothed Black Hole
[ "astro-ph" ]
For an accretion disk around a black hole, the strong relativistic effects affect every aspect of the radiation from the disk, including its spectrum, light-curve, and image. This work investigates in detail how the images of a thin disk around a black hole will be distorted, and what the observer will see from different viewing angles and in different energy bands.
astro-ph
astro-ph
RELATIVISTIC EFFECTS ON THE APPEARANCE OF A CLOTHED BLACK HOLE Xiaoling Zhang, S.Nan Zhang, Yuxin Feng, Yangsen Yao Physics Department, University of Alabama in Huntsville, Huntsville, AL, USA National Space Science and Technology Center, Huntsville, AL, USA [email protected] Abstract For an accretion disk around a black hole, the strong relativistic effects affect every aspect of the radiation from the disk, including its spectrum, light-curve, and image. This work investigates in detail how the images will be distorted, and what the observer will see from different viewing angles and in different energy bands. Keywords: Relativistic effects, black hole -- image Introduction Black holes are believed to be very common in the universe. A black hole itself, as indicated by its name, is invisible. But the interactions between a black hole and other objects can generate spectacular phe- nomena and make the hole "visible". A "clothed black hole", a black hole with an accretion disk (Luminet, 1978), is one of the most important laboratories to study relativity. Strong relativistic modifications on energy spectra, light curves, power spectra etc., have been observed with the many detectors on the ground and in space, yet we have never observed any black hole image due to the limited spatial resolution of currently available instruments. People have been curious about what a black hole or a clothed black hole would look like. Luminet (Luminet 1978) studied the Schwarzschild black hole case and gave a simulated photograph of a spherical black hole with a thin accretion disk. Fanton et al.(1997) developed a semi-analytical approach and a code of calculating photon trajectories in Kerr metric, calculated the redshift maps of both direct and secondary images, and the temperature map. We used their code in the calculation of the disk images in this work. 1 2 The observable disk image is actually the flux map. However, the to- tal flux (in the whole energy range) map might not be a proper object, because each instrument has its sensitive energy range; it can only de- tect photons in this range, thus instruments sensitive to different energy ranges will see different pictures. In section 1 we demonstrate how the relativistic effects in a Kerr space will work on various elements of a black hole system. In section 2 we give the images in different energy ranges, followed by summary and discussion. 1. Red-shift maps of disks Figures 1, 2 and 3 are all redshift maps. Figures 1 shows the redshift c2 ) and an inner maps of disks with an outer radius 20R radius as the last stable orbit. For a small viewing angle, there is only significant (gravitational) red-shift in the inner region; whereas for large viewing angles, Doppler shift becomes important, significant blue- and red-shift can be seen on the whole disk. (It would be clearer in color figures.) Also, for large viewing angles, far end of the disk bends toward the observer. g (R GM ≡ g Red-shift map and over- Figure 1. all shape of accretion disks around black holes. g is the ratio between the observed energy of a photon and the energy in the rest frame of the disk. Top to bottom: θ = 5◦, 45◦, 70◦, 85◦, θ is the inclination angle (the angle between the normal of the disk and the line of sight). Left to right: a = 0.01, 0.5, 0.998, where a is the specific angular momentum of the black hole. The white zones stand for the regions with zero red-shift (following Fanton et al.(1997) ). Left-hand side of the disk is approaching the observer and blue-shifted. Figure 2 shows the overall shape of disks due to Doppler and gravita- tional effects at a large inclination angle (85◦). The inner radius takes the value of the last stable circular orbit. In the first row, the white circles and ellipses show the last stable orbit. The inner contours are greatly distorted by the focusing effect. Figure 3 shows the comparison between Shwarzschild black holes and Kerr black holes, which shows clearly additional distortion due to frame- dragging effect. Spin results in a smaller last stable orbit, and the inner contour changes from symmetric for a = 0.01 to asymmetric for a = Relativistic Effects on the Appearance of a Clothed Black Hole 3 0.998. Therefore in principle, we can tell a rotating black hole from a non-rotating one from the inner contours of their accrection disks. Redshift maps for disks Figure 2. at inclination angle 85◦ with different outer radii. Top to bottom: disks with outer radii 10Rg, 20Rg, 100Rg. Left to right: a = 0.01, 0.5, 0.998. See text for details. Redshift maps for 10 Rg Figure 3. disks with different inner radii. Top inner radii are the radii of the row: last stable orbits; Bottom row: inner radii are 6 Rg. Left to right: a = 0.01, 0.5, 0.998. 2. Images of thin accretion disks With the redshift map, we calculated images of accretion disks around black holes in several energy ranges. In the calculations, we assume 1) the disks are standard thin disks (Shakura & Sunyaev, 1973); and 2) lo- cal radiation results from gravitational potential energy release through blackbody radiation (Page & Thorne 1974). The inner disk radii are the respective last stable orbits. a) a = 0.01 b) a = 0.5 Figure 4. Disk images of 20Rg disks for different spin of the black holes, viewing in certain energy ranges. In all figures, same energy range in a column, and same inclination angle in a row. Left to right: 0.01 -- 100, 0.1 -- 1, 1 -- 2, 2 -- 5, 5 -- 10. Top to bottom: θ = 5◦, 45◦, 70◦, 85◦, 85◦. The last image row of c) is the same as the row above it, expect that it is in log scale. c) also shows the scale bar. c) a = 0.998 4 In all the images, energy ranges are in the unit of the energy cor- responding to the maximum temperature in a non-rotating black hole case. For example, if the maximum temperature in a non-rotating black hole case is 1 keV, then the above energy ranges are 0.1 to 1 keV, 1 to 2 keV etc.. The maximum temperature in the extremal Kerr black hole case is about 5 keV. The images show the brightness in gray scale, with energy flux linearly (if not otherwise specified) mapped to the [0,1] shown on the scale bar in Figure 4c. Figure 5. Disk images for 100Rg disks in the energy range 0.5 -- 2. Top to θ = 5◦, 45◦, 70◦, 85◦. bottom: Left to right: a = 0.01, 0.5, 0.998 respec- tively. 0.5 to 2 keV is the energy range of NASA's next generation mission MAXIM (Micro-Arcsecond X-ray Imag- ing Mission) pathfinder. (The images in this figure have a lower contrast than the other images.) 3. Summary and discussion We studied the images of standard thin accretion disks around black holes under the strong relativistic effects. The expected images for sys- tems with different black hole spin, viewing from different viewing angles and in different energy bands are given. The spin of the black hole not only determines how close the disk can extend to the center, but also causes additional distortion due to the dragging of the frame. The secondary image are left out because in most cases it will be blocked by the disk, though it might be important in some cases. Acknowledgments The authors would like to thank Dr. Fanton for providing us his code of calculating the redshift map and line profile of a thin disk. This study is supported in part by NASA's Marshall Space Flight Center and through NASA's Long Term Space Astrophysics Program. References Cadez, A., C. Fanton, and M. Calvani, M. (1998). New Astronomy, 3, (No. 8), 647 Fanton, Claudio, Massimo Calvani, Fernando de Felice, and Andrej Cadez. (1997). PASJ, 49, 159 Luminet, J.-P.. (1978). A&A, 75, 228 Page, D.N. & K.S. Thorne. 1974. ApJ, 191, 499 Shakura, N. I. & Sunyaev, R. A.. 1973. A&A, 24, 337 This figure "figure1.png" is available in "png"(cid:10) format from: http://arxiv.org/ps/astro-ph/0212055v1 This figure "figure2.png" is available in "png"(cid:10) format from: http://arxiv.org/ps/astro-ph/0212055v1 This figure "figure3.png" is available in "png"(cid:10) format from: http://arxiv.org/ps/astro-ph/0212055v1 This figure "figure4a.png" is available in "png"(cid:10) format from: http://arxiv.org/ps/astro-ph/0212055v1 This figure "figure4b.png" is available in "png"(cid:10) format from: http://arxiv.org/ps/astro-ph/0212055v1 This figure "figure4c-scaleBar.png" is available in "png"(cid:10) format from: http://arxiv.org/ps/astro-ph/0212055v1 This figure "figure4c.png" is available in "png"(cid:10) format from: http://arxiv.org/ps/astro-ph/0212055v1 This figure "figure5.png" is available in "png"(cid:10) format from: http://arxiv.org/ps/astro-ph/0212055v1
astro-ph/9411021
1
9411
1994-11-05T17:23:17
The EINSTEIN On-Line Service
[ "astro-ph" ]
The Einstein On-Line Service (EOLS) is a simple menu-driven system which provides an intuitive method of querying over one hundred database catalogs. In addition, the EOLS contains over 30 CDROMs of images from the Einstein X-ray Observatory which are available for downloading. The EOLS provides all of our databases to the NASA Astrophysics Data System (ADS) and our documents which describe each table are written in the ADS format. In conjunction with the IAU working group on Radioastronomical Databases, the EOLS serves as an experimental platform for on-line access to radio source catalogs. The number of entries in these catalogs exceeds half a million.
astro-ph
astro-ph
Proc. Astronomical Data Analysis Software and Systems { IV, Baltimore, Sept. {,   ASP Conference Series ( ) , eds. H. Payne, D. Shaw, & J. Hayes The EINSTEIN On-Line Service D. E. Harris and C. S. Grant Center for Astrophysics,  Garden St., Cambridge, MA  H. Andernach Observatoire de Lyon, F{  Saint-Genis-Laval Cedex, France Abstract. The Einstein On-Line Service (EOLS) is a simple menu- driven system which provides an intuitive method of querying over one hundred database catalogs. In addition, the EOLS contains over  CDROMs of images from the Einstein X-ray Observatory which are avail- able for downloading. The EOLS provides all of our databases to the NASA Astrophysics Data System (ADS) and our documents which de- scribe each table are written in the ADS format. In conjunction with the IAU working group on Radioastronomical Databases, the EOLS serves as an experimental platform for on-line access to radio source catalogs. The number of entries in these catalogs exceeds half a million. . INTRODUCTION In January   , SAO established an on-line service to help astronomers prepare ROSAT proposals by providing access to the preliminary source list from the \Einstein Observatory Catalog of IPC X-ray Sources". In the intervening years, we have updated the source list, added to the documentation, included many more Einstein databases as well as a number of tables from other wavebands, provided access to images for downloading from all of the Einstein CDROMs, and installed new software for more sophisticated (cid:12)ltering and retrieval. Although we have improved the functionality and made signi(cid:12)cant additions to the databases, we still maintain a simple menu interface accessible from any type of terminal. An instruction manual does not exist: the on-line help facility provides the necessary information. . DATABASE RETRIEVAL OPTIONS The EOLS provides three options for data retrieval. The DATABASE QUERY is a menu-driven system which allows users to build their own query and select which output columns should be retrieved. The QUICK QUERY is a canned output format which requires only one speci(cid:12)cation of selection criteria, and returns up to  characters per row (preselected by us). The MULTIPLE QUICK QUERY is a positional search on several tables. We have endeavored to make  the selection process simple and the easiest option is to select one or more a(cid:14)nity groups. However, the user can specify particular tables from di(cid:11)erent groups by choosing the \Custom" option or choose to select all tables. Here are  of the  rows retrieved from  catalogs with a single query: --------------------------------- eos_source ---------------------------------- SEQ FLDCAT RA DEC +/-COR C/S # # # h m s d ' ''asccts/sec +/- S/N SIZCORRECOID ------------------------------------------------------------------     .+   .e-.e- . . Q ------------------------------------------------------------------------------- Total:  rows retrieved. ------------------------------ qso_veron ------------------------------ RA DEC radio fdr fdr name h m s d ' '' z vmagsource? (cgs) (cgs) ----------------------------------------------------------------    .+  .. yes.e-.e+   .+  .. yes.e-.e+ ------------------------------------------------------------------------- Total:  rows retrieved. ----------------------------- rad_cm_bwe  ------------------------------ RA DEC S(.GHz)ExtensionSpectralSeparation Name h m s d ' '' (mJy) Flag Index Flag ------------------------------------------------------------------ +   .+    -. -------------------------------------------------------------------------- Total:  rows retrieved. --------------------------------- rad_gb ---------------------------------- RA RA+/- DEC DEC+/-S(.)S +/-EXTWARNCONFFWHMFWHM PA h m s sec d ' '' arcs (mJy) (mJy)FLGFLAGFLAGMAJ MIN ----------------------------------------------------------------   . .+      ..- ----------------------------------------------------------------------------- Total:  rows retrieved. ---------------------------------- rad_pks ---------------------------------- +/- fdr fdr fdr RA DEC arcs id mag z (Jy) (Jy) (Jy) h m s d ' '' alias -------------------------------------------------------------------- Xs .    .+  C. ------------------------------------------------------------------------------- Total:  rows retrieved. . DOWNLOADING X-RAY IMAGES All of our Einstein CDROMs are permanently available for downloading images. They may be obtained with ftp or with the VMS \copy" command. As of  June  , we have: The E Catalog of IPC X-ray Sources: FITS smoothed arrays The Database of HRI X-ray Images: FITS photon arrays The IPC Slew Survey:  Jan   The HRI Images in Event List Format: FITS binary tables The IPC Images in Event List Format: FITS binary tables The IPC Unscreened Data Archive: FITS binary tables  . OTHER FEATURES (cid:15) Low maintenance operation. (cid:15) The ADS connection { The EOLS provides all of our databases to the NASA Astrophysics Data System (ADS) and our documents which de- scribe each table are written in the ADS format. (cid:15) DOCUMENTATION { Documentation is available for each database table and provides extensive detail in some areas (e.g. Vol  of the IPC catalog), archival material such as the Einstein Revised User's Manual, and \Help" (cid:12)les which provide information on system operations. (cid:15) COMMUNICATION { This section of the EOLS features a bulletin board (includes ROSAT news) and an email facility for sending yourself docu- ments and the results of database searches. (cid:15) TOOLS { Precession of coordinates and column densities of galactic HI (north of DEC = - degrees) are available. (cid:15) ACCESS { There are four methods to log onto the EOLS: . internet { telnet einline.harvard.edu . decnet { set host  . modem { dial ()  - or  - . mosaic { http://hea-www.harvard.edu/einline/einline.html Once successful, the login name is \einline", no password required. NB: Since UNIX is case sensitive, make sure you use lower case when logging in! . THE RADIO SOURCE CATALOG INITIATIVE For many years we have noticed the di(cid:14)culty of (cid:12)nding radio source data in public archives or on-line databases. In  , the IAU Working Group on Ra- dioastronomical Databases was formed with one of us (H.A.) as the chair. By now the WG has secured some  catalogs of radio sources with altogether (cid:24), entries. Although the working group recognized that a long term so- lution to this problem should be provided by an on-going commitment from the National Radio Astronomy Observatory (NRAO) or the Strasbourg Astro- nomical Data Center (CDS), the EOLS volunteered to serve as an experimental platform for access to the catalog collection. This \provisional" solution is now by far the most complete on-line facility for radio source data, o(cid:11)ering access to  radio catalogs with , searchable entries. Contributions of new data tables are welcome, but can only be incorporated into the system if authors provide proper documentation and formatted ASCII tables. . STATISTICS OF USAGE The histogram shows the average number of logins per month from outside the Center for Astrophysics from   to mid  .  Figure . Logins per month from outside CfA:   {  
astro-ph/0303280
1
0303
2003-03-13T20:37:45
A Turbulent Interstellar Medium Origin of the Binary Period Distribution
[ "astro-ph" ]
In this paper, we present a semi-empirical model of isolated binary star formation. This model includes the effects of turbulence in the initial state of the gas, and has binary orbital parameters consistent with observation. Our fundamental assumption is that the angular momenta of binary star systems is directly related to the net angular momentum induced by turbulence in parent molecular cloud cores. The primary results of this model are as follows. (i) A quantitative prediction of the initial width of the binary period distribution ($\sigma_{\log {P}_d} = 1.6 - 2.1$ for a star formation efficiency in the range $\epsilon_* = 0.1 - 0.9$). (ii) A robust, negative anticorrelation of binary period and mass ratio. (iii) A robust, positive correlation of binary period and eccentricity. (iv) A robust prediction that the binary separation of low-mass systems should be more closely separated than those of solar-mass or larger. These predictions are in good agreement with observations of PMS binary systems with periods $P > 10^3$ d, which account for the majority of all binaries. We conclude with a brief discussion of the implications of our results for observational and theoretical studies of multiple star formation.
astro-ph
astro-ph
A Turbulent Interstellar Medium Origin of the Binary Period Distribution Robert T. Fisher [email protected] Lawrence Livermore National Laboratory, Mail Code L-023, 7000 East Avenue, Livermore Ca. 94550 Received ; accepted 3 0 0 2 r a M 3 1 1 v 0 8 2 3 0 3 0 / h p - o r t s a : v i X r a -- 2 -- ABSTRACT In this paper, we present a semi-empirical model of isolated binary star for- mation. This model includes the effects of turbulence in the initial state of the gas, and has binary orbital parameters consistent with observation. Our funda- mental assumption is that the angular momenta of binary star systems is directly related to the net angular momentum induced by turbulence in parent molecular cloud cores. The primary results of this model are as follows. (i) A quantitative prediction of the initial width of the binary period distribution (σlog Pd = 1.6−2.1 for a star formation efficiency in the range ǫ∗ = 0.1− 0.9). (ii) A robust, negative anticorrelation of binary period and mass ratio. (iii) A robust, positive correla- tion of binary period and eccentricity. (iv) A robust prediction that the binary separation of low-mass systems should be more closely separated than those of solar-mass or larger. These predictions are in good agreement with observations of PMS binary systems with periods P > 103 d, which account for the majority of all binaries. We conclude with a brief discussion of the implications of our results for observational and theoretical studies of multiple star formation. Subject headings: stars: formation, binaries:general, gravitation, hydrodynamics 1. Introduction To fully elucidate the mechanisms underlying the origin of binary systems, it is crucial to compare theory with observations of the field and of nearby star-forming regions. However, a myriad of complex physical effects have been advanced by theorists to account for the formation and evolution of binary and multiple star systems, including fragmentation -- 3 -- (Hoyle 1953; Inutsuka & Miyama 1992), turbulence (Klein, Fisher, & McKee 2001; Bate, Bonnell, & Bromm 2002), ideal magnetohydrodynamics (MHD) (Galli et al. 2001), non-ideal MHD, including ambipolar diffusion (Mouschovias 1977; Boss 2000), radiative transfer (Boss et al. 2000), dust physics (Whitworth, Boffin, & Francis 1998), tidal torquing (Larson 2002), capture (Clarke & Pringle 1991), competitive accretion (Bonnell, Bate, Clarke, & Pringle 1997), and accretion disk processes (Laughlin & Bodenheimer 1994). However, to date, no simulation has included all of these varied physical effects, and few have been run to a substantial evolutionary state. Moreover, as other authors have pointed out (Larson 2001), once one includes the effect of turbulence, star formation is inherently a stochastic process, implying that many such simulations need to be performed before their results can be meaningfully compared to observation. In this paper, we present a simplified, semi-empirical model of isolated binary star formation. This model includes the effects of turbulence in the initial state of the gas, and has binary orbital parameters consistent with observation. Our fundamental assumption is that the angular momenta of binary star systems is directly related to the net angular momentum induced by turbulence in parent molecular cloud cores. We do not require that the binary angular momentum be equal to that of its parent core; indeed, we shall find that angular momentum loss from the initial core is crucial in determining the median binary period, as previous authors suggested (Simon 1992; Bodenheimer 1995). While we do not in this work determine the mechanism underlying the loss of angular momentum, we find that we can describe a number of observed properties simply by specifying that the efficiency of transfer of mass and angular momentum from the initial gaseous core to the final binary are both constant. This model therefore provides an excellent framework with which to examine the importance of turbulence in the context of isolated star formation, and to explore the role which the stochastic nature of the initial turbulent state has on the statistical properties of binary systems. While it is certainly true that a number of other -- 4 -- physical effects other than turbulence and angular momentum dissipation may play a role in the star formation process, our model can elucidate their relative importance by serving as a strawman against which other results can be compared. The paper begins with an overview of observational evidence (§2), including both pre-stellar molecular cloud core (§2.1) and binary (§2.2) properties. Next, we describe our pre-stellar core model in §3. In §4, we present the methodology which connects the pre-stellar core model with the binary properties. The results of this model are detailed in §5. We conclude the main body of the paper with a discussion of the implications of this semi-empirical model for observational and theoretical studies of star formation (§6). Lastly, appendix A includes an elementary derivation of characteristic binary scales, in terms of their orbital parameters. 2. Observational Evidence 2.1. Pre-Stellar Core Properties 2.1.1. Density Structure The observations of Motte & Andr´e (2001) consisted of a complete 1.3 mm survey of both pre-stellar cores and protostellar envelopes using the IRAM 30m telescope and the MPIfR bolometer array (MAMBO). Their observations resolved structures from 1,500 AU to 15,000 AU. Motte & Andr´e (2001) concluded that their observations of pre-stellar core backgrounds were flatter than predicted by a singular isothermal sphere model (Shu 1977), but were consistent with Bonnor-Ebert spheres (Bonnor 1957; Ebert 1955). -- 5 -- 2.1.2. Velocity Structure on the Molecular Cloud Core Scale In 1981, in an extension of an earlier paper (Larson 1979) focusing on the large-scale structure of the interstellar medium, Larson (1981) published a seminal paper on the internal velocity dispersion in molecular clouds. Using data already published in the literature in a wide variety of studies, he accumulated about 50 data points for a number of different star forming regions, on scales ranging from 10−1 to 102 pc. He established scaling laws for both the mean density and the total internal velocity dispersion over projected lengthscales on the sky. Plotting the total internal velocity dispersion (sum of thermal and non-thermal linewidths) versus the maximum projected linear size of the region, he found that the internal velocity dispersion obeyed a power law of the form σ1D(km s−1) = 1.10 L(pc)0.38 (1) Here we have designated the one-dimensional linewidth dispersion, as inferred from velocities along the line of sight by σ1D, to properly distinguish it from the fully 3D velocity dispersion. Significantly, the power law nature of the turbulent linewidth-size scaling laws supports the premise, which Larson originally suggested, that over a wide range of scales, the observed interstellar turbulence is part of a scale-free hierarchy of turbulent eddies. This scale-free cascade should extend all the way down to about the characteristic scale where such eddies are damped -- either the minimum Jeans mass in the cloud (Larson 1995) or the ion-neutral damping length (Myers & Lazarian 1998), depending on whether thermal or magnetic damping effects dominate, which in turn depends on the relative strength of the magnetic field and the ionization state of the gas. Subsequent authors investigated the linewidth-size scaling relationship in more detail for molecular cloud cores using more precise techniques. In particular, Larson's original -- 6 -- study contained only 4 cores observed with rather poor angular resolution, and in the absence of actual data, made blanket assumptions regarding the kinetic temperature of the gas used in computing the local soundspeed. Leung, Kutner, & Mead (1982) and Myers (1983) re-investigated Larson's linewidth- size relationship and found the exponent closer to 1/2, rather than 1/3. In the most extensive study of molecular cloud cores to date, Jijina, Myers, & Adams (1999) studied 264 cores mapped in NH3 in a wide variety of regions and environmental conditions. They conclude that the linewidth-size relationship for all cores has an exponent of .63 ± .10, though a slight difference of low statistical significance existed between subsamples with embedded IRAS sources (exponent .49 ± .12) and without (exponent .83± .18). Moreover, Jijina, Myers, & Adams (1999) found a significant variation in the median value of the non-thermal linewidth ∆vN T across regions : from a value of 0.22 km/s in Taurus (thermal linewidth median ∆vT = .44 km/s), to a non-thermal linewidth ∆vN T median value of .86 km/s in Orion A (thermal linewidth median ∆vT = .58 km/s). Over all regions, cores without IRAS embedded sources or clusters had ∆vN T /∆vT = .65+.35 −.25 (within one quartile of the median). Cores without clusters, but with IRAS embedded sources had ∆vN T /∆vT = 1.0+.40 −.30. Cores with clusters had higher non-thermal linewidths still. We note that the most complete observations to date suggest that even in low-mass cores, the level of turbulent support is comparable to that of thermal support, although there are large variations both within and across individual star-forming regions.1 Moreover, there is little evidence that the non-thermal linewidths diminish during the process of star formation; if anything, the non-thermal linewidths in cores with IRAS embedded sources is higher than in their non-IRAS counterparts. However, we note that the non-thermal 1Although these cores are often termed "quiescent", such a designation is somewhat misleading. -- 7 -- linewidths measured in cores with embedded objects will in general also include infall motion towards one or more protostars in a complex flow field. Such infall motions are extremely difficult, if not impossible, to separate from turbulent motions on the scale of the core without a full spatial mapping of the velocity field, and so it is remains unclear (from an observational standpoint) what the timescale for the dissipation of turbulence is, or indeed, even whether turbulence decays or increases during star formation. 2.2. Binary Star Properties 2.2.1. Binarity Fraction Prompted by inconsistencies in multiplicity fractions in earlier surveys (Kuiper 1935, 1942; Jaschek & Jaschek 1957; Petrie 1960; Jaschek & G´omez 1970), Abt & Levy (1976) determined the multiplicity fraction of a sample of 135 bright field stars of types F3-G2, and found a binarity fraction (the fraction of all primaries having a detectable secondary companion) of 57%. Later, Duquennoy & Mayor (1991) (DM91 hereafter) using a complete sample of 164 primary G dwarf stars with types F7-G9, found a remarkably similar binarity fraction of 58% (although their conclusions regarding the period distribution differed substantially; see below). 2.2.2. Period Distribution DM91 demonstrated that the binary periods in a complete sample of field G dwarf stars fits a broad log-normal distribution remarkably well, with a median period of log ¯Pd = 4.8 and a standard deviation of log σPd = 2.3, where Pd is the period measured in days. DM91's result is substantially different than the median period log ¯Pd = 3.7 found previously by Abt & Levy (1976), whose survey was biased by their use of a magnitude-limited sample. -- 8 -- Similar period distributions were found for field K dwarfs (Mayor, Duquennoy, Halbwachs, & Mermilliod 1992) and field M dwarfs (Fischer & Marcy 1992), though the K dwarf sample was found to have a somewhat narrower distribution (log σPd = 1.9). Mathieu (1994) compiled data of pre-main sequence (PMS) binaries extant at that time. While the statistics of PMS binaries are not as complete as those in the field, the compiled PMS binary period distribution appears to fit a log-normal distribution similar to that of Duquennoy & Mayor (1991). More recently, other observers have begun to produce tentative evidence in favor of the hypothesis that the period distribution itself varies from region to region. For instance, in observations of Upper Scorpius A and B, Brandner & Koehler (1998) argue that the period distribution median varies significantly between the two regions, with wider binaries being preferred in the low-mass star forming region. Similarly, Scally, Clarke, & McCaughrean (1999) argue that a deficiency of wide (a > 1000 AU) binaries exists in Orion. However, such studies of individual regions and distribution tails reduce the counting statistics by necessity (some 20 systems in the case of Brandner and Koehler's study; some 3 systems in the case of Scally et al's study), and so we must consider them with some caution. To date, the period distribution has not been quantitatively explained by theory. Mouschovias (1977) envisioned dense cores as non-turbulent, highly subcritical, slowly contracting regions initially endowed with angular velocity induced by galactic shear. In his picture, magnetic braking becomes inefficient at some critical density at which the core becomes supercritical, and thereafter the angular momentum of the gas is conserved during the collapse. He invoked variations in this critical density parameter across all star-forming regions to explain the broad variation in the binary period distribution. However, it remains unclear whether this picture is quantitatively consistent with the observed period distribution. -- 9 -- More recently, Larson (2002) offered the possibility that tidal interactions among binaries in a stellar cluster would tend to transfer angular momentum among the binaries, and thereby broaden the period distribution. However, Kroupa & Burkert (2001) demonstrated that even under extreme conditions (stellar densities ∼ 106 pc−3), gravitational encounters were not able to explain the observed breadth of the binary period distribution. Indeed, even when a very small amount of broadening of the period distribution was obtained in their densest configurations, disruption dominated over broadening, so that the binarity fraction dropped from an initial value of unity to ∼ .2 in about 3 · 104 yr, in sharp conflict with observation. Indeed, Kroupa & Burkert (2001) concluded that the binary period distribution was a remnant of the initial conditions of star formation, and was not due to subsequent evolution. However, Bate, Bonnell, & Bromm (2002) found that close systems could form in a gas-rich environment through dynamical friction. A corollary of Bate's findings is that in a turbulent environment, the angular momentum dissipation rate should be sensitive to initial conditions -- as had been suggested earlier by Larson (2001). We will return to this point later (§6). 2.2.3. Eccentricity Distribution Another important diagnostic of binary stars is the eccentricity distribution. DM91 divided their binary catalog into short (P < 103 d) and long (P > 103 d) samples. They found that while the short-period binaries had a distribution which peaked at e = .3, long-period binaries followed a distribution which was roughly linear (f (e) = 2e), up to eccentricities of about .8, at which point the eccentricity distribution tapers off significantly. Mathieu (1994) reviewed the data extant at that time for MS and PMS binaries, and compared plots of the eccentricity e versus period P for each population of binaries separately. One of his principal findings was that the eccentricities of longer period systems -- 10 -- (both MS and PMS) were positively correlated with the period of the system, to periods of at least 103 days. Moreover, the MS eccentricity distribution appears to be roughly consistent with the PMS eccentricity distribution. This suggests that the eccentricities of MS binaries are established by age of the PMS stage (about 106 years) or earlier. 2.3. Mass Ratio An accurate observational determination of the the mass ratio distribution requires a number of corrections due to observational biases. DM91 found a mass ratio distribution which was consistent with drawing both individual members of the binary system independently from the IMF, and could be fitted to a Gaussian. In particular, their mass ratio distribution exhibited no maximum towards q = 1. However, a later paper (Mazeh, Goldberg, Duquennoy, & Mayor 1992), pointed out that once additional selection-effect corrections for spectroscopic binaries with periods less than 3 · 103 d are taken into account, the resultant period distribution for short period binaries is nearly flat, with a slight rise towards q = 1. 3. Initial Molecular Cloud Core Model Formulation 3.1. Overview Our model for turbulent molecular cloud cores will consist of two essential elements. The first, the background density model (§3.2) is based upon a Bonnor-Ebert density profile of isothermal gas, with pressure support enhanced through turbulent pressure. As we will show, this density profile is consistent both with Larson's mean density relationship and with resolved observations of molecular cloud cores (§2.1.1). The second element, a turbulent velocity field (§3.3), is superposed upon this background density. This velocity -- 11 -- field is consistent with Larson's linewidth-size relation (§2.1.2). In contrast to our assumptions of isothermal gasdynamics and a turbulent velocity field, other authors have have assumed that turbulence can effectively be accounted for by using a logotropic (McLaughlin & Pudritz 1997) or polytropic pressure component (McKee & Holliman 1999), and derived density profiles for their models under the assumption that gravity was balanced by the thermal and turbulent pressure of the core. Our approach differs fundamentally in that it is essentially a non-equilibrium model -- while the kinetic, thermal, and gravitational energies are in virial balance, our cores are never static. This difference is absolutely critical in the context of the current work, since we seek to tie the binary properties to the angular momentum of the turbulent model cores. 3.2. Background Density Model Since our models include turbulent support, we define an effective soundspeed ceff, which includes contributions from both thermal and turbulent pressure, added in quadrature. Because pressure depends on only the 1D rms of the velocity, we include the one-dimensional turbulent support through the effective soundspeed : ceff = cisos1 +(cid:18) M√3(cid:19)2 (2) It is both convenient and physically meaningful to scale the radius r to a dimensionless unit ξ = r/r0, where r0 proportional to the Jeans length at the central density ρc and to scale the density itself to a dimensionless unit Θ in terms of the central density : r0 = ceff√4πGρc = ξ = r/r0, ceff p4πGχρedge (3) (4) -- 12 -- and Θ = ρ/ρc, (5) where ρedge is the edge density of the core, and χ = ρc/ρedge is the density contrast between the center of the core and its edge. Combining the Poisson equation with the momentum equation, assuming the pressure support is derived from a polytropic equation of state, and demanding both spherical symmetry and force balance in the radial direction, gives us the Lane-Emden equation : 1 ξ2 d dξ (cid:18)ξ2 dΘ dξ(cid:19) + Θn = 0 (6) where γ = 1 + 1/n, and γ is the ratio of specific heats. We take the boundary conditions at ξ = 0 as Θ(ξ = 0) = 1 and Θ′(ξ = 0) = 0, appropriate for centrally condensed initial conditions. While there is no closed-form solution to (6) for γ = 1, we utilize a simple analytic approximation originally suggested by Natarajan & Lynden-Bell (1997): Θ(ξ) =(cid:18) A C 2 + ξ2 − B D2 + ξ2(cid:19) (7) where the parameters A, B, C, and D are numerical constants parameterizing the solution. Natarajan & Lynden-Bell (1997) suggest A = 50, B = 48, C = √10, D = √12. We assume the density distribution is critically stable, with a contrast χ ≃ 14 (Bonnor 1957; Ebert 1955). The resultant approximation accurately describes the critical Bonnor-Ebert sphere to within 1% at all spatial points. In the absence of turbulence, a critically stable core will have the familiar Bonnor-Ebert mass and radius (Bonnor 1957; Ebert 1955) : Mth = 1.18 Rth = .485 c3 iso pG3ρedge ciso pGρedge (8) (9) -- 13 -- Where ρedge is the edge density of the core. Using the effective soundspeed in place of the isothermal soundspeed, we have expressions for the turbulently-supported Bonnor-Ebert mass and radius: MBE = 1.18 RBE = .485 c3 eff pG3ρedge ceff pGρedge (10) (11) Larson's mean-density law states that the mean density within a given volume scales as the inverse size of that region (Larson 1981). Hence the column density through a given region should remain roughly independent of the size of the region. Applying this mean-density scaling to our MBE and RBE, we find MBE R2 BE ∝ ceffr ρedge G = const. (12) Hence, because the effective soundspeed ceff must scale with the turbulent linewidth M, we find that the edge density of the cloud core cannot remain a constant, but must scale as ρedge = (13) ρ0 p1 + (M/3)2 where ρ0 is the central density, or equivalently, the edge density as one approaches M → 0. In the remainder of the paper, we will set ρ0 = 5 · 10−20 gm cm−3, and ciso = 2 · 104 cm s−1, which imply RBE = .07 pc and MBE = 2M⊙ for a transonic (M = 1) molecular cloud core, which is typical for cores in the Taurus region (Jijina, Myers, & Adams 1999). We note that in a large-scale, isothermal, supersonic molecular cloud, shocks will compress and rarefy the mean flow of the gas, producing a log-normal distribution in the density field, where the width of the distribution depends on the Mach number of the region (Padoan & Nordlund 2002). In the current work, we treat only the median density, and not the tails -- 14 -- of the density distribution. As a result, the minimum mass binary system treated will be proportional to the thermal Jeans mass (see §4). This approximation is justified for systems whose masses are each close to the median stellar mass, but above the substellar limit. We will extend the semi-empirical method to systems containing at least one substellar component in a future paper (Fisher 2003). 3.3. Velocity Field Model As we noted earlier (§2.1.2), Larson's linewidth-size relation, as determined by recent authors (Jijina, Myers, & Adams 1999) states that the velocity dispersion ∆v along a line of sight through a region of size R scales roughly as R1/2 on molecular cloud core scales and above. We assume that each Fourier mode in the velocity decomposition is uncorrelated. By the Central Limit Theorem, uncorrelated Fourier modes are equivalent to demanding that the probability distribution of the amplitude of each mode is drawn from a Gaussian distribution, and the phase of each mode is drawn from a uniform distribution. Decomposing an individual mode δk in Fourier space in terms of its amplitude and phase as δk = rk exp(iφk), the Gaussian probability distribution gk for the amplitude rk and phase φk can be written as gk(rk, φk)drkdφk = 2(rkdrk) σk 2π (cid:19) exp(cid:18)−r2 (cid:18)dφk k (cid:19) k σ2 (14) The standard deviation σk is referred to as the power spectrum of the perturbation spectrum. As is conventional, we write P (k) = σk. By examining the successive moments of a Gaussian distribution, one can show that only the first two moments -- the mean and standard deviation -- are needed to entirely specify the distribution. Hence, specifying the mean and standard deviation completely determines the Gaussian spectrum. In the -- 15 -- following discussion, we assume that the mean of the perturbing field is zero, and that the perturbing field is homogeneous and isotropic, so that the amplitude of the power spectrum depends solely on the magnitude of the k−vector. By Parseval's Theorem, the total power introduced in k-space is identical to the total power in real space. Hence, the 3D rms velocity ∆v in a 3D spherical volume in real space extending from Lmin to Lmax in the radial direction can be easily evaluated by integrating the power spectrum in k-space from kmin = 2π/Lmax to kmax = 2π/Lmin: (∆v)2 =Z 4π 0 dΩZ kmax kmin k2dkP (k) (15) Adopting a power-law power spectrum of the form P (k) = Ak−n, then integrating over angles and wavenumbers, we have (∆v)2 = 4πA (3 − n)(cid:0)kmin−n+3 − kmax−n+3(cid:1) Assuming that most of the power is on large scales (n > 3) and that Lmin ≪ Lmax, the maximum wavenumber kmax corresponding to the minimum scale in the problem can be neglected, so we have (16) (17) ∆v =(cid:18) 4πA n − 3(cid:19)1/2 max (cid:0)2πL(−n+3)/2 (cid:1) When n = 4, we can recover Larson's Law (∆v ∝ L1/2 perturbation power spectrum provides the basis for our turbulent models. Specifically, we max). This n = 4 Gaussian velocity assume all background velocities are zero, and apply three separate turbulent velocity fields to our model molecular cloud core via vx(~r) = vy(~r) = vz(~r) = δx(~r)cisoM√3 δy(~r)cisoM√3 δz(~r)cisoM√3 (18) (19) (20) -- 16 -- where δx(~r), δy(~r), and δz(~r) are three realizations of a k−4 Gaussian field, ciso is the isothermal soundspeed, and M = ∆v/ciso is the 3D rms turbulent Mach number. While Gaussian perturbation fields and the structures which arise from them have been a topic of active investigation in the cosmological context for several decades, our use of them in the context of star formation differs in two critical respects. First, application of a perturbation to the density field requires that the total power must be small : a nonlinear perturbation with sufficient total power will drive the density negative in some regions. By applying perturbations to the velocity field directly, the total power can be made arbitrarily large without encountering unphysical negative densities. Second, whereas cosmological density fluctuations typically introduce most power on small scales (the inflationary Harrison-Zel'dovich spectrum uses n = 1), we introduce most power on large scales, in accord with observations of star forming regions. One should appropriately ask how the ratio of rotational to gravitational binding energy, β, depends on the 3D turbulent Mach number M, while keeping the realization of the turbulence fixed. Interestingly enough, under the assumption of a critical Bonnor-Ebert density distribution, for transonic and supersonic cores, a fixed realization yields a β that is completely independent of the edge-density of the core, and only weakly dependent on the turbulent Mach number. Specifically, β ∝(cid:18) J 2 I (cid:19)(cid:18) R M 2(cid:19) β ∝(cid:18)M2R M (cid:19) β ∝ M2 1 +(cid:16) M√3(cid:17)2 (21) (22) (23) ie, for transonic and supersonic cores with M ≥ 1, β varies by roughly a factor of 3 from M = 1 to M = ∞. Here I is the moment of inertia of the initial core. This scaling -- 17 -- explains why turbulent core models naturally produce the same median value of β as is seen in observation without any fine-tuning of parameters: in critical Bonnor-Ebert spheres dominated by turbulent pressure support, Mach scaling applies, and as a result the models are scale-free. Therefore, in the supersonic regime, for a fixed turbulent realization, β does not depend on any model parameters. This scaling holds approximately even down to the transonic regime, so that the value of β is primarily determined from the slope of the turbulent spectrum. The robust agreement between the predicted value of β and observation gives strong support for our model of the turbulent velocity field. Similar Gaussian turbulent spectra imposed on the velocity field have been used in a variety of numerical simulations of turbulence in the interstellar medium. Dubinski, Narayan, & Phillips (1995) were apparently the first to suggest that Larsonian turbulence in the interstellar medium could be generated using Gaussian random velocity fields with index n = 4. A number of authors, beginning with Gammie & Ostriker (1996), studied the evolution of turbulence in molecular clouds using the same turbulent spectrum -- initially establishing an Alfv´enic spectrum of waves by an initial velocity fluctuation on a constant density and magnetic field background. Burkert & Bodenheimer (2000) computed linewidth gradients induced by turbulent eddies on the scale of molecular cloud cores. Hujeirat, Myers, Camenzind, & Burkert (2000) computed the initial decay of turbulence and subsequent collapse (followed to an early evolutionary time) of molecular cloud cores supported by Alfv´enic turbulence. Lastly, Bate, Bonnell, & Bromm (2002) computed the evolution of a large, highly supersonic (M ∼ 7) turbulent core. 4. Semi-Empirical Model for Binary Formation via Turbulent Fragmention We may now predict the distribution of binary periods implied by our isolated turbulent core models. Since our knowledge of the detailed physics of turbulent fragmentation is still -- 18 -- rudimentary at best, in order to predict binary properties from our initial molecular cloud core models, we must make several semi-empirical assumptions motivated by observation. We will assume that the star-formation efficiency ǫ∗ = (M1 + M2)/Mcore is constant. Further, in direct analogy to the star-formation efficiency, we define the conversion efficiency of angular momentum of cores to binaries, ǫJ = J/Jcore, and assume it is constant as well. In addition, we will assume that the individual stellar masses are uncorrelated, so that both may be randomly drawn from the IMF. Lastly, we will assume that the orbital eccentricities of all binaries are drawn from a distribution f (e) = 2e. We recognize that these assumptions may not be rigorously correct; in particular, based on observations of field stars, there is evidence (Duquennoy & Mayor 1991) that short (P < 103 d) period binary systems follow a different eccentricity distribution than long period binaries, and may also have a different mass-ratio distribution (Mazeh, Goldberg, Duquennoy, & Mayor 1992). However, because such short-period systems will be strongly affected by disk-star interactions, it is likely that their orbital parameters undergo significant evolution. Hence it, in the context of obtaining initial binary properties in the current work, we will explore the consequences of the hypothesis that all binary systems are formed with uncorrelated masses drawn from the IMF, and obeying a thermal eccentricity distribution. We proceed as follows. 1) First, we randomly draw two masses M1 and M2 from the IMF. We uniformly draw ξ over the interval from 0 - 1 twice, and assign masses (scaled to solar) and mass ratio q according to m(ξ) = .08 + γ1ξγ2 + γ3ξγ4 (1 − ξ).58 , (24) where γ1 = .19, γ2 = 1.55, γ3 = .050, γ4 = .6 (Kroupa, Gilmore, & Tout 1991). We note that this IMF introduces a cutoff at .08M⊙, and so does not extend into the substellar range. We define M1 to be the greater of these two masses; hence q = M2/M1. -- 19 -- 2) Next, using the star formation efficiency factor, set the mass of the initial protostellar core: Mcore = (M1 + M2) ǫ∗ . (25) Since we now know the parent core mass as well as the thermal Bonnor-Ebert mass, we may also set the 3D turbulent Mach number, using equation (3.2) : M = 3"(cid:18)Mcore Mth (cid:19)4/7 − 1#1/2 (26) In cases where Mcore < Mth, the computed core mass is less than the thermal Bonnor-Ebert value, implying that the initial core is inconsistent with the selected binary masses and assumed star formation efficiency. In these instances, we simply reject the model binary and draw both stars again. 3) We then stochastically generate three perturbation cubes, and perturb the velocity field of our model core using the known Mach number M, thereby setting the resultant core angular momentum Jcore as well as the binary system angular momentum J = ǫJ Jcore. 4) Next, we draw the eccentricity of the binary orbit from the thermal distribution f (e) = 2e. We do this by drawing a number ξ from a uniform distribution ranging from 0 to 1, and setting e = √ξ. 5) Finally, knowing J, e, M and q, we may compute the period P and semi-major axis a of the binary system, from and P =(cid:18) 2π G2(cid:19)(cid:18) J 3 M 5(cid:19) 1 (1 + q)6 (1 − e2)3/2 q3 a = 1 G(cid:18) J M 2(cid:19) 1 (1 − e2) 1 M (1 + q)4 q2 (27) (28) (See Appendix A for derivation.) -- 20 -- These relations are often written down for the special case of an equal-mass, circular binary (e = 0, and q = 1). Clearly, in that case, the larger the value of J/M, the greater the amount of angular momentum in the system, and hence, the longer the period. In addition, the smaller the total mass M of the system, the less the influence of gravity, and hence, the wider the binary. However, although it is commonly not recognized, for fixed mass and angular momentum, the eccentricity and the mass ratio can also have a substantial impact on the orbital properties of binaries. For instance, for a fixed angular momentum, the larger the eccentricity e of a system, the wider the binary. Similarly, for a fixed mass, the smaller the mass ratio q, the wider the binary. Although apparently trivial, these scaling relationships will have an important bearing on the statistical properties of binary star systems, as we shall soon see. Our models introduce two free parameters: ǫ∗ and ǫJ , the mass and angular momentum star formation efficiencies, respectively. The first parameter is relatively well-constrained by both theory and observation (Matzner & McKee 2000). For a given ǫ∗, the ǫJ parameter is obtained by requiring that the resultant model binary period distribution median agree with the observed value. In effect, the imposition of this constraint reduces the parameter space to one free parameter, which we take to be ǫ∗. We note, from equation (27), that with the constraint that the model period median must agree with observation, that ǫJ ∝ ǫ5/3 . ∗ The nonlinearity of the dependence of ǫJ on ǫ∗ has a simple interpretation: for low star formation efficiency, a larger amount of mass must be accreted onto the binary, which implies a greater angular momentum loss efficiency due to the larger turbulent linewidth on larger scales. Note that our model implicitly assumes that the mass of the system M, the mass ratio q, and the orbital eccentricity e are uncorrelated, whereas the angular momentum J (and hence the period P ) is explicitly dependent upon M : the higher the mass of -- 21 -- the binary system, the greater the mass and Mach number M of the initial core. These assumptions differ from those adopted by Kroupa (1995) in his formulation of the inverse binary population synthesis problem, where he assumed that P , q, and e were uncorrelated. In contrast, in our formulation, the period P depends explicitly on q and e through angular momentum conservation (equation 27), which implies that our model predicts non-trivial correlations between P and q, and between P and e. We return to these points in later sections (§5.2 and §5.3), where we detail the nature of these correlations. 5. Results 5.1. Binary Period Distribution Figure 1 shows the binary period distribution obtained for the star formation efficiency ǫ∗ = 0.26. The standard deviation of the computed distribution is σPd = 1.7; slightly narrower than that of the field, though consistent with the PMS distribution. The computed distribution is in remarkably good agreement with Mathieu's cataloged visual binaries with periods > 104 d. The computed distribution exhibits a somewhat smaller binarity fraction at periods shorter than 103 d than either the field or the PMS distribution, although given the numbers involved, this result is of marginal statistical significance. There is a small, but systematic increase in the width of the distribution with increasing star formation efficiency. We interpret this increasing width as a consequence of the decrease in angular momentum loss efficiency with increasing star formation efficiency. As a result, higher star formation efficiency models have slightly wider distributions of pre-stellar core specific angular momenta. The distribution of initial core specific angular angular momenta for the case of ǫ∗ = 0.26 is shown in figure 2. -- 22 -- 5.2. Eccentricity Versus Period As we noted in §4, in our semi-empirical model of turbulent star formation, the angular momentum of the binary system is directly related to the initial angular momentum of the turbulent core. As a result, the period and eccentricity are no longer uncorrelated quantities, and the period is explicitly dependent upon the eccentricity; for a fixed angular momentum and mass, the greater the eccentricity, the longer the period. From observation (§2.2.3), we know that in binaries wider than the tidal circularization limit (P ∼ 10 d), the period tends to be correlated with the eccentricity, which is in fact a robust feature of our model (see equation 27 ). We plot eccentricity versus the log (base 10) of the period for in figure 5, for the case of intermediate star-formation efficiency (ǫ∗ = 0.5). For comparison, binaries from the field (Duquennoy & Mayor 1991), and from PMS regions (Mathieu 1994), are shown in star and plus symbols, respectively, in the plot on the right. The vertical dashed line indicates the tidal circularization limit found by Duquennoy & Mayor (1991) at about Pd ∼ 11 d. The key result here is that the model systems exhibit a positive correlation between eccentricity and period, which is qualitatively similar to that observed. However, it is certainly true that our models predict an overabundance of highly eccentric binaries at shorter periods. It is quite likely that additional angular momentum dissipation mechanisms, including coupling to circumstellar disks, may help circularize such systems. 5.3. Mass Ratio Versus Period The semi-empirical model also predicts a correlation between mass ratio and period. The reason for the correlation is, once again, simply related to conservation of angular momentum. A system with two unequal masses, and a small mass ratio (q ≪ 1) will have -- 23 -- a longer period than a similar system with the same total angular momentum and mass, but with equal masses. The results of 200 model systems for the case of intermediate star-formation efficiency (ǫ∗ = 0.5) are shown in figure 4. We emphasize that in order to explain the correlation between mass ratio and period, the semi-empirical model framework does not require that the masses themselves be correlated. Instead, the model suggests that it is possible to form two stars independently within a single turbulent core, with the resultant period anti-correlated with the mass ratio simply through angular momentum conservation. 5.4. Binary System Mass Versus Semimajor Axis As we noted previously, observations of low-mass systems have revealed that such systems do not follow the same statistical trends observed in solar-mass and higher binaries. Perhaps the most noticeable difference is the trend for low-mass systems to occur in tighter binaries. In the semi-empirical model, two factors contribute to the explanation of this effect. First, lower-mass binary systems originate from parent cores of lower mass, and hence, lower specific angular momentum. Second, drawing stars independently from the IMF, a low-mass system will naturally tend to have a more equal-mass ratio. Taken together, the model predicts that low-mass systems should naturally tend to occur in shorter-period systems, which is in fact what is observed (Mart´ın & Basri 2001). In figures 6 - 7, we display the results from two model systems of variable star formation efficiency. The minimum system mass represented for a given star-formation efficiency ǫ∗ is indicated with a dashed horizontal line; no model systems form below the line. For comparison, we also plot the largest observable binary separations amax, as determined by Close, Siegler, Freed, Biller (2003) in their figure 15. They found that the two straight lines -- 24 -- of constant orbital velocity amax = 23.2(Mtot/.185) AU and amax = 1000(Mtot/.185) AU (where Mtot is in solar masses) describe the maximal separation upper envelopes relatively well. We note that, above the horizontal minimum mass line, our model systems fit the Close et al. upper envelopes remarkably well; very few systems lie to the right of the envelopes, quite independent of the star-formation efficiency. In the context of the current paper, our assumption of a fixed Jeans mass requires that low mass systems must form in regions of low star-formation efficiency. Hence, it appears that a combination of a variety of star-formation efficiencies can explain the results of Close et al. More general models including constant star-formation efficiency and supersonic turbulence on large scales may also be able to explain the observations; a topic we are investigating further (Fisher 2003). 6. Discussion One limitation of our current approach is the choice of a fixed central density ρ0. By specifying ρ0, we have implicitly restricted our attention to systems above a mass ǫ∗Mth. As we pointed out previously (§3.2), shock compressions in a turbulent molecular cloud will produce a log-normal distribution of ρ0. We expect that a set of binary systems formed from cores with varying Mth resulting from to shock compression will in fact have a slightly broader period distribution than that computed here. Similarly, the IMF adopted introduces a cutoff at .08M⊙, even though substellar objects are also formed in star-forming regions. These effects may partially account for the slightly smaller period distribution standard deviations computed here (σPd ≃ 1.6 − 2.1), in comparison to that found in the field (σPd ≃ 2.3) (Duquennoy & Mayor 1991). Detailed theoretical work treating wind-driven outflows (Matzner & McKee 2000) -- 25 -- predicts that the star-formation efficiency on the molecular cloud scale should lie in the range 25 - 70%. For star-formation efficiencies in this range, our models predict core properties consistent with observation. In particular, with our fiducial scaling, the median Mach number ¯M lies in the range of 1.2 and 0.9 over the range ǫ∗ = 0.1 − 0.9 -- quite close to the actual median value of .9 observed in Taurus (Jijina, Myers, & Adams 1999). We note that since both our pre-stellar core and final binary properties are relatively insensitive to the star-formation efficiency (the median prestellar core Mach number ¯M varies only by a factor of 2 for star-formation efficiencies over the range ǫ∗ = 0.1 − 0.9) our models do not provide further constraints on the star-formation efficiency itself. A key consequence of the semi-empirical model for star formation described here stems from the fact that many observational properties can be quantitatively described with a constant angular momentum loss factor. In contrast, several authors have advocated a "chaotic" description star formation (Larson 2001; Bate, Bonnell, & Bromm 2002), in which a variety of cluster interactions play a key role in setting the initial properties of binaries. In this paper, we have explicitly demonstrated that chaotic interactions are not required to explain many properties of binaries wider than P > 103 d, provided that turbulent fragmentation can in fact directly produce the mass and eccentricity distributions which we have assumed. Note that we distinguish the fact that our initial conditions are turbulent, and therefore inherently stochastic (though deterministic) from the sensitivity upon initial conditions implied by a chaotic model. We would, however, agree that dynamical friction is likely to play an important role in the formation of tight binaries, and that subsequent tidal interactions may also act to broaden the initial period distribution found here. Another key consequence of our model is the importance which angular momentum loss from initial molecular cloud cores plays in determining the properties of binaries. Although this problem has been identified by a number of authors (Simon 1992; Bodenheimer 1995; -- 26 -- Larson 2002), virtually all simulations of multiple star formation on the molecular cloud core scale done to date remain purely hydrodynamic, and have neglected the influence of the magnetic field entirely.2 As a result, the only means of angular momentum transport treated in most calculations is gravitational torquing. Given the importance of angular momentum loss in determining binary properties, and the relative effectiveness of magnetic braking in axisymmetric simulations of single-star formation (Basu & Mouschovias 1994), caution must be applied in comparing the results of purely hydrodynamic simulations directly against observation; in all likelihood, the magnetic field plays a critical role in determining the binary angular momentum. RTF would like to thank Richard Klein and Christopher McKee for their prescient suggestion to investigate the physical role of turbulence in multiple star formation as a thesis topic. Thanks also go to Marc Davis and Matt Craig for the use of their Gaussian perturbation generation program. This research has made use of NASA's Astrophysics Data System Bibliographic Service. A. Characteristic Scales In this appendix, we derive the characteristic scales of a binary system consisting of two masses M1 and M2, in terms of its angular momentum J, its mass ratio q = M2/M1, and its orbital eccentricity e. From standard two-body classical mechanics in an inverse square law potential 2To the best of our knowledge, the only exceptions are the work of Hujeirat, Myers, Camenzind, & Burkert (2000) and Boss (2000), although Boss did not include a fully self- consistent treatment of the MHD equations. -- 27 -- (Goldstein 1980), we know the angular momentum of a binary system is completely specified by the semi-major and semi-minor axes of the bound orbit, as well as the reduced mass of the two-body system µ = M1M2/(M1 + M2), and the inverse square law constant (A1) of proportionality k : J = b a1/2pµk, where µ = M1M2/(M1 + M2) is the reduced mass of the two body system, and k = GM1M2 for the gravitational force. Hence, since b = a√1 − e2, we can determine the final orbital separation as a function of the eccentricity e, total mass M = M1 + M2, and specific angular momentum J/M : J/M = a1/2√1 − e2s G(M1M2)2 (M1 + M2)3 (A2) Defining the ratio of secondary to primary mass q = M2/M1, for e ≤ 1, the semi-major axis of a bound orbit is a = 1 M(cid:19)2 G(cid:18) J 1 1 − e2 1 M (1 + q)4 q2 (A3) For a typical molecular cloud core specific angular momentum value of J/M = 3· 1020 cm2/s, a = 50AU(cid:18) J/M 3 · 1020 cm2/s(cid:19)2 1 1 − e2 (cid:18)M⊙ M (cid:19) (1 + q)4 q2 (A4) Then, from Kepler's third law, we can determine the period of the final binary system : P = 2π G2 1 (1 − e2)3/2 (cid:18) J M(cid:19)3 (1 + q)6 q3 1 M 2 P = 300 years 1 (1 − e2)3/2 (cid:18) J/M 3 · 1020 cm2/s(cid:19)3 (1 + q)6 q3 (A5) (A6) M (cid:19)2 (cid:18)M⊙ We anticipate that explicit dependence of P on e in equation (A6) may cause some readers, who are accustomed to the fact that the Keplerian period depends only on a, and not on e, some undue consternation. These concerns will be rapidly alleviated when it is recognized that, for fixed masses and angular momentum, a more eccentric binary system must have a wider separation (eqn. A4). Our equation (A6) expresses the period in terms of the -- 28 -- angular momentum J in favor of the semimajor axis a, and hence the origin of the explicit dependence of P on e. Lastly, using the classical two-body inverse square law result for the total energy E of the system E = − k 2a , one can also determine the final energy E of the binary system : E = 1 2 G2M 3(e2 − 1) q3 J (cid:19)2 (1 + q)6 (cid:18) M (A7) (A8) -- 29 -- REFERENCES Abt, H. A. & Levy, S. G. 1976, ApJS, 30, 273 Basu, S. & Mouschovias, T. C. 1994, ApJ, 432, 720 Bate, M. R., Bonnell, I. A., & Bromm, V. 2002, MNRAS, 332, L65 Bodenheimer, P. 1995, ARA&A, 33, 199 Bonnell, I. A., Bate, M. R., Clarke, C. J., & Pringle, J. E. 1997, MNRAS, 285, 201 Bonnor, W. B. 1957, MNRAS, 117, 104 Boss, A. P., Fisher, R. T., Klein, R. I., & McKee, C. F. 2000, ApJ, 528, 325 Boss, A. P. 2000, ApJ, 545, L61 Brandner, W. & Koehler, R. 1998, ApJ, 499, L79 Burkert, A. & Bodenheimer, P. 2000, ApJ, 543, 822. Clarke, C. J. & Pringle, J. E. 1991, MNRAS, 249, 584 Close, L. M, Siegler, N., Freed, M., & Biller, B., ApJ, to appear. Duquennoy, A. & Mayor, M. 1991, A&A, 248, 485 Dubinski, J., Narayan, R., & Phillips, T. G. 1995, ApJ, 448, 226 Ebert, R. 1955, Zeitschrift Astrophysics, 37, 217 Fischer, D. A. & Marcy, G. W. 1992, ApJ, 396, 178 Fisher, R. T. 2003, in preparation. Galli, D., Shu, F. H., Laughlin, G., & Lizano, S. 2001, ApJ, 551, 367 -- 30 -- Gammie, C. F. & Ostriker , E. C. 1996, ApJ, 466, 814 Ghez, A. M., White, R. J., & Simon, M. 1997, ApJ, 490, 353 Goldstein, H. 1980, Classical Mechanics, Addison-Wesley Goodman, A. A., Benson, P. J., Fuller, G. A., & Myers, P. C. 1993, ApJ, 406, 528 Hoyle, F. 1953, ApJ, 118, 513 Hujeirat, A., Myers, P., Camenzind, M., & Burkert, A. 2000, New Astronomy, 4, 601 Inutsuka, S. & Miyama, S. M. 1992, ApJ, 388, 392 Jaschek, C. & G´omez, A. E. 1970, PASP, 82, 809 Jaschek, C. & Jaschek, M. 1957, PASP, 69, 546 Jijina, J., Myers, P. C., & Adams, F. C. 1999, ApJS, 125, 161 Klein, R. I., Fisher, R., & McKee, C. F. 2001, IAU Symposium, 200, 361 Kroupa, P. 1995, MNRAS, 277, 1491 Kroupa, P., Gilmore, G., & Tout, C. A. 1991, MNRAS, 251, 293 Kroupa, P. & Burkert, A. 2001, ApJ, 555, 945 Kuiper, G. P. , PASP, 47, 15 Kuiper, G. P. 1942, ApJ, 95, 201 Larson, R. B. 1979, MNRAS, 186, 479 Larson, R. B. 1981, MNRAS, 194, 809 Larson, R. .B. 1995, MNRAS, 272, 213 -- 31 -- Larson, R. B. 2001, IAU Symposium, 200, 93 Larson, R. B. 2002, MNRAS, 332, 155 Laughlin, G. & Bodenheimer, P. 1994, ApJ, 436, 335 Leung, C. M., Kutner, M. L., & Mead, K. N. 1982, ApJ, 262, 583 Looney, L. W., Mundy, L. G., & Welch, W. J. 2000, ApJ, 529, 477 Mart´ın, E. L. & Basri, G. 2001, IAU Symposium, 200, 55 Mathieu, R. D. 1994, ARA&A, 32, 465 Matzner, C. D., & McKee, C.F. 2000, ApJ, 545, 364 Mayor, M., Duquennoy, A., Halbwachs, J.-L., & Mermilliod, J.-C. 1992, ASP Conf. Ser. 32: IAU Colloq. 135: Complementary Approaches to Double and Multiple Star Research, 73 Mazeh, T., Goldberg, D., Duquennoy, A., & Mayor, M. 1992, ApJ, 401, 265 McKee, C. F. & Holliman, J. H. 1999, ApJ, 522, 313 McLaughlin, D. E. & Pudritz, R. E. 1997, ApJ, 476, 75 Motte, F. & Andr´e, P. 2001, A&A, 365, 440. Mouschovias, T. C. 1977, ApJ, 211, 147 Myers, P. C. 1983, ApJ, 270, 105 Myers, P. C. & Lazarian, A. 1998, ApJ, 507, L157 Natarajan, P. & Lynden-Bell, D. 1997, MNRAS, 286, 268 -- 32 -- Padoan, P. & Nordlund, A. 2002, ApJ, 576, 870 Petrie, R. M. 1960, AJ, 65, 55 Rodriguez, L. F. et al. 1998, Nature, 395, 355. Scally, A., Clarke, C., & McCaughrean, M. J. 1999, MNRAS, 306, 253 Shu, F. H. 1977, ApJ, 214, 488 Simon, M. 1992, ASP Conf. Ser. 32: IAU Colloq. 135: Complementary Approaches to Double and Multiple Star Research, 41 Whitworth, A. P., Boffin, H. M. J., & Francis, N. 1998, MNRAS, 299, 554 This manuscript was prepared with the AAS LATEX macros v5.0. -- 33 -- ǫJ log ¯Pd log σPd ǫ∗ 0.1 .002 4.9 0.3 .01 0.5 .04 0.7 .06 0.9 .10 4.9 4.8 4.9 5.0 1.6 1.7 1.8 2.0 2.1 .31 .51 .72 .92 34 30 35 45 ǫ∗Mth/M⊙ ¯a/AU ¯j (cm2 s−1) .14 30 ¯M 1.7 1.2 1.0 .89 .87 1.3 · 1021 7.7 · 1020 4.5 · 1020 3.5 · 1020 3.4 · 1020 Table 1: A table of model parameters, for a variety of star formation efficiencies (ǫ∗ = 0.1 − 0.9) fitted such that the median model period agrees with that of the field. Shown are the mass and angular momentum star-formation efficiencies ǫ∗ and ǫJ , the median of the log of the period distribution (in days) log ¯Pd, the standard deviation of the log of the period distribution (in days) log σPd, the minimum mass binary system (in solar masses) ǫ∗Mth/M⊙, the median of the semi-major axis (in AU) ¯a, the median prestellar core angular momentum ¯j (in cm2 s−1), and the median prestellar core Mach number ¯M. -- 34 -- 0.4 0.35 0.3 0.25 n o i t c a r F 0.2 0.15 0.1 0.05 0 −1 0 1 2 3 4 log (Pd) 5 6 7 8 9 10 0.4 0.35 0.3 0.25 n o i t c a r F 0.2 0.15 0.1 0.05 0 −1 0 1 2 3 4 log (Pd) 5 6 7 8 9 10 Fig. 1. -- Histogram of log Pd. The left panel shows our numerical results for 200 model systems for the case of star-formation efficiency ǫ∗ = 0.26. For comparison, the right-hand panel shows the period distribution inferred from PMS stars (Mathieu 1994) and from field stars (Duquennoy & Mayor 1991), in solid and dashed lines, respectively. -- 35 -- 0.25 0.2 0.15 0.1 0.05 n o i t c a r F 0 19.5 20 20.5 21 21.5 log (Jcore (cm2 s−1) 22 22.5 23 Fig. 2. -- Histogram of the log of the model initial core specific angular mom entum j = J/M, measured in cm2 s−1, for 200 binary-producing cores, shown for a low star-formation efficiency (ǫ∗ = 0.26). -- 36 -- 0.25 0.2 0.15 0.1 0.05 n o i t c a r F 0 −2 −1 0 1 2 log (a / AU) 3 4 5 6 Fig. 3. -- Histogram of log a(AU). The numerical results for 200 model systems for the cases of low star formation efficiency (ǫ∗ = 0.26), are shown, with Poisson error bars drawn. -- 37 -- 1 0.9 0.8 0.7 0.6 1 M / 2 0.5 M = q 0.4 0.3 0.2 0.1 0 −1 0 1 2 3 4 log (Pd) 5 6 7 8 9 10 Fig. 4. -- Binary mass ratio versus log Pd. The circles show the results of 200 model systems for the cases of intermediate star-formation efficiency (ǫ∗ = 0.5). -- 38 -- 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 y t i c i r t n e c c E 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 y t i c i r t n e c c E 0 −1 0 1 2 3 4 log (Pd) 5 6 7 8 9 10 0 −1 0 1 2 3 4 5 6 log (Pd) Fig. 5. -- Eccentricity versus log Pd. The circles show the results of 200 model systems for the case of intermediate star-formation efficiency (ǫ∗ = 0.5). For comparison, binaries from the field (Duquennoy & Mayor 1991), and from PMS regions (Mathieu 1994), are shown in star and plus symbols, respectively, in the plot on the right. The vertical dashed line indicates the tidal circularization limit found by Duquennoy & Mayor (1991) at about Pd ∼ 11 d. -- 39 -- 101 ) 100 n u s M / t o t M ( g o l 10−1 100 101 102 103 104 105 log (a / AU) Fig. 6. -- Log semimajor axis versus log total mass. The circles show the results of model systems for the very low star-formation effficiency case ǫ∗ = 0.1. The two solid slanted lines indicate the upper envlopes indicated by Close, Siegler, Freed, Biller (2003) (see text). The horizontal dashed line is drawn to delineate the minimum mass binary system for this star-formation efficiency. -- 40 -- 101 ) 100 n u s M / t o t M ( g o l 10−1 100 101 102 103 104 105 log (a / AU) Fig. 7. -- Log semimajor axis versus log total mass, as in fig. 6. The circles show the results of model systems for the intermediate star formation effficiency case ǫ∗ = 0.5.
0811.0570
1
0811
2008-11-04T19:02:04
The Full Spectrum Galactic Terrarium: MHz to TeV Observations of Various Critters
[ "astro-ph" ]
Multi-wavelength studies at radio, infrared, optical, X-ray, and TeV wavelengths have discovered probable counterparts to many Galactic sources of GeV emission detected by EGRET. These include pulsar wind nebulae, high mass X-ray binaries, and mixed morphology supernova remnants. Here we provide an overview of the observational properties of Galactic sources which emit across 19 orders of magnitude in energy. We also present new observations of several sources.
astro-ph
astro-ph
The Full Spectrum Galactic Terrarium: MHz to TeV Observations of Various Critters Mallory S.E. Roberts∗, C. Brogan†, S. Ransom†, M. Lyutikov∗∗, E. de Oña Wilhelmi‡, A. Djannati-Ataï‡, R. Terrier‡, S.M. Dougherty§, E.D. Grundstrom¶, J.W.T. Hesselsk, S. Johnston††, M.V. McSwain‡‡, P.S. Ray§§, K.S. Wood§§, G.G. Pooley¶¶ and A. Weinstein∗∗∗ ∗Eureka Scientific, Inc., 2452 Delmer Street, Suite 100, Oakland, CA 94602-3017 †National Radio Astronomy Observatory, 520 Edgemont Road, Charlottesville, VA 22903 ∗∗Department of Physics, Purdue University, West Lafayette, IN 47907, USA ‡Astroparticule et Cosmologie (APC), CNRS, Universite Paris 7 Denis Diderot, 10, Paris, France §National Research Council of Canada, Herzberg Institute for Astrophysics, Dominion Radio Astrophysical Observatory, PO Box 248, Penticton, British Columbia V2A 6J9, Canada ¶Physics and Astronomy Department, Vanderbilt University, 6301 Stevenson Center, Nashville, TN 37235 kAstronomical Institute "Anton Pannekoek," University of Amsterdam, 1098 SJ Amsterdam, Netherlands ††Australia Telescope National Facility, CSIRO, PO Box 76, Epping, NSW 1710, Australia ‡‡Department of Physics, Lehigh University, Bethlehem, PA 18015 §§Space Science Division, Naval Research Laboratory, Washington, DC 20375-5352 ¶¶Cavendish Laboratory, Cambridge University ∗∗∗Department of Physics and Astronomy, University of California, Los Angeles, CA 90095 Abstract. Multi-wavelength studies at radio, infrared, optical, X-ray, and TeV wavelengths have discovered probable counterparts to many Galactic sources of GeV emission detected by EGRET. These include pulsar wind nebulae, high mass X-ray binaries, and mixed morphology supernova remnants. Here we provide an overview of the observational properties of Galactic sources which emit across 19 orders of magnitude in energy. We also present new observations of several sources. Keywords: pulsar, pulsar wind nebula, supernova remnant, EGRET, FERMI, GLAST, gamma-ray, VLA, XMM-Newton, Binary, ATCA, PSR J2021+3651, HESS, MILAGRO, MGRO J1908+06, RCW 49, LS 5039 PACS: 95.85.Pw, 95.85.Nv, 95.85.Bh, 97.60.Gb, 97.60.Jd, 97.80.Jp, 98.38.Dq, 98.38.Mz ALL ACROSS THE YOUNG PULSARS AND PULSAR WIND ELECTROMAGNETIC SPECTRUM NEBULAE For many years, the Crab nebula was the only Galac- tic source detected across the entire spectrum, from ra- dio to TeV energies [1]. The EGRET telescope on board the Compton Gamma-Ray Observatory detected dozens of Galactic sources at GeV energies [2], most of which were not firmly identified by the end of the mission. Since then, multi-wavelength studies of the EGRET er- ror boxes at radio optical and X-ray wavelengths have discovered probable counterparts for most of the sources bright above 1 GeV, most of them the products of su- pernovae (pulsars, pulsar wind nebulae (PWN), and su- pernova remnants (SNR), [3]). Furthermore, surveys of the Galactic plane by HESS, MILAGRO, and VERITAS (among others) have discovered TeV emitting regions which can plausibly be associated with many of these sources. Here we provide an overview of the observa- tional properties of sources which emit across 19 orders of magnitude in energy. Young spin-powered pulsars were the only firmly iden- tified (through their pulsations) class of Galactic GeV sources during the EGRET era. Pulsed emission from the longest radio wavelengths up to a few tens of GeV has been detected, although in general the g −ray cut-off for pulsed emission seems to be in the few GeV range [see 4, for a general review of pulsar and pulsar wind nebu- lae properties]. The winds of young pulsars create nebu- lae which emit synchrotron radiation from radio through soft gamma-rays, and through inverse Compton scatter- ing produce photons with energies up to tens of TeV. The transition from synchrotron to inverse Compton emission in PWN seems to be in the ∼ 0.01 − 1 GeV range, al- though only for the Crab and Vela is the synchrotron cut- off energy fairly well established. The cut-off energies of both the pulsed and PWN emission should be greatly constrained by the Fermi (formerly known as GLAST) LAT over the coming years. Inverse Compton (IC) emission is a very efficient means of producing high-energy photons from an accel- erated particle with a large fraction of the electron en- ergy transferred to the seed photon in a single scatter. The peak of the IC emission is typically 5-10 orders of magnitude higher in energy than the corresponding syn- chrotron radiation. It is also not dependent on magnetic field strength, and so can produce significant amounts of radiation even in regions with very low magnetic field strength. The implication of this for emission from PWN is that the synchrotron lifetimes of the IC emitting elec- trons can be very long, and it is not necessarily the case that where the synchrotron emission is bright, the IC emission is bright. In fact, it is typically the case that the TeV emission is significantly offset from the current position of the pulsar. This offset is caused by either the motion of the pulsar or the interaction of the supernova remnant reverse shock with the PWN. The MILAGRO telescope detected two previously un- known regions of significant emission above 10 TeV [5]: a complex of sources in the Cygnus region with the brightest peak known as MGRO J2019+37 and another extended source called MGRO J1908+06. The nearest EGRET source to MGRO J2019+37 is GeV J2020+3658 which is associated with the young pulsar and PWN PSR J2021+3651 [6], which is near the central region of the Cygnus cross. For this reason, and the fact that the pulsar was discovered with Arecibo in Puerto Rico, we some- times refer to it as Cisne (swan in Spanish). We have performed X-ray observations with XMM- Newton and 20cm radio observations with the VLA of the region around PSR J2021+3651. Outside of the bright inner X-ray nebula there structured hard, faint X- ray emission extending out ∼ 10 − 15 arcminutes. The brightest region of this larger X-ray nebula is an exten- sion to the west of about 8′ in length. The VLA image shows a radio nebula coincident with this X-ray exten- sion, with a suggestion of a conical morphology extend- ing out around 10′, at which point there is a decrease in surface brightness and a further broadenig of the neb- ula. The radio nebula extends out at least 20′ to the west, right to the center of the best fit ellipse to MGRO J2019+37. This connection, for any reasonable distance to PSR J2021+3651 (> 6 kpc, see discussion in [6]) re- sults in a TeV luminosity several times that of the Crab nebula, making this the most luminous TeV source in the Galaxy. If we include the EGRET flux, the broad-band spectral energy distribution of the PWN would appear to peak in the GeV range. However, much if not most of the emission seen by EGRET is undoubtably magne- tospheric pulsed emission, therefore it is likely the syn- chrotron emitting PWN cuts off in the MeV range. Un- fortunately, VERITAS has not yet been able to confirm the TeV source, but should ultimately give a much more detailed understanding of the TeV emission and its mor- FIGURE 1. XMM-Newton X-ray image of Cisne, the PWN around PSR J2021+3651 (for color version, red=0.5-1.5 keV, green=1.5-2.5, blue=2.5-7.5) with VLA contours FIGURE 2. VLA 20cm image of Cisne PWN with XMM- Newton X-ray contours phological connection to the PWN of PSR J2021+3651. MGRO J1908+06 is near GeV J1907+0557, for which a potentially extended X-ray counterpart was discovered with ASCA [7]. A brief (10 ks) Chandra ACIS obser- vation showed most of the flux is from a hard (photon index ∼ 1) moderately absorbed (nH ∼ 2 × 1022cm−2) point source with no compact nebula and just a faint hint of the larger nebula seen by ASCA. A moderately deep I band image with the MDM 2.4m telescope shows no ev- idence of an optical counterpart. Both HESS and VERI- TAS (see other contributions, this proceedings) have ver- ified and resolved the MILAGRO source, with the ASCA source at the southern edge. AGILE data of the region verifies the GeV emission is consistent with the position of GeV J1907+0557 and suggests the third EGRET cata- log source 3EG J1903+0550 was a composite of the GeV source and another, unrelated source. VLA ASCA EGRET MILAGRO 1e-09 1e-10 1e-11 1e-12 1e-13 1e-14 s / 2 ^ m c / g r e ) v ( F v 1e-15 1e+08 1e+10 1e+12 1e+14 1e+16 1e+18 1e+20 1e+22 1e+24 1e+26 1e+28 Frequency (Hz) FIGURE 3. Spectral Energy Distribution of Cisne. Note that the EGRET flux undoubtably has a significant pulsed compo- nent. FIGURE 4. GB6 4850MHz image [8] of MGRO J1908+06 region. The white contour is the HESS source, the black con- tour is the ASCA 2-10 keV image (nb. that the circle is due to exposure correction of noise at the edge of the FOV, the source under discussion lies along the outside edge of the HESS contour) and the grey contour is the AGILE > 300 MeV image. Several of the unidentified EGRET sources contain- ing PWN appear to be variable on timescales of months according to the Nolan et al. (2003) variability test [9] with t = Frms/Fmean ∼ 1. This suggests the GeV flux may be partially due to particle acceleration in the inner nebula rather than being pulsed emisssion from the mag- netosphere. Some of these sources (eg. the Rabbit/GeV J1417-6100 and the Eel/GeV J1825-1310 nebulae, see [10]) appear to emit TeV emission as well. However, three potentially variable GeV sources, CTA 1, GeV J1809-2327 (Taz) and SNR W44, have PWN inside of mixed-morphoogy SNR [11]. These supernova remnants have radio shells but are filled with thermal X-rays. Most SNR of this type also contain young pulsars with pul- sar wind nebulae. Two mixed-morphology SNR associ- ated with GeV sources, W28 and IC 443, have had TeV hotspots reported [12, 13], suggesting the three contain- ing variable GeV sources are likely to be TeV emitters as FIGURE 5. Top: ATCA 20cm image of RCW 49. Bottom: MSX 8.3m image. well. Complicating our understanding of these sources are their environments. Many are near molecular clouds, which may both enhance the the production of g −ray emission by particles accelerated in pulsar winds and SNR shocks as well as produce g −ray emission on their own. In addition, there have been a few claims of ex- tended TeV sources associated with massive stars or col- liding wind binaries contained within molecular clouds despite the TeV emission being somewhat offset from the stellar systems. One should keep in mind that the bright diffuse radio emission from molecular clouds can obscure PWN or SNR shells. We have performed ATCA radio imaging of one such source, the Westerlund 2 com- plex (also known as RCW 49) near GeV J1025−5809 which is coincident with HESS J1023−575 [14]. While much of the radio emission can be associated with mid- infrared emission seen by MSX, there are some arcs in the ATCA image that do not correlate with infrared struc- tures. This suggests a non-thermal, i.e. SNR shell, iden- tification for these arcs. RADIO-LOUD HIGH MASS X-RAY BINARIES There are at least 3 strange radio emitting binary sys- tems with massive stars in eccentric orbits. PSR 1259- work was partially supported by SAO Grant No. G07- 8072 and GO6-7136X, NASA Grant No. NNX08AV70G and NNG06EJ54P. G. McSwain acknowledges institu- tional support from Lehigh University. M. Roberts would like to thank the CNRS and the APC for support during his visit. REFERENCES 1. T. C. Weekes, et al. ApJ 342, 379 -- 395 (1989). 2. R. C. Lamb, and D. J. Macomb, ApJ 488, 872 -- + (1997). 3. M. S. E. Roberts, "Pulsars Everywhere! A Galactic EGRET Source Retrospective," in 40 Years of Pulsars: Millisecond Pulsars, Magnetars and More, edited by C. Bassa, Z. Wang, A. Cumming, and V. M. Kaspi, 2008, vol. 983 of American Institute of Physics Conference Series, pp. 621 -- 623. 4. V. M. Kaspi, M. S. E. Roberts, and A. K. Harding, Isolated neutron stars, Compact stellar X-ray sources, 2006, pp. 279 -- 339. 5. A. A. Abdo, et al. ApJL 664, L91 -- L94 (2007) 6. J. W. T. Hessels, M. S. E. Roberts, S. M. Ransom, V. M. Kaspi, R. W. Romani, C. . Ng, P. C. C. Freire, and B. M. Gaensler, ApJ 612, 389 -- 397 (2004). 7. M. S. E. Roberts, R. W. Romani, and N. Kawai, ApJS 8. 133, 451 -- 465 (2001) J. J. Condon, J. J. Broderick, G. A. Seielstad, K. Douglas, and P. C. Gregory, AJ 107, 1829 -- 1833 (1994). P. L. Nolan, W. F. Tompkins, I. A. Grenier, and P. F. Michelson, ApJ 597, 615 -- 627 (2003). 10. M. S. E. Roberts, "What Will Be The Brightest GLAST Sources in the Galaxy?," in The First GLAST Symposium, edited by S. Ritz, P. Michelson, and C. A. Meegan, 2007, vol. 921 of American Institute of Physics Conference Series, pp. 385 -- 386. 11. Roberts, M. S. E., & Brogan, C. L. 2008, ApJ, 681, 320 12. F. Aharonian, et al. A& A 481, 401 -- 410 (2008). 13. J. Albert, et al. ApJL 664, L87 -- L90 (2007) 14. F. Aharonian, et al. A& A 467, 1075 -- 1080 (2007) 15. S. Johnston, L. Ball, N. Wang, and R. N. Manchester, MNRAS 358, 1069 -- 1075 (2005) 16. A. Neronov, and M. Chernyakova, AP& SS 309, 253 -- 259 (2007) 17. E. D. Grundstrom, S. M. Caballero-Nieves, D. R. Gies, W. Huang, M. V. McSwain, S. E. Rafter, R. L. Riddle, S. J. Williams, and D. W. Wingert, ApJ 656, 437 -- 443 (2007) 18. J. Casares, I. Ribas, J. M. Paredes, J. Martí, and C. Allende Prieto, MNRAS 360, 1105 -- 1109 (2005) 19. G. Dubus, A& A 456, 801 -- 817 (2006) 63 is a young pulsar in a 3.6 year highly eccentric orbit around a Be Star. Near periastron, it becomes a moder- ately bright unpulsed broad-band (radio-TeV) source (al- though not yet seen in GeV) [15, 16]. LSI+61 303 and LS 5039 are in much smaller eccentric orbits (26.5 and 3.9 days [17, 18]) which have moderately bright hard X- ray emission but with luminosities about 2-3 orders of magnitude smaller than known accreting systems. They also have moderately bright and variable radio emission (on the order of 100 mJy) extended on milliarcsecond scales and are coincident with point sources of TeV emis- sion [see 19, for an overview]. Their emission appears modulated at the orbital period at all frequencies. Pulsa- tions from LS 5039 and LSI+61 303 have not yet been detected. We are performing deep, high frequency pulse searches with the GBT. Since both of these sources are prime targets for the Fermi telescope, we are monitoring these two sources twice weekly in X-rays with RXTE. In addition, there are supporting optical and radio monitor- ing observations planned for LSI+61 303 and LS 5039 lasting through the first year of GLAST. A CATALOG OF GALACTIC SOURCES The space here is too limited to discuss all of the mul- tiwavelength data of the bright Galactic GeV sources that EGRET detected. An online catalog of multiwave- length observations of Galactic EGRET sources is being produced by M.S.E. Roberts, and can be found at: http://www.physics.mcgill.ca/~pulsar/unidcat.html This catalog is an expansion of the ASCA Catalog of Potential Counterparts of GeV source [7] and is a work in progress. 9. ACKNOWLEDGMENTS 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. The National Radio Astronomy Ob- servatory Very Large Array is a facility of the National Science Foundation operated under cooperative agree- ment by Associated Universities, Inc. This research made use of data products from the Midcourse Space Experi- ment, processing of which data was funded by the Ballis- tic Missile Defense Organization with additional support from NASA Office of Space Science. The Australia Tele- scope Compact Array is part of the Australia Telescope which is funded by the Commonwealth of Australia for operation as a National Facility managed by CSIRO. Based on observations obtained with XMM-Newton, an ESA science mission with instruments and contributions directly funded by ESA Member States and NASA. This
0709.3988
2
0709
2008-04-13T02:43:16
Do f(R) theories matter?
[ "astro-ph" ]
We consider a modified action functional with a non-minimum coupling between the scalar curvature and the matter Lagrangian, and study its consequences on stellar equilibrium. Particular attention is paid to the validity of the Newtonian regime, and on the boundary and exterior matching conditions, as well as on the redefinition of the metric components. Comparison with solar observables is achieved through numerical analysis, and constraints on the non-minimum coupling are discussed.
astro-ph
astro-ph
Do f (R) theories matter? O. Bertolami∗ and J. P´aramos† Instituto Superior T´ecnico, Departamento de F´ısica‡, Av. Rovisco Pais 1, 1049-001 Lisboa, Portugal (Dated: October 30, 2018) We consider a modified action functional with a non-minimum coupling between the scalar cur- vature and the matter Lagrangian, and study its consequences on stellar equilibrium. Particular attention is paid to the validity of the Newtonian regime, and on the boundary and exterior match- ing conditions, as well as on the redefinition of the metric components. Comparison with solar observables is achieved through numerical analysis, and constraints on the non-minimum coupling are discussed. PACS numbers: 04.20.Fy, 04.80.Cc, 97.10.Cv Preprint DF/IST-7.2007 I. INTRODUCTION Modern cosmology faces two outstanding challenges, namely the existence and nature of dark energy and dark matter. Many theories have been put forward to address both issues: for dark matter, several candidates are avail- able, such as weak-interacting particles (WIMPs) aris- ing from extensions to the Standard Model (e.g. axions, neutralinos), etc.; for dark energy, "quintessence" mod- els consider the slow-roll of a scalar field [1, 2], amongst other candidates; others suggest that the averaging of inhomogeneities at a cosmological scale may yield an ef- fective scalar field, thus accounting for the dark energy component of the Universe [3]. A possible unification of both "dark" components has also been suggested, resort- ing to a scalar field model [4] or an exotic equation of state, as featured by the so-called modified Chaplygin gas [5]. A different approach assumes that no extra energy content is needed and that the fundamental laws and tenets of gravitation may be incomplete, perhaps just a low-energy approximation; as a consequence, modifica- tions of the Friedmann equation to include higher order terms in the energy density ρ (see e.g. [6] and references therein) have been proposed or, at a more fundamental level, changes to the action functional. A rather straight forward approach lies in replacing the linear scalar cur- vature term in the Einstein-Hilbert action by a function of the scalar curvature, f (R); alternatively, one could re- sort to other scalar invariants of the theory [7]. This has led to some success in replicating the accelerated expan- sion of the Universe, while comparison with its evolu- tion throughout the different ages has yielded some con- straints and exclusions to the form of f (R); perhaps the most well-known proposal of this type is the Starobin- sky inflationary model, where a quadratic term in the curvature is added to the usual linear form (plus cosmo- ‡Also at Instituto de Plasmas e Fusao Nuclear, IST, Lisbon ∗Electronic address: [email protected] †Electronic address: [email protected] logical constant), f (R) = R − Λ + αR2 [8]. Solar system tests could also bring further insight, mostly arising from the parameterized post-Newtonian (PPN) metric coeffi- cients derived from this extension of general relativity (GR). However, some disagreement exists in the commu- nity, with some arguing that no changes are predicted at a post-Newtonian level (see e.g. [9] and references therein); amongst other considerations, this mostly stems from an approach based either in the more usual metric affine connection (that is, where the affine connection is taken a priori as depending on the metric), or in the so-called Palatini approach [10] (where both the met- ric and the affine connection are taken as independent variables). As an example of a clear phenomenological consequence of this extension of GR, it has been shown that f (R) = f0Rn theories yield a gravitational poten- tial which displays an increasing, repulsive contribution, added to the Newtonian term [11]. Notwithstanding the significant literature on these f (R) models, few steps have been taken to address an- other interesting possibility: not only that the curvature is non-trivial in the Einstein-Hilbert Lagrangian, but also that the coupling between matter and geometry is not minimum; indeed, these are only implicitly related in the action functional, since one expects that covariantly in- variant terms in Lm should be constructed by contrac- tion with the metric (e.g. the kinetic term of a real scalar field, gµνφ,µφ,ν ). A non-minimum coupling would imply that geometric quantities (such as the scalar invariants) would explicitly show in the action; asides from theo- retical elegance, this could have deep phenomenological implications: indeed, in regions where the curvature is high (which, in GR, are related to regions of high en- ergy density or pressure), the implications of such the- ory could deviate considerably from those predicted by Einstein's theory [12]. Related proposals have been put forward previously to address the problem of the accel- erated expansion of the Universe [13] and the existence of a cosmological constant [14]. In this sense, the immediate question, posed in the ti- tle of this work, is: what are the implications for the be- haviour of matter under such conditions? Some work has been put forward concerning this issue, namely changes to geodetic behaviour [12], the possibility of modelling dark matter [15] and the violation of the highly con- strained equivalence principle [16]. Perhaps the most important consequence of the mentioned studies is that energy may no longer be covariantly conserved, that is, ∇µTµν 6= 0, where Tµν is the energy-momentum tensor of matter; this occurs because, due to the presence of extra terms in the equations of motion, the Bianchi iden- tities no longer imply in the covariant conservation of the energy-momentum tensor. This work addresses what we believe is the natu- ral proving ground for a non-minimally coupled gravity model: regions where the density may be high enough, to evidence some deviation from GR, although moderate enough so that effects are still perturbative -- a star. The results rely upon and expand the methodology followed by the authors in previous studies [17, 18]. This paper is divided in the following sections: first, we present the model upon which the subsequent work is based; then, we develop the equations of motion, aiming at the modified Tolman-Oppenheimer-Volkoff (TOV) equation, with due care taken regarding the validity of the Newtonian regime and resulting modified hydrostatic equilibrium equation; afterwards, we insert the polytropic equation of state into the latter, and compute the necessary observables; a nu- merical session follows, where profiles and bounds are computed for the relevant quantities; finally, a discussion of our results is presented. II. THE MODEL Following the discussion of the previous section, one postulates the following action for the theory [12], S =Z (cid:20) 1 2 f1(R) + [1 + λf2(R)]Lm(cid:21) √−gd4x (1) where fi(R) (with i = 1, 2) are arbitrary functions of the scalar curvature R, Lm is the Lagrangian density of mat- ter and g is the metric determinant. For convenience, the contribution of the non-minimum coupling of f2 is gauged through the coupling constant λ (which has dimensions [λ] = [f2]−1). The standard Einstein-Hilbert action is recovered by taking f2 = 0 and f1 = 2κ(R − 2Λ), where κ = c4/16πG and Λ is the cosmological constant (from now on, one works in a unit system where c = 1). Variation with respect to the metric gµν yields the modified Einstein equations of motion, here arranged as 1 2 f1gµν = (F1 + 2λF2Lm) Rµν − ( µν − gµν ) (F1 + 2λF2Lm) + [1 + λf2] Tµν , µν ≡ ∇µ∇ν for convenience, as well where one defines as Fi(R) ≡ f ′(R), and omitted the argument. The mat- ter energy-momentum tensor is, as usually, defined by (2) Tµν = − 2 √−g δ (√−gLm) δgµν . 2 (3) As stated before, the Bianchi identities, ∇µGµν = 0 imply the non-(covariant) conservation law ∇µTµν = λF2 1 + λf2 (gµνLm − Tµν )∇µR , (4) and, as expected, in the GR limit λ → 0, one recovers the conservation law ∇µTµν = 0. A. Scope of application (2), It is our stated purpose to arrive at the modified form of the TOV equation, the relativistic version of the hy- drostatic equilibrium condition. From Eq. it is clear that a full treatment of the equations of motion is unattainable unless some specific form for f1(R) and f2(R) is provided. Furthermore, the presence of both the pure curvature and non-minimum coupling, respectively, still constitutes a daunting analytical challenge; hence, and since one is mainly interested in the relevance of the effects within a high curvature and pressure medium, where f2 should overwhelm the modification of the pure curvature term, f1 − 2κR (neglecting the contribution of the cosmological constant), it appears sensible to discard the latter; a treatment of the standard f (R) scenario with f2 = 0 may be found in Ref. [19]. Mathematically, the chosen approximation reads as 1 2f1 − κR ≪ λf2Lm F1 − 2κ ≪ 2λF2Lm , , ( µν − gµν ) F1 ≪ 2 λ ( µν − gµν (5) ) (F2Lm) . where the second and third inequalities indicate that this regime stems not just from the comparison between con- tributions to the action functional, but also those in- volved in the modified Einstein equations (2); also, notice that the perturbative condition λf2(R) ≪ 1 has not been enforced yet. The last inequality is satisfied if the follow- ing stronger conditions hold, d(F2Lm) dr d2(F2Lm) λ dr2 (cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) λ dF1 d2F1 dr (cid:12)(cid:12)(cid:12)(cid:12) ≪ 2(cid:12)(cid:12)(cid:12)(cid:12) dr2 (cid:12)(cid:12)(cid:12)(cid:12) ≪ 2(cid:12)(cid:12)(cid:12)(cid:12) This said, one considers the simplest form f2 = R; this linear coupling could arise from a first order expansion of a more general f2(R) function, in the weak field en- viron of the Sun. Also, one takes Lm = p, a natural , (6) (cid:12)(cid:12)(cid:12)(cid:12) . (cid:12)(cid:12)(cid:12)(cid:12) choice for the Lagrangian density of an ideal fluid [20] -- characterised by the standard energy-momentum tensor Tµν = (ρ + p)uµuν − pgµν , (7) where p is the fluid's pressure, ρ its matter-energy density and u its 4-velocity vector, with uµ = (u0,~0), uµ µ = 1, so that u0 = g1/2 Inserting the above expressions for f2 00 . and Lm, the inequalities (5) become (cid:12)(cid:12)(cid:12)(cid:12) dr (cid:12)(cid:12)(cid:12)(cid:12) , f1 ≪ 2 (κ + λp) R , F1 ≪ 2 κ + λp dF1 ≪ 2 λp′(r) , (8) (9) ≪ 2λp′′(r) , (10) (cid:12)(cid:12)(cid:12)(cid:12) d2F1 dr2 (cid:12)(cid:12)(cid:12)(cid:12) where the prime denotes derivation with respect to the radial coordinate. Clearly, if f1 = 2κR → F1 = 2κ, these are trivially satisfied. Also, notice that f2 = R implies that [λ] = [f −1 2 ] = [R]−1 = M −2. [2] One must now ascertain the form of f1(R). In Ref. it is shown that acceptable models with f1a = 2κ(R − αR1−m) (α > 0, [α] = [Rm] = M 2m and 0 < m < 1 or f1b = 2κ(R + αR2 − Λ) (with αΛ ≪ 1 and [α] = [R−1] = M −2) are cosmologically viable; the latter may arise from the renomalizability of the theory near the Planck scale, with α ∼ M −2 and M = 1012 GeV [2]. For the perturbative regime to be valid, one must have, for the f1 = f1a case 3 1 2 R (gµν + aTµν) = (13) (1 + ap) Rµν − a( µν − gµν )p + 1 2κ Tµν , where one defines the parameter a ≡ λ/κ = 16πGλ, with dimension [a] = M −4; accordingly, both ap and aρ are dimensionless quantities. In the above form, the physi- cal meaning of the proposed model is more transparent: aside from a pressure-dependent term on the r.h.s., the most interesting modification occurs on the l.h.s.; firstly, the contribution of the Riemann tensor is modified by a factor 1 + ap; secondly, the scalar curvature is cou- pled not only to the metric gµν, but also to the energy- momentum tensor Tµν. One thus gets a clear picture of the matter-geometry interaction, which occurs via scalar- tensor combinations. By taking the trace of the above equation, one obtains R = − −3aκ p + T κ [2 + a (T − 2p)] = 3p − ρ + 3aκ p κ [2 + a(ρ − 5p)] , (14) having inserted T = T µ µ = ρ − 3p. Interestingly enough, in the "strong" a → ∞ regime it yields the asymptotic re- sult R = 3 p/(ρ−5p): in this regime, a varying, low den- sity can give origin to extremely high curvatures, while an almost uniform, even if a high density fluid might yield a vanishingly small curvature. (11) C. Perturbative regime , 1 λ ≫ ακ(cid:12)(cid:12)(cid:12)(cid:12) Rmp(cid:12)(cid:12)(cid:12)(cid:12) λ ≫ (1 − m)mακ(cid:12)(cid:12)(cid:12)(cid:12) λ ≫ (1 − m)mακ(cid:12)(cid:12)(cid:12)(cid:12) , 1 1 dR R1+mp′ dr(cid:12)(cid:12)(cid:12)(cid:12) dr(cid:1)2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) R (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) dr2 − (1 + m)(cid:0) dR R1+mp′′(cid:12)(cid:12)(cid:12)(cid:12) d2R . Foreseeing a later comparison with solar observables, which one knows are well predicted by GR, it is now assumed that the effect of f2 yields only perturbative corrections: in the approximation λf2 = λR ≪ 1, the scalar curvature is given by R ≈ (3p − ρ)/κ, resulting in (λ/κ)(3p − ρ) = a(3p − ρ) ≪ 1; anticipating the New- tonian approximation p ≪ ρ, this yields aρ ≪ 1 and ap ≪ 1. Inserting Eq. (14) in the Einstein equation, one obtains, after some algebraic manipulation, while, for the f1 = f1b case (which can be derived from the above inequalities, setting m = −1, α → −α and ignoring the cosmological constant term), , R dR p(cid:12)(cid:12)(cid:12)(cid:12) λ ≫ ακ(cid:12)(cid:12)(cid:12)(cid:12) λ ≫ 2ακ(cid:12)(cid:12)(cid:12)(cid:12) dr(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) p′′(cid:12)(cid:12)(cid:12)(cid:12) dr2 (cid:12)(cid:12)(cid:12)(cid:12) λ ≫ 2ακ(cid:12)(cid:12)(cid:12)(cid:12) 1 p′ 1 d2R , (12) κ[2 + a(ρ − 3p)]Rµν = (3p − ρ)gµν + 2(1 − ap)Tµν + aκ(4 µν − gµν (15) )p , keeping only first order terms in a. This shall be the main tool for the derivation of the TOV equation. . III. STELLAR EQUILIBRIUM B. Equation of motion A. Static, spherical symmetric scenario With the above choice for f1 and f2, the equation of motion becomes Since one is dealing with an ideal, spherically symmet- ric system, in which temporal variations are assumed to occur only at the cosmological scale H −1 0 , and hence neg- ligible at an astrophysical time scale, one considers the Birkhoff metric (in its anisotropic form), given by the line element ds2 = eν(r)dt2 −(cid:16)eσ(r)dr2 + dΩ2(cid:17) , with dΩ = r2(dθ2 + sin2θ dφ2), so that √−g = r2 sinθ e(ν+σ)/2. (16) B. Newtonian limit 4 Before continuing the derivation, it is opportune to ad- dress the issue of the validity of the Newtonian regime; clearly, establishing this limit will enable many fruitful simplifications in the calculations ahead. In the stan- dard derivation of the hydrostatic equilibrium equation, this arises from a set of simplifications imposed on the relativistic TOV equation; the later reads, One can resort to the expression of the Riemann tensor and Eq. (15) to obtain the following intermediary step (the full algebraic derivation is shown in Appendix A): p′(r) = − G r2 [ρ(r) + p(r)][me(r) + 4πp(r)r3] 1 − 2Gme(r)/r , (22) κ[2 + a(ρ − 3p)](cid:20) 1 ap 2 (ρ + 3p) + 1 r2 + e−σ(cid:18) σ′ r − 2 (cid:18)5e−ν 00 + 3e−σ r2(cid:19)(cid:21) = ρ − (17) r2 (cid:19) p . rr + 2 θθ aκ One now defines the parameter me, here called the ef- fective mass, to distinguish it from its identification with the gravitational mass of the unperturbed GR scenario, derived from the Schwarzschild metric; it is given by the usual expression e−σ = 1 − 2Gme r , which yields r2 + e−σ(cid:18) σ′ (cid:20) 1 r − 1 r2(cid:19)(cid:21) = 2Gm′ e r2 , (18) (19) where the prime denotes differentiation with respect to the radial coordinate. Inserting the above in Eq. (17) and solving for m′ e, one gets and the Newtonian approximation is valid if the following inequalities are satisfied r ≫ 2Gme(r) , ρ(r) ≫ p(r) , me(r) ≫ 4πp(r)r3 , (23) yielding the non-relativistic hydrostatic equation of state p′(r) = −G ρ(r)me(r) r2 . (24) Since is is assumed that the coupling between mat- ter and geometry will only produce a perturbative effect, leading to the redefinition of mass through Eq. (21), it is clear that no changes occur regarding the validity of the Newtonian approximation. Hence, foreseeing the ap- plication of the following results to the Sun, where such regime is valid, one may simplify the intermediate calcu- lations and, when convenient, insert the inequalities (23) in order to simplify a-dependent terms. By the same to- ken, terms involving the coupling between a-dependent quantities and any covariant derivatives may be evalu- ated by taking their Newtonian counterparts, eν−σ 00p = − 2 rrp = p′′ + θθp = e−σrp′ ≈ rp′ ν′p′ ≈ 0 , σ′p′ ≈ p′′ . 1 2 (25) , m′ e = 4πr2ρ ar2 4G (cid:0)5e−ν 2 − ap(cid:16)1 + 3p ρ (cid:17) 2 + a(ρ − 3p) 00 + 3e−σ 2 + a(ρ − 3p) rr + 2 r2 + (20) θθ(cid:1) p . Notice that this does not imply that one is directly de- riving the Newtonian hydrostatic equilibrium equation, since leading order terms will not be approximated. In the perturbative regime, one may take only first-order terms in ap and aρ, obtaining C. Tolman-Oppenheimer-Volkoff equation p2 m′ e ≈ 4πr2ρ(cid:20)1 + a(cid:18)p − 8G (cid:18)5e−ν 00 + 3e−σ ρ 2 − rr + 2 θθ ρ (cid:19)(cid:21) + r2 (cid:19) p , 3 2 ar2 using 4Gκ = 1/4π. The above expression clearly shows the perturbation to the purely gravitational mass, de- fined by m′ g = 4πr2ρ. (21) By following a procedure similar to the one leading to Eq. (17) (depicted in Appendix A), one obtains the following equation: κ [2 + a (ρ − 3p)](cid:20)(cid:18)1 − 2 (cid:18)3e−ν (p − ρ) + ap 2 p + aκ 2Gme r (cid:19) ν′ r − 00 + 5e−σ 2Gme r3 (cid:21) = (26) r2 (cid:19) p . rr − 2 θθ Substituting by the expressions for the covariant deriva- tives, one gets 00p ≈ 0 , rrp ≈ p′′ , θθp ≈ rp′ . (33) 5 (27) This approximation, together with p ≪ ρ, implies that rr − 2 θθ r2 (cid:19) p = 00 + 5e−σ 2Gme r (cid:19) p′′ + (cid:18)3e−ν 5(cid:18)1 − (cid:20)5 1 Gme Gm′ e 2Gme r (cid:18)2 + r (cid:19) 3ν′ 2 − r −(cid:18)1 − r (cid:19)(cid:21) p′ One now introduces the approximations discussed in the previous subsection. Also, the approximation m′ e ≈ g = 4πρr2 is taken, since the perturbative corrections m′ would produce second order terms in a. One obtains, after a little algebra, . 2Gme κ (2 + aρ)(cid:18) ν′ r3 (cid:19) = r − 2 (cid:20)5p′′ +(cid:18)5 aκ ρ(cid:17) + p(cid:16)1 − a 2 Solving for ν′, one gets (28) Gm′ e r − 3ν′ 2 − 2 r(cid:19) p′(cid:21) . ν′ 2 = G (2 + aρ) me + 4 (2 − aρ) πpr3 r (r − 2Gme)(cid:2)2 + a(cid:0)ρ + 3 4 rp′(cid:1)(cid:3) 4 p′′ +(cid:0)5πGρr − 1 2r(cid:1) p′ 2 + a(cid:0)ρ + 3 4 rp′(cid:1) r − 2Gme r2 5 a When a = 0, one recovers the standard expression + (29) with . m′ e ≈ 4πr2ρ + ar2(cid:18) 3p′′ 8G + p′ 4Gr − 2πρ2(cid:19) . (34) Similarly, one obtains κ (2 + aρ) R ≈ −ρ − 3aκ(cid:0)p′′ + 2 r p′(cid:1) 2(cid:20)3(cid:18)p′′ + ρ 2κ − − 2 r a (35) ≈ p′(cid:19) − ρ2 2κ(cid:21) . Although the expression for the scalar curvature R is somewhat involved, one can approximate its derivative by the unperturbed value R′ = −ρ′/2κ, losing only terms of order O(a2); indeed, one could write R′ = − λR′ = − ρ′ 2κ λρ′ 2κ + aF (ρ, p) → + aλF (ρ, p) = − a 2 (36) ρ′ + a2 F (ρ, p) κ , F (ρ, p) = − 1 2(cid:20)3(cid:18)p′′ + 2 r p′(cid:19) − ρ2 2κ(cid:21)′ . (37) To write the modified equation of hydrostatic equilib- rium, one now resorts to the non-(covariant )conservation of the energy-momentum tensor, Eq. (4). The expression for the covariant derivative is purely geometric; in order to derive an expression for p′, one aims at the ν = r com- ponent of the above equation; with our choice of f2 = R and Lm = p, one gets the modified TOV equation ν′ 2 (ρ + p) = 2λ 1 + λR (cid:18)1 − 2Gme r (cid:19) R′ → (38) (ρ + p) ≈ p′ − 2λp 1 + λR R′ ≈ p′ + apρ′ , p′ + ν′ 2 − dropping higher order terms and using R′ = −ρ′/2κ. equations for the problem at hand: Thus, one finally obtains a set of three differential (ρ + p) , ν′ 2 p′ − 2λpR′ = − e = 4πr2ρ + ar2(cid:18) 3p′′ m′ 8G + ν′ 2 = G me + 4πpr3 r2 − 2Gmer + a(cid:20)(cid:18) 5 p′ 4Gr − 2πρ2(cid:19) , p′′ − 4πGpρ(cid:19) r − 8 (39) (40) p′ 4(cid:21) .(41) me + 4πpr3 r2 − 2Gmer Linearizing with respect to a yields ν′ 2 = G . (30) ν′ 2 = G me + 4πpr3 r2 − 2Gmer + ah(p, ρ) , (31) where the function h(p, ρ) is defined through h(p, ρ) = (32) , 1 8 8 2 3 8 p′ 4 Gρr2 − Gme r − 4(cid:19) p′ ≈ p′′ − 4πGpρ(cid:19) r +(cid:18) 5π p′′ − 4πGpρ(cid:19) r − (cid:18) 5 (cid:18) 5 after considering the inequality r ≫ 2Gme and also 5πGρr2/2 ≪ 1; taking the Sun's maximum, central, density ρc = 1.622 × 105 kg/m3, and the radius R⊙ = 6.955 × 108 m, one gets 5πGρr2/2c2 = 4.57 × 10−4 ≪ 1. Before continuing, one can rewrite the expression for the gravitational mass, obtained before, but imposing the limit e−σ ≈ e−ν ≈ 1, ν′ ≈ σ′ ≈ 0, which yields Substituting Eq. (39) into (41), after some algebra one gets the modified TOV equation, p′ + G(ρ + p) = me + 4πpr3 r2 − 2Gmer p′ p′′ − 4πGpρ(cid:21) r − a(cid:20)(cid:18)(cid:20) 5 8 4(cid:19) ρ + pρ′(cid:21) . (42) This yields the non-relativistic hydrostatic equilibrium equation, p′ + Gmeρ r2 = a(cid:20)(cid:18)(cid:20) 5 8 p′′ − 4πGpρ(cid:21) r − p′ 4(cid:19) ρ + pρ′(cid:21) . (43) D. Polytropic equation of state Realistic stellar models rely on four differential equa- tions, together with appropriate definitions [21]; aside from the mass conservation condition (40) and the hy- drostatic equilibrium equation (42) or (43), these express energy conservation and transport, through and dL(r) dr = 4πr2ǫρ(r) dT (r) dr = − L(r) 4πr2λc , (44) (45) respectively. In the above, L(r) is the energy flow across a sphere of radius r, ǫ is the energy generation rate per mass unit, and λc is the conductivity coefficient. For given ǫ and λc, which account for the processes ongoing inside the star, one is left with five unknowns, m′ e, p, ρ, T and L; an additional relation is needed, in the form of a suitable equation of state p = p(ρ). Many candidate equations of state and solar models are available, with varying degrees of sofistication, account- ing for effects such as chemical composition, solar mat- ter mixing, discontinuities between layers, heat difusion, etc.; two outstanding examples are the Mihalas-Hummer- Dappen and OPAL equations of state [22]. However, solving the above set of differential equations with a re- alistic equation of state requires heavy-duty numerical integration with complex code designs, and an analysis of the perturbations induced by a non-minimum coupling between matter and curvature is beyond the scope of the present study. Instead, one may resort to a very simplistic assump- tion, the so-called polytropic equation of state. This is commonly given by p = Kρ(n+1)/n , where K is the polytropic constant, ρ0 is the baryonic mass density and n is the polytropic index [21, 23, 24]. The polytropic B 6 index n interpolates between the basic thermodynami- cal processes: n = −1 for isobaric, n = 0 for isomet- ric, and an infinite polytropic index n for an isothermal sphere. Adiabatic processes yield n = 1/(γA − 1), with γA = cp/cV the adiabatic coefficient. Several crude ap- proximations to relevant astrophysical systems are also obtained: n = 3/2 may model the degenerate star cores found in giant (gaseous) planets, white or brown dwarfs,and red giants; a polytropic index n = 5 yields a boundless system (that is, with non-vanishing density ev- erywhere), which was taken by Schuster as the first can- didate for a stellar system. Finally, a polytropic equation of state was used by Eddington in his proposal for the first solar model, with n = 3; naturally, it does not offer an accurate description of the solar interior, and has been deprecated by the following advancements. Nonetheless, the use of a polytropic equation of state is still of interest, due to its simplicity and ease of ma- nipulation, which render it a valuable tool in more the- oretically driven studies, as is the present case (as an example, the generalized Chaplygin gas is a polytrope with index n = −1/(1 + α) [18]). Clearly, more realis- tic assumptions regarding the structure of the Sun would improve the final results; also, the procedure outlined in this work could also be applied to more exotic bodies, either through the use of an adequate polytropic index, or via a more realistic equation of state, possibly yielding a more stringent constraint on the coupling parameter λ. Recall that the effective mass me is defined in terms of the energy density ρ, which appears in the energy- momentum tensor; these two quantities are related through ρ = ρB + np, yielding the relation ρ =(cid:16) p K(cid:17)n/(n+1) + np . (46) However, since one is interested in probing the Newtonian regime which occurs in the Sun, the condition p ≪ ρ holds; therefore, ρ ≃ ρB, and one may take the form p = Kρ(n+1)/n for the equation of state [25]. In order to transform the modified hydrostatic equilib- rium Eq. (43) into a differential equation with a single variable, one first rewrites Eq. (43) as 1 r2 (cid:20) r2 ρ (cid:18)p′ + a(cid:20)(cid:18)(cid:20) 5 = −4πGρ − a(cid:18) 3p′′ 8 8 p′′ − 4πGpρ(cid:21) r − 4r − 2πGρ2(cid:19) , p′ + p′ 4(cid:19) ρ + pρ′(cid:21)(cid:19)(cid:21)′ (47) and inserts the polytropic equation of state, p = Kρ(n+1)/n, written as ρ = ρcθn(ξ) and p = pcθn+1(ξ), with ξ = r/r0 a dimensionless variable and r2 0 ≡ (n + 1)pc/4πGρ2 c; as stated before, ρc = 1.622 × 105 kg/m3 is the central density, and pc = 2.48× 1016 P a is the central pressure. One obtains the perturbed Lane- Emden equation for the function θ(ξ): 1 ξ2 (cid:20)ξ2θ′(cid:18)1 + Acθn(cid:20)(cid:20) 5 8(cid:18)θ′′ + n θ′2 θ (cid:19) − Ncθn+1(cid:21) ξ θ′ + 4(n + 1)(cid:21)(cid:19)(cid:21)′ 3n − 1 = −θn(cid:20)1 + Ac(cid:18) 3 8(cid:20)θ′′ + n θ′2 θ (cid:21) + θ′ 4ξ − 7 θn 2 (cid:19)(cid:21) , (48) where the prime now denotes derivation with respect to the dimensionless radial coordinate ξ, and one has de- fined Ac ≡ aρc and Nc ≡ pc/ρc = 1.7 × 10−6, for con- venience. Obviously, setting Ac = 0 one recovers the unperturbed Lane-Emden equation [21] 1 ξ2 (cid:0)ξ2θ′(cid:1)′ = −θn . (49) One may try to solve Eq. (48) analytically around ξ = 0, and compare with the solution of the unperturbed equation, given (in the vicinity of ξ = 0) by [21] θ(ξ) ≈ 1 − 1 6 ξ2 + n 120 ξ4 . This calculation (outlined in Appendix B) yields θ(ξ) ≈ 1 − Aξ2 + Bξ4 , A = 1 6(cid:18)1 − Ac 13 + 25n 12(n + 1)(cid:19) , with and B = n 120(cid:18)1 − Ac 11 60 39 + 59n n + 1 (cid:19) , (50) (51) (52) (53) neglecting the Nc ≪ 1 term and taking the limit Ac ≪ 1. Taking Ac = 0 gives back the unperturbed values A = 1/6 and B = n/120, as expected. For n = 0 (a constant density model), one finds θ(ξ) ≈ 1 − 1 6(cid:18)1 − 11 6 Ac(cid:19) ξ2 . (54) For n = 3 (the first proposed model for the Sun), one obtains θ(ξ) ≈ 1− 1 6(cid:18)1 − 11 6 Ac(cid:19) ξ2 + 1 40(cid:18)1 − 99 10 Ac(cid:19) ξ4 . (55) Finally, for n → ∞ (an isothermal sphere), one gets Ac(cid:19) ξ4 . Ac(cid:19) ξ2 + 40(cid:18)1 − 6(cid:18)1 − θ(ξ) ≈ 1 − 1837 180 23 12 1 1 (56) E. Perturbative solution Inspection of Eq. (48) shows that it is a third-order equation, while the unperturbed Lane-Emden equation is only second-order. Thus, the initial conditions θ(0) = 1 and θ′(0) = 0 alone are not sufficient. Indeed, an extra condition regarding the initial behaviour of the second derivative must be provided; this is derived from the se- ries expansion of θ around ξ = 0 taken before, that is θ′′(0) = − 1 3(cid:18)1 − Ac 13 + 25n 12(n + 1)(cid:19) . (57) In order to numerically solve the perturbed Lane- (48) (after a long Emden equation, one rewrites Eq. derivation) as θ′′′ = − 24 5 Nc + 4 5 θn(cid:20) 1 ξ . 2 8 θn ξ θ′ 5Acξ " θ′′ + 2 θn+1 5(cid:20)4 ξ − + 2(2n + 1)Ncθ′(cid:21) − n + 1# − n(n − 1) 7n − 1 n + 1 θ (cid:20)3θ′′ + 3n + 2 n + 1 θ′′ ξ θ′ + θ′3 θ2 (58) θ′ ξ2(cid:21) + 8 5 3n + 2 n + 1 θ′ ξ (cid:21) Clearly, when Ac → 0, the first term blows up unless it is compensated by the condition θ′′ + 2θ′/ξ = −θn, which is precisely the unperturbed Lane-Emden equation. By the same token, the first term expresses the deviation from the unperturbed case, and should be of order Ac, therefore cancelling out the divergence. However, implementing the above third order differ- ential equation proves too computationally demanding; thus, one must first approach the perturbed Lane-Emden equation and, given that one is searching for a perturba- tion, expand it in terms of θ = θ0(1 + Acδ), where θ0 is the solution to the unperturbed Lane-Emden equation and δ is the (relative) perturbation, obeying Acδ ≪ 1. Thus, one may write, for n > 0, θn = θn 0 (1 + Acδ)n ≈ θn 0 (1 + nAcδ) . (59) If this expansion is introduced in the perturbed Lane- Emden equation, and considering only terms of order Ac, one obtains a differential equation for δ, with θ0 as source: δ′′ + 2(cid:18) θ′ 0 θ0 + 1 ξ(cid:19) δ′ + (n − 1)θn−1 0 δ = 5n 2 ξθ2n−2 0 θ′ 0 (60) σ′ −(R⊙) = −2G 8 , (63) +(2n + 1)Ncξθ2n−1 0 θ′ 0 + θ2n−1 0 + 3Ncθ2n 0 9n + 5 4(n + 1) 5n(n − 1) 8 − ξθn−3 0 θ′3 0 + n(3n + 7) 4(n + 1) θn−2 0 θ′2 0 + θ′ 0 1 2 θn−1 0 ξ . having eliminated the second derivative of the unper- turbed solution through Eq. (49). From Eq. (51), one concludes that this differential equation is supplemented by the initial conditions δ(0) = 0 and δ′(0) = 0. Note that δ does not depend on Ac: one must only find the unique δ (for each Nc and n); one aims at a simultaneous variation of the model's parameter Ac and the polytropic index n in the vicinity of the standard solar value n = 3, enabling the plotting of an exclusion graph in the (Ac, n) plane. For n = 3, the differential equation for δ simplifies to δ′′ + 2(cid:18) θ′ 0 θ0 + 1 ξ(cid:19) δ′ + 2θ2 7Ncξθ5 0θ′ 0 + ξθ4 0θ′ 0 + 0δ = 3Ncθ6 0 + 2θ5 0 + (61) 15 2 1 2ξ θ2 0θ′ 0 + 3θ0θ′2 0 − 15 4 ξθ′3 0 . F. Mass budget and matching conditions One now computes the deviation between the effective mass me, defined in Eq. (21), and the gravitational mass, defined by m′ g = 4πr2ρ; some algebra yields m′ e − m′ g = n + 1 G AcNcξ2θn 0 (cid:20)− 3n 8 θ′2 0 θ0 + θ′ 0 2ξ + 7 8 θn (62) 0(cid:21) , which is explicitly written in terms of Ac = aρc, the model's dimensionless parameter, and Nc ≡ pc/ρc, the ratio that measures the validity of the Newtonian regime. One has also to eliminate δ′′ 0 through Eq. (49); one con- cludes that not only is the above mass difference small (due to the perturbative regime, Ac ≪ 1), but that it is further suppressed by the factor Nc ≪ 1. Clearly, the model's parameters should be adjusted to the known observables: the star's radius R⊙ and mass M . The issue of this identification is rather delicate; indeed, M should not be identified with the gravitational mass mg(R⊙), but with the total effective mass me(R⊙). Asides from the physical interpretation of the full energy content of the star affecting geodetic motion around it, one can also resort to the matching conditions of the inner metric with the outer Schwarzschild metric; indeed, writing Me ≡ me(R⊙), Mg ≡ mg(R⊙), M ′ e(R⊙) and M ′ e ≡ m′ g(R⊙), one has g ≡ m′ Me − M ′ eR⊙ R⊙(R⊙ − 2GMe) , GM R⊙(R⊙ − 2GM ) + GMe σ′ + = −2 ν′ −(R⊙) = 2 R⊙(R⊙ − 2GMe) 4 (cid:18) 5 a ν′ + = −σ′ + = 2 2 GM p′′(R⊙)R⊙ − p′(R⊙)(cid:19) , R⊙(R⊙ − 2GM ) , where the + and − subscripts indicate inner or outer boundary condition. If one identifies M = Me, it follows that M ′ e = 0, which implies 3p′′(R⊙) 2 + p′(R⊙) R⊙ = 0 , (after resorting to Eq. (21), with ρ(R⊙) = 0), and 5 2 p′′(R⊙)R⊙ − p′(R⊙) = 0 . Clearly, both conditions hold, since (64) (65) p′(R⊙) ∝ θ′(ξ1)θn(ξ1) = 0 , p′′(R⊙) ∝ θ′′(ξ1)θ(ξ1)n + nθ′(ξ1)2θ(ξ1)n−1 = 0 , (66) where ξ1 signals the star's boundary, through θ(ξ1) = 0. G. A solution for the divergence problem Given the identification M = Me ≡ me(R⊙), one can now deduce the perturbation to the central density ρc, which is a model dependent parameter. However, before using Eq. (48) to evaluate the perturbation δ and extract relevant quantities, it should be noticed that, by inspec- tion, it is clear that δ diverges when θ0 approaches zero. Hence, one cannot extend the perturbational approach to the full range of the star, and should instead deal with the full differential equation for the perturbed θ. In order to circumvent this issue, recall that there is a pronounced deviation between the predictions of the polytropic model and the realistic Standard Solar Model, for r > Rr = 0.713 R⊙; this reflects the crossing from the radiative zone, where the n = 3 polytrope is a good ap- proximation for the Sun, and the convection zone, where this approach fails. Hence, the irregular behaviour of δ near ξ1 may be safely disregarded, since the fundamen- tal equation for θ is not valid there. Instead, one shall consider only the range 0 ≤ r ≤ Rr or, equivalently, In doing so, one is of course 0 ≤ ξ ≤ ξr = 0.713 ξ1. neglecting the contribution of the latter for all relevant 9 quantities: however, although the density and pressure are still significant at that point, both the polytropic and the Standard Solar Model show that up to 99.1% of the Sun's mass is located within the radiative zone. Also,it can be numerically shown that the boundary condition θ(ξ) = θ0(ξ)(1 + Acδ(ξ)) = 0 does not shift significantly from the unperturbed θ0(ξ1) = 0 case, so that no problem arises from neglecting any changes to the scale factor r0 (which, recall, relates the dimension- less coordinate ξ with the physical distance to the center r). Furthermore, the matching of the inner and outer derivatives of the metric should not be taken as a real- istic condition, but merely a consistency check for the developed model. H. Model-dependent parameters FIG. 1: Profiles for the unperturbed solution θ0 (top) and absolute perturbation ∆ = δθ0 (bottom), for 2.8 ≤ n ≤ 3.2. One proceeds with the calculation of the total mass of the Sun or, more accurately, the mass of the radiative zone: for the unperturbed case, one has (see Appendix C for the derivation of the following results) M = −4π R3 r ξr ρc0θ′ 0r → ρc0 = − M 4πθ′ 0r ξr R3 r . (67) where one defines θ′ 0(ξr) and ρc0 is the unperturbed central density (that is, obtained from M , R and the numerical results for the unperturbed solution θ0). 0r ≡ θ′ In the perturbed case, one obtains R3 r ξr ρcθ′ M = −4π "1 − Ac r θ′ ξ2 0r Z ξr 0 0r × 0 (cid:18)nδ + ξ2θn (68) 3n 8 θ′2 0 θ0 − θ′ 0 2ξ − 7 8 θn 0(cid:19) dξ# . Notice that there are two Ac-dependent contributions: one arising from θ0 and its derivatives, since including the perturbation δ would only amount to second-order terms O(A2 c ), and other involving the integral of δ. Since the total mass M does not change (only its in- terpretation as a purely gravitational mass or a sum of gravitational plus "active" components), one gets 1 − Ac r θ′ ξ2 ρc0 ρc = 0r Z ξr 0 (69) ξ2θn 0 (cid:20)nδ + 3n 8 θ′2 0 θ0 − θ′ 0 2ξ − 7 8 θn 0(cid:21) dξ . Clearly, taking Ac = 0 yields ρc = ρc0. Also, notice that ξr is not perturbed: this would reflect a change in ξ1 (since ξr = 0.713 ξ1), which has already been ruled out. Also, notice that the denominator affecting the integral is equal to ξ2 0 : thus, one can 0 (ξ2θ′ 0 θn r θ′ 0r =R ξr 0)′ = −R ξr interpret this integral as the volume averaging of the ex- pression in brackets within the integrand, with the den- sity as weighting function, that is, θn ∝ ρ. T n+1, which enables writing From the polytropic equation of state, one gets ρ ∝ Tc (cid:19)n+1 1 −(cid:18) Tc0 0r Z ξr Ac r θ′ ξ2 ξ2θn 0 = (70) 0 (cid:20)nδ + 3n 8 θ′2 0 θ0 − θ′ 0 2ξ − 7 8 θn 0(cid:21) dξ . IV. NUMERICAL RESULTS The above results allow us to construct an exclusion plot for the central temperature on the (n, Ac) parameter space, by imposing the constraint 1−Tc/Tc0 ≤ 0.06 (6%), the uncertainty of the central temperature of the Sun. Solving numerically the differential equation for δ for a varying n, and computing all relevant quantities for vary- ing Ac, and defining the absolute perturbation ∆ ≡ θ0δ, so that θ = θ0(1 + Acδ) = θ0 + Ac∆, one obtains the results shown in the Figures 1 to 7. First, notice that, from Fig. 1, the absolute perturba- tion ∆ is fairly insensitive to the value of n; clearly, given that the profile of θ0 for varying n differs more sharply as ξ → ξr (as seen in Fig. 1), a approximately constant ∆ translates into a relative perturbation δ that also exhibits this behaviour, as can be seen from Fig. 2. Furthermore, notice the indication of the divergence of the relative per- turbation δ in Fig. 2 (which does not depend on n, since θ0 → 0 and ∆ → const. yields δ → ∞, as previously discussed); as stated before, this divergence is avoided by dealing only with the radiative region r ≤ Rr. Also, as can be seen from Fig. 2, the relative perturba- tion δ peaks at δmax ≈ 1.1 (for n = 2.9, and a smaller value of δmax ≈ 0.8 for n = 3); hence, the perturbative condition Acδ ≪ 1 translates to Ac ≪ 1, leading to the 10 0.2 0 −0.2 −0.4 −0.6 −0.8 −1 δ n = 2.8 n = 3 n = 3.2 −1.2 0 0.2 0.4 r/R o 0.6 0.8 FIG. 2: Profile of the relative perturbation δ for 2.8 ≤ n ≤ 3.2. FIG. 5: Relative deviation of the central temperature Tc/Tc0 − 1 as a function of Ac and n. 1.15 1.1 1.05 1 0.95 M / M g n = 2.8 n = 3 n = 3.2 3.2 3.1 0 . 0 0 5 0 . 0 0 2 5 n 3 0.0 0 7 5 0 . 0 0 5 0 . 0 0 2 5 2.9 0 . 0 5 2 0 0.0 − 5 2 0 0.0 − 5 0 0.0 − −0.005 −0.0075 −0.01 −0.0125 0.5 1 0 0 0 0 A c 0.9 −1 −0.5 0 A c 0.5 1 1 2 50 . 0 1 5 2.8 −1 0 . 0 0 . 0 . 0 0 1 7 0 0 5 5 −0.5 FIG. 3: Fraction of gravitational to total mass as a function of Ac, for 2.8 ≤ n ≤ 3.2. FIG. 6: Contour plot for the relative deviation of the central temperature Tc/Tc0−1 as a function of Ac and n, with contour lines of step 0.1%. 1 − T / T 0 c c 0.02 0.01 0 −0.01 −0.02 −1 n = 2.8 n = 3 n = 3.2 ) g K ( M + M a g 30 x 10 2 1.5 1 −0.5 0 A c 0.5 1 0.5 0 0 A = −1 c A = 1 c A = 0 c 1 2 3 4 5 r/R o FIG. 4: Relative deviation of the central temperature Tc/Tc0 − 1 as a function of Ac, for 2.8 ≤ n ≤ 3.2. FIG. 7: Mass profiles, for n = 3 and Ac = −1, 0, 1. chosen (extreme) interval for simulation −1 ≤ Ac ≤ 1. Hence, according to Figs. 4, 5 and 6, no relative devia- tion of the central temperature above the experimentally determined level of 6% is attained. However, the values found, of the order of 1%, indicate that any future refine- ment of the experimental error of Tc could yield a direct bound on the parameter Ac. The above results show that Ac is presently uncon- strained by solar observables, aside from the perturbative condition Ac ≪ 1 → λ ≪ κ/ρc; taking κ = c4/16πG = 2.41 × 1042 kg.m/s2 and ρc = 1.622 × 105 kg/m3, this yields λ ≪ 1.48× 1037 m4/s2 or, in natural units, λ ≪ 4.24 × 1033 eV −2. However, this study assumes that the effects arising from the non-trivial matter-geometry cou- pling supersede those from the non-linear curvature term in the modified Hilbert-Einstein action (1), as expressed by the inequalities (8), (11) and (12). Having numerically determined the perturbative solution δ, one can now re- evaluate these conditions. For this, one first replaces R by its unperturbed value (with the Newtonian approach ρ ≫ p), R ≈ −ρ/2κ, since corrections would amount to second-order terms in λ. Considering the scaling laws, ρ = ρcθn(ξ) and p = pcθn+1(ξ), and the inequalities (11) and (12), some al- gebra (see Appendix D) yields, for the f1 = f1a ≡ 2κ(R − αR1−m) case, n 2Ncθ λ ≫ (cid:18) 2κ λ ≫ ρcθn(cid:19)m+1 α n + 1(cid:18) 2κ n + 1(cid:18) 2κ ρcθn(cid:19)m+1 α(1 − m)m ρcθn(cid:19)m+1 α(1 − m)m while, for the f1 = f1b ≡ 2κ(R + αR2) case, λ ≫ 2Ncθ 2Ncθ n Cn,m(ξ) , λ ≫ λ ≫ λ ≫ α 2Ncθ n n + 1 n n + 1 , α Ncθ α Ncθ , Cn,−1(ξ) , (72) where one defines the quantity m 3 , C 1.5 1 0.5 0 −0.5 −1 −1.5 0 11 θ 0 ′ θ 0 A = 1 c A = −1 c 0.2 0.4 r/R o 0.6 0.8 FIG. 8: Profiles of C3,m, for m = −1, 0, 1, superimposed on profile of θ0 and θ′ 0, with n = 3. Varying m in its domain 0 < m < 1 (for the f1 = f1a case) and m = −1 (for the f1 = f1b case); its profile is depicted in Figure 8. One concludes that Cn,m < ∼ 1.5, for both the f1 = f1a case (0 < m < 1) as well as the f1 = f1b case (m = −1). This, together with the constraint θ(ξ) > ∼ 0.1 (as can be seen from Fig. 1) and the quantity Nc ≡ pc/(ρcc2) = 1.7 × 10−6 yields, for the f1 = f1a case, λ ≫ 6.6 × 106α . (76) Using the previously discussed value, α = (1012 GeV )−2 , which arises from the Planck-scale renormalization of the theory, one gets λ ≫ 6.6 × 10−36 eV −2 = (3.9 × 108 GeV )−2 . Also, a recent paper has reported a relation between the coefficient of the quadratic term in f1 and the PPN parameter γ [26], (77) , (71) , λ ≫ 2.8 × 1043(8.48 × 1036)m(1 − m)mα eV −2(1+m) while, for the f1 = f1b case, , (75) . (73) α = . (78) Cn,m(ξ) ≡(cid:12)(cid:12)(cid:12)(cid:12) (1 + nm)θ′2(ξ) − θθ′′(ξ) nθ′2(ξ) + θθ′′(ξ) (cid:12)(cid:12)(cid:12)(cid:12) 1 2κs(cid:12)(cid:12)(cid:12)(cid:12) 1 − γ 2γ − 1(cid:12)(cid:12)(cid:12)(cid:12) One can plot the function Cn,m(ξ) by evaluating θ by the unperturbed solution θ0, using Eq. (49) and taking n ≈ 3. One obtains Hence, the current experimental constraint γ−1 = (2.1± 2.3) × 10−5 [27] yields Cn,m(ξ) ≈(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) θ′ 0h(1 + 3m)θ′ 0 + 2 θ′ 0(cid:16)3θ′ 0 + 2 ξ θ0i + θ4 0 ξ θ0(cid:17) + θ4 0 . (74) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) α ≤ 0.17M −2 P l =(cid:0)3.0 × 1019 GeV(cid:1)−2 where MP l = p¯hc/G = 1.2 × 1019 GeV is the Planck mass. This yields , (79) (80) λ ≫ 7.4 × 10−51 eV −2 = (1.2 × 1016 GeV )−2 . Clearly, the f1 = f1b case does not impose any sig- nificant bound on λ, with the above value laying many orders of magnitude below the upper bound, λ ≪ 4.24 × 1033 eV −2, arising from the perturbative treat- ment. Notice that the above inequalities (71), (72), (75), (76), (77) and (80) are not to be considered as restrictions on the parameters, but as conditions for the validity of the regime where the effects of the matter-geometry coupling supersede those of the the non-linear curvature term in Eq. (1). V. DISCUSSION In this work we have examined a model where the usual Einstein-Hilbert action is modified, not only by allowing non-linear curvature terms to appear, but by also en- abling an explicit, non-minimum coupling between the curvature and the Lagrangian of matter fields. We first assume that this matter-geometry coupling is linear in curvature and its effect allows the non-linear curvature terms to be neglected -- an assertion that is qualified in the end of the study, for two relevant non-linear f (R) models. We then proceed and perform the necessary calcula- tions to ascertain the effect of the latter in the hydro- static equilibrium of an n ≈ 3 polytrope such as the Sun. We assume a perturbative regime to the usual Tolman- Oppenheimmer-Volkoff equation, and take into consider- ation the validity of the Newtonian regime in this mod- ified theory, as well as the redefinition of relevant quan- tities, which are computed numerically and compared with Solar observables. This goal is achieved through the use of the (non-relativistic) polytropic equation of state p = Kρ(n+1)/n : as stated before, this is a very simplistic description of the complex behaviour of solar matter, and has been superseded by much more elaborate equations of state; we use it in our study so to better illustrate the effects of the matter-curvature coupling on an eas- ily understandable, closed model. However, we should remark that this restrictive treatment, although advan- tageous from the theoretical point of view, might conceal some of the more intricate phenomenology found in stel- lar systems, which could affect our results. Clearly, a subsequent study should consider a more realistic solar model. B The results allow us to conclude that no strong con- straints on the matter-geometry coupling from the com- parison between the model's predictions and current ex- perimental sensitivity, aside from the perturbative ap- proach λf2(R) = λR ≈ (λ/κ)ρc ≪ 1, which yields λ ≪ 4.24 × 1033 eV −2. However, the numerically ob- tained results show that a slight increase in accuracy 12 would allow an upper bound to be placed on λ; this close- ness between the prediction of the perturbative model and experiment also seems to validate the latter. APPENDIX A Given the line element (16), one first writes g00R00 − grrRrr = e−σ ν′ + σ′ r . (A1) Resorting to Eq. (15), one gets 2κ[2 + a(ρ − 3p)](g00R00 − grrRrr) = 2(1 − ap)(g00T00 − grrTrr) + 4aκ(g00 κ[2 + a(ρ − 3p)]e−σ ν′ + σ′ = ρ + p − ap(ρ + p) + 4aκ(e−ν 00 + e−σ r 00 − grr (A2) rr)p → rr)p , since the terms depending on the metric cancel out (that is, g00g00 − grrgrr = 0). Next, compute (A3) Rθθ r2 = 2κ[2 + a(ρ − 3p)] gθθ r2 + 2(1 − ap) (3p − ρ) 2κ[2 + a(ρ − 3p)](cid:20) 1 ρ − 3p + 2(1 − ap)p + aκ(cid:18) 4 κ[2 + a(ρ − 3p)](cid:20) 2 ρ − p − 2ap2 + aκ(cid:18) 4 1 Tθθ r2 + aκ r2 (4 θθ − gθθ )p → r2 + e−σ(cid:18) σ′ − ν′ 2r − θθ + (cid:19) p → r2(cid:19)(cid:21) = r2(cid:19)(cid:21) = − r2 2 r2 + e−σ(cid:18) σ′ − ν′ θθ + (cid:19) p . r2 r Adding Eqs. (A2) to (A3), one obtains Eq. (17) κ[2 + a(ρ − 3p)](cid:20) 1 − (ρ + 3p) + ap 2 r2 + e−σ(cid:18) σ′ r − 2 (cid:18)5e−ν aκ (A4) θθ(cid:19) p , 2 r2 1 r2(cid:19)(cid:21) = ρ 00 + 3e−σ rr + with ≡ e−ν Eq. (18), one gets Eq. (26) 00 − e−σ rr − 2 θθ/r2. By subtracting Eqs. (A2) from (A3), and substituting κ [2 + a (ρ − 3p)](cid:20)(cid:18)1 − 2 (cid:18)3e−ν (p − ρ) + ap 2 aκ + 2Gm r (cid:19) ν′ r − 00 + 5e−σ 2Gm r3 (cid:21) = p (A5) θθ(cid:19) p . rr − 2 r2 APPENDIX B In order to obtain an approximation to the solution of the Lane-Emden equation (48) in the vicinity of ξ = 0, one writes θ(ξ) ≈ 1−Aξ2+Bξ4. Inserting this in Eq. (48) yields, after some algebra and expanding up to fourth order in ξ, 13 Ac 4(n + 1) (cid:20)20 + (cid:20)−1 + Ac 2 + 5A 4 (A [4Nc(n + 1)(2n + 1) + An(13 + 21n)] + 10B(13 + 21n))(cid:21) ξ4 = (cid:21) ξ2 +(cid:20)An − Ac(cid:18)An(cid:20)1 + B(cid:19)(cid:21) ξ4 . A(cid:21) + 11 4 11 2 (B1) (C4) Equating second and fourth-order terms, one gets A ≈ 1 6(cid:18)1 − Ac 13 + 25n 12(n + 1)(cid:19) , and B ≈ n 120(cid:18)1 − Ac 11 60 39 + 59n n + 1 (cid:19) . APPENDIX C (B2) (B3) In the unperturbed case, the total mass M is purely gravitational, and hence defined as usual: r2ρ dr = 4πr3 0ρc0Z ξr 0)′ dξ = −4πr3 0 (ξ2θ′ ξ2θn 0 dξ = (C1) 0ρc0ξ2 r (θ′ 0)ξ=ξr = 0 M = 4πZ Rr 0ρc0Z ξr −4πr3 R3 r ξr 0r ≡ θ′ −4π ρc0θ′ 0 0r , defining θ′ density ρc0, a model dependent parameter, is given by 0(ξr); hence, the (unperturbed) central 0 0 0 G 4π ρ2 c m′ n + 1 4πr3 R3 r ξ3 r e dr = r2ρ dr + ξ2θn dξ + (cid:18)4πr2 M =Z Rr 4πZ Rr 0ρcZ ξr ρcZ ξr pc(cid:19) AcNcr0Z ξr ρcZ ξr 0ρcAcZ ξr ρc Z ξr ρc −ξ2 R3 r ξ3 r R3 r ξ3 r R3 r ξ3 r 4πr3 ξ2θn ξ2θn r θ′ 4π 4π 4π 0 0 0 0 0 R3 r ξr −4π ρcθ′ 0r 1 − ξ2θn 0 Θ dr = AcNcZ Rr 0 n + 1 G AcNcr0Z ξr 0 ξ2θn 0 Θ dξ = ξ2 (θ0(1 + Acδ))n dξ + ξ2θn 0 Θ dξ ≈ 0 0 (1 + nAcδ) dξ + 0 Θ dξ = 0 ξ2θn 0 dξ + AcZ ξr 0r + AcZ ξr 0r Z ξr Ac r θ′ ξ2 ξ2θn 0 0 ξ2θn 0 (nδ + Θ) dξ! = 0 (nδ + Θ) dξ! = 0 (nδ + Θ) dξ! . ξ2θn ρc0 = − M 4πθ′ 0r ξr R3 r . (C2) In the perturbed case, one first uses Eq. (62) to define, for clarity, Notice that there are two Ac-dependent contributions: one arising from θ0 and its derivatives (embodied in Θ), and other involving the integral of δ. Θ(ξ) = − 3n 8 θ′2 0 θ0 + θ′ 0 2ξ + 7 8 θn 0 . (C3) APPENDIX D In the above, θ0 is used instead of θ, since Θ is coupled to Ac, so that including the perturbation δ would only amount to second-order terms O(A2 c ). Using the defini- tions Nc ≡ pc/ρc and r2 0 ≡ (n + 1)pc/4πGρ2 c, one gets By considering the scaling laws ρ = ρcθn(ξ) and p = pcθn+1(ξ) and using Eq. (11) one gets, for the f1 = f1a case, where the prime denotes denotes differentiation with re- spect to ξ (the factors r0 arising from changing derivation with respect to r to derivation with respect to ξ = r/r0 cancel out). , (D1) 14 2Ncθn(1+m)+1 α = , n ρmp ρ′(r) λ ≫ 2mακ1+m ρc(cid:19)m+1 =(cid:18) 2κ λ ≫ 2m(1 − m)mακ1+m(cid:12)(cid:12)(cid:12)(cid:12) ρ1+mp′(cid:12)(cid:12)(cid:12)(cid:12) ρc(cid:19)m+1 α(1 − m)m n + 1(cid:18) 2κ 2m(1 − m)mακ1+m (cid:12)(cid:12)(cid:12)(cid:12) λ ≫ ρc(cid:19)m+1 α(1 − m)m 2Ncθn(1+m)+1 (cid:12)(cid:12)(cid:12)(cid:12) (cid:18) 2κ 2Ncθn(1+m)+1 ρ1+mp′′ ρ′2(r) ρ′′(r) − (1 + m) (1 + nm)θ′2(ξ) − θθ′′(ξ) nθ′2(ξ) + θθ′′(ξ) ρ (cid:12)(cid:12)(cid:12)(cid:12) = (cid:12)(cid:12)(cid:12)(cid:12) , while, for the f1 = f1b case, Eq. (12) yields λ ≫ λ ≫ λ ≫ α 2Ncθ n , α n + 1 2Ncθ n n + 1 α 2Ncθ (cid:12)(cid:12)(cid:12)(cid:12) , (1 − n)θ′2(ξ) − θθ′′(ξ) nθ′2(ξ) + θθ′′(ξ) (D2) , (cid:12)(cid:12)(cid:12)(cid:12) ACKNOWLEDGMENTS The authors would like to thank F. Lobo for fruit- is sponsored by ful discussions. The work of J.P. the FCT under the grant BP D23287/2005. O.B. ac- knowledges the partial support of the FCT project P OCI/F IS/56093/2004. [1] S. Carroll, V. Duvvuri, M. Trodden and M. Turner, Phys. Rev. D 70, 043528 (2004); S. Capozziello, S. Nojiri, S.D. Odintsov and A. Troisi, Phys. Lett. B 639, 135 (2006); S. Nojiri and S.D. Odintsov, Phys. Rev. D 74, 086005 (2006), Int. J. Geom. Meth. Mod. Phys. 4, 115 (2007). [2] L. Amendola, R. Gannouji, D. Polarski and S. Tsujikawa, Phys. Rev. D 75, 083504 (2007). [3] A. Coley, N. Pelavas and R. Zalaletdinov, Phys. Rev. Lett. 95, 151102 (2005); S. Rasanen, Int. J. Mod. Phys. D 15. 2141 (2006); T. Buchert, gr-qc/0612166. [4] O. Bertolami and R. Rosenfeld, hep-ph/0708.1784. [5] M. Bento, O. Bertolami and A. Sen, Phys. Rev. D 66, 043507 (2002); N. Bilic, G. Tupper and R. Viollier, Phys. Lett. B 535, 17 (2002); V. Gorini, A. Kamenshchik and U. Moschella, Phys. Rev. D 67, 063509 (2003). [6] R. Maartens, Living Rev. Rel. 7, 7 (2004). [7] S. Capozziello, V. Cardone and A. Troisi, Phys. Rev. D 71, 043503 (2005); G. Allemandi, A. Borowiec and M. Francaviglia, Phys. Rev. D 70, 103503 (2004). [8] A. Starobinsky, Phys. Lett. B 91, 99 (1980). [9] S. Capozziello, A. Stabile and A. Troisi, gr-qc/0708.0723. [10] T. Sotiriou, Class. Quantum Gravity 23, 5117 (2006); N. Poplawski, Class. Quantum Gravity 23, 2011 (2006); T. Sotiriou and S. Liberati, Annals Phys. 322, 935 (2007). [11] S. Capozziello, V. Cardone, S. Carloni and A. Troisi, Phys. Lett. A 326, 292 (2004); S. Capozziello, V. Car- done and A. Troisi, Mon. Not. R. Ast. Soc. 375, 1423 (2007); J. Mbelek, Astron. and Astrophys. 424, 761 (2004). [14] S. Mukohyama and L. Randall, Phys. Rev. Lett. 92, 211302 (2004); A. Dolgov and M. Kawasaki, astro- ph/0307442; astro-ph/0310822. [15] L. Amendola, Phys. Rev. D 62, 043511 (2000); Phys. Rev. Lett. 86, 196 (2001); L. Amendola and D. Tocchini- Valentini, Phys. Rev. D 64, 043509 (2001), Phys. Rev. D 66, 043528 (2002); A Billyard and A. Coley, Phys. Rev. D 61, 083503 (2000); D. Holden and D. Wands, Phys. Rev. D 61, 043506 (2000); O. Bertolami and P. Martins, Phys. Rev. D 61, 064007 (2000); L. Chimento, A. Jakubi, D. Pavon and W. Zimdahl, Phys. Rev. D 67, 083513 (2003); L. Chimento and D. Pavon, Phys. Rev. D 73, 063511 (2006); C. Boehmer, T. Harko and F. Lobo, gr-qc/0709.0046. [16] V. Faraoni, Cosmology in Scalar-Tensor Gravity (Kluwer Academis Publishers, 2004); O. Bertolami. J. P´aramos and S. Turyshev, gr-qc/0602016; O. Bertolami, F. Pedro and M. Le Delliou, astro-ph/0703462; astro- ph/0705.3118. [17] O. Bertolami and J. P´aramos, Phys. Rev. D 71, 023521 (2005). [18] O. Bertolami and J. P´aramos, Phys. Rev. D 72, 123512 (2005). [19] T. Multamaki and I. Vilja, astro-ph/0612775; K. Kain- ulainen, J. Piilonen, V. Reijonen and D. Sunhede, Phys. Rev. D 76, 024020 (2007); A. Bustelo and D. Barraco, Class. Quantum Gravity 24, 2333 (2007); E. Barausse, T. Sotiriou and J. Miller, gr-qc/0703132; S. Capozziello, A. Stabile and A. Troisi, gr-qc/0709.0891. [12] O. Bertolami, C. Boehmer, T. Harko and F. Lobo, Phys. [20] B. Schutz, Phys. Rev. D 2, 2762 (1970); J. Brown, Class. Rev. D 75, 104016 (2007). Quantum Gravity 10, 1579 (1993). [13] S. Nojiri and S. Odintsov,Phys. Lett. B 599, 137 (2004); hep-th/0412030; G. Allemandi, A. Borowiec, M. Francav- iglia, and S. Odintsov, Phys. Rev. D 72, 063505 (2005). [21] "The Fundamentals of Stellar Astrophysics", G. Collins (W.H. Freeman and Co., 1989); J.N. Bahcall, Phys. Rev. D33 47 (2000); "Textbook of Astronomy and As- trophysics with Elements of Cosmology", V. Bhatia (Narosa Publishing House, 2001); "Theoretical Astro- physics: Stars and Stellar Systems", T. Padmanabhan (Cambridge University Press, 2001). [22] J. Bahcall, A. Serenelli and S. Basu, Ap. J. Supp. 165, 400 (2006); R. Trampedach, W. Dappen and V. Baturin, Ap. J. 646, 560 (2006). [23] R. Tooper, Ap. J. 140, 434 (1964). 15 [24] N. Stergioulas, Living Rev. Rel., 6 3 (2003). [25] R. Tooper, Ap. J. 142, 1541 (1965). [26] S. Capozziello, A. Stabile and A. Troisi, Mod. Phys. Lett. A 21, 2291 (2006). [27] B. Bertotti, L. Iess and P. Tortora, Nature 425, 374 (2003).
astro-ph/0701255
1
0701
2007-01-09T16:31:45
Study on polarization of high-energy photons from the Crab pulsar
[ "astro-ph" ]
We investigate polarization of high-energy emissions from the Crab pulsar in the frame work of the outer gap accelerator. The recent version of the outer gap, which extends from inside the null charge surface to the light cylinder, is used for examining the light curve, the spectrum and the polarization characteristics, simultaneously. The polarization position angle curve and the polarization degree are calculated to compare with the Crab optical data. We show that the outer gap model explains the general features of the observed light curve, the spectrum and the polarization by taking into account the emissions from inside of the null charge surface and from tertiary pairs, which were produced by the high-energy photons from the secondary pairs. For the Crab pulsar, the polarization position angle curve indicates that the viewing angle of the observer measured from the rotational axis is greater than 90 degrees.
astro-ph
astro-ph
Proceedings of the 363. WE-Heraeus Seminar on: "Neutron Stars and Pulsars" (Posters and contributed talks) Physikzentrum Bad Honnef, Germany, May. 14 − 19, 2006, eds. W. Becker, H.H. Huang, MPE Report 291, pp. 120-123 7 0 0 2 n a J 9 1 v 5 5 2 1 0 7 0 / h p - o r t s a : v i X r a Study on polarization of high-energy photons from the Crab pulsar J. Takata1, H.-K. Chang2, and K.S. Cheng3 1 ASIAA/National Tsing Hua University - TIARA, PO Box 23-141, Taipei, Taiwan 2 Department of Physics and Institute of Astronomy, National Tsing Hua University, Hsinchu 30013, Taiwan 3 Department of Physics, University of Hong Kong, Pokfuam Road, Hong Kong, China Abstract. We investigate polarization of high-energy emissions from the Crab pulsar in the frame work of the outer gap accelerator. The recent version of the outer gap, which extends from inside the null charge surface to the light cylinder, is used for examining the light curve, the spectrum and the polarization characteristics, simultane- ously. The polarization position angle curve and the polar- ization degree are calculated to compare with the Crab op- tical data. We show that the outer gap model explains the general features of the observed light curve, the spectrum and the polarization by taking into account the emissions from inside of the null charge surface and from tertiary pairs, which were produced by the high-energy photons from the secondary pairs. For the Crab pulsar, the po- larization position angle curve indicates that the viewing angle of the observer measured from the rotational axis is greater than 90◦. 1. Introduction Young pulsars such as the Crab pulsar are strong γ-ray sources. The EGRET instrument revealed that the light curve with double peaks in a period and the spectrum extending to above GeV are typical features of the high- energy emissions from the γ-ray pulsars. Although these data have constrained proposed models, the origin of the γ-ray emission is not yet conclusive. On important reason is that various models have successfully explained the fea- tures of the observed spectra and/or light curves. For ex- ample, the polar cap model (Daugherty & Harding 1996), the caustic model (Dyks et al. 2004) and the outer gap model (Cheng et al. 2000, hereafter CRZ00), all expect the main features of the observed light curve. So, we can- not discriminate the three different models using the light curve. Furthermore, both polar cap and outer gap models have explained the observed γ-ray spectrum (Daugherty & Harding 1996; Romani 1996). Polarization measurement will play an important role to discriminate the various models, because it increases observed parameters, namely, polarization degree (p.d.) and position angle (p.a.) swing. So far, only the optical polarization data for the Crab pulsar is available (Kan- bach et al. 2005) in high energy bands. For the Crab pul- sar, the spectrum is continuously extending from optical to γ-ray bands. In addition, their pulse positions are con- sidered so well that the optical emission mechanism is re- lated to higher energy emission mechanisms. In the future, the next generation Compton telescope will probably be able to measure polarization characteristics in MeV bands. These data will be useful for discriminating the different models. In this paper, we examine the optical polarization char- acteristics of the Crab pulsar with the light curve and the spectrum in frame works of the outer gap model. CRZ00 has calculated the synchrotron self-inverse Compton scat- tering process of the secondary pairs produced outside the outer gap and has explained the Crab spectrum from X- ray to γ-ray bands. In CRZ00, however, the outer-wing and the off-pulse emissions of the Crab pulsars cannot be reproduced, because the traditional outer gap geom- etry, which extends from the null charge surface of the Goldreich-Julian charge density to the light cylinder, is assumed. Furthermore, the spectrum in the optical band was not considered. In this paper, on these grounds, we modify the CRZ00 geometrical model into a more realis- tic model, following recent 2-D electrodynamical studies (Takata et al. 2004; Hirotani 2006), and calculate the light curve, the spectrum and the polarization characteristics of the Crab pulsar. 2. Calculation method The outline of the outer gap model for the Crab pul- sar is as follows. The charge particles are accelerated by the electric field parallel to the magnetic field lines in so called gap, where the charge density is different from the Goldreich-Julina charge density. The high energy particles accelerated in the gap emit the γ-ray photons (called pri- mary photons) via the curvature radiation process. For the Crab pulsar, most of the primary photons escaping from the outer gap will convert into secondary pairs outside the gap, where the accelerating electric field vanishes, by col- liding with synchrotron X-rays emitted by the secondary J. Takata, H.-K. Chang, & K.S. Cheng: Study on polarization of high-energy photons from the Crab pulsar 121 pairs. The secondary pairs emit optical - MeV photons via the synchrotron process and photons above MeV with the inverse Compton process. The high-energy photons emit- ted by the secondary pairs may convert into tertiary pairs at higher altitude by colliding with the soft X-rays from the stellar surface. The tertiary pairs emit the optical-UV photons via the synchrotron process. This secondary and tertiary photons appear as the observed radiations from the Crab pulsar 2.1. Outer gap We consider the outer gap geometry that is extending from inside null charge surface to near the light cylinder (Takata et al 2004; Hirotani 2006). Because the Crab pulsar has a thin gap, we describe the accelerating electric field (Cheng et al. 1986) with E(r) = ΩB(r)f 2(r)R2 lc cs(r) , (1) where f (r) is the local gap thickness in units of the light radius, Rlc = c/Ω, and s(r) is the curvature radius of the magnetic field line. The typical fractional size of the outer gap is given by f (Rlc/2) ∼ 5.5B−4/7P 26/21 (∼ 0.04 for the Crab pulsar) and the local fractional size is estimated by f (r) ∼ f (Rlc/2)(2r/Rlc)1.5 (Zhang & Cheng 1997). The electric field described by equation (1) can be adopted for the acceleration beyond the null charge sur- face. Inside null charge surface, the electric field rapidly decreases because of the screening effects of the pairs. To simulate the accelerating electric field inside null charge surface, we assume the following field, E(r) = En (r/ri)2 − 1 (rn/ri)2 − 1 , ri ≤ r ≤ rn, (2) where En is the strength of the electric field at the null charge surface and rn and ri are the radial distances to the null charge surface and the inner boundary of the gap, respectively. The local Lorentz factor of the accel- erated particles in the outer gap is described by Γe(r) = [3s2(r)E/2e]1/4 with assuming the force balance between the acceleration and the curvature radiation back reaction. We assume that the outer gap extends around the whole polar cap. In the calculation, we constrain the boundaries of the axial distance and radial distance for the emission regions with ρmax = 0.9Rlc and r = Rlc, respectively. 2.2. Synchrotron radiation from the pairs We calculate the synchrotron radiation at each radiating point following CRZ00. The photon spectrum of the syn- chrotron radiation is (CRZ00) Fsyn(Eγ, r) = 31/2e3B(r) sin θp(r) mc2hEγ × Z (cid:20) dne(r) dEe (cid:21) F (x)dEedVrad, (3) e is the pitch angle of the particle, F (x) = xR ∞ where x = Eγ/Esyn, Esyn is the typical photon energy, θp x K5/3(y)dy, where K5/3 is the modified Bessel function of order 5/3, dne/dEe ∝ E−2 is the distribution of the pairs and dVrad is the volume element of the radiation region considered. The pitch angle of the secondary pairs is estimated from sin θp(Rlc) ∼ λ/s(Rlc), where λ is the mean free path of the pair-creation between the primary γ-rays and the non-thermal X-rays from the secondary pairs. For the Crab pulsar, the mean free path becomes λ ∼ (nX σγγ)−1 ∼ 107cm, where we used the typical num- ber density nX ∼ LX(< EX >)/δΩR2 lcc < EX >∼ 8 × 1017 cm−3, the luminosity LX ∼ 1035erg/s, the typ- ical soft-photon energy for the pair-creation, < EX >∼ (2mec2)2/10 GeV ∼ 100 eV, and σγγ ∼ σT /3. There- fore the pitch angle of the secondary pairs at the light cylinder is estimated as sin θp ∼ λ/s(Rlc) ∼ 0.06, and the local pitch angle is calculated from sin θp(r) = sin θp(Rlc)(r/Rlc)1/2. Some high-energy photons emitted by the inverse Compton process of the secondary pairs may convert into tertiary pairs at higher altitude by colliding with thermal X-ray photons from the star. In this paper, we assume that the maximum energy of and the local number den- sity of the tertiary pairs are smaller than about 10% of those of the secondary pairs. Because the pitch angle of the pairs increases with altitude, we use sin θp = 0.1 for the pitch angle of the tertiary pairs. In fact, the results are not sensitive to the pitch angle of the tertiary pairs. 2.3. Stokes parameters For a high Lorentz factor, we can anticipate that the emis- sion direction of the particles coincides with the direction of the particle's velocity. In the inertial observer frame, the particle motion may be described by n = β0 cos θpb + β0 sin θpb⊥ + βcoeφ, (4) where the first term in the right hand side represents the particle motion along the field line, for which we use the rotating dipole field, and b is the unit vector of the mag- netic field line. The second term in equation (4) repre- sents gyration motion around the magnetic field line and b⊥ ≡ cos δφk + sin δφk × b is the unit vector perpendicu- lar to the magnetic field line, where δφ refers the phase of gyration motion and k = (b ·∇)b/(b·∇)b is the unit vec- tor of the curvature of the magnetic field lines. The third term is co-rotation motion with the star, βco = ρΩ/c. The emission direction of equation (4) is described in terms 122 J. Takata, H.-K. Chang, & K.S. Cheng: Study on polarization of high-energy photons from the Crab pulsar of the viewing angle measured from the rotational axis, ξ = cos−1 nz, and the rotation phase, Φ = −Φn − r · n, where nz is the component of the emission direction par- allel to the rotational axis, Φn is the azimuthal angle of the emission direction and r is the emitting location in units of the light radius. Because the particles distribute on the gyration phase δφ, the emitted beam at each point must become cone like shape with opening angle θp(r). We calculate the ra- diations from the particles for all of the gyration phase δφ = 2πi/n (i = 1, · · · , n − 1). We assume that the radiation at each point linearly polarizes with degree of Πsyn = (p + 1)/(p + 7/3), where p is the power law index of the particle distribution, and circular polarization is zero, that is, V = 0 in terms of the Stokes parameters. The direction of the electric vec- tor of the electro-magnetic wave toward the observer is parallel to the projected direction of the acceleration of the particle on the sky, that is, Eem ∝ a − (n · a)n (Blaskiewicz et al. 1991). In the present case, the ac- celeration with equation (4) is approximately written by a ∼ β0ωB sin θp(− sin δφk + cos δφk × b). We define the position angle χi to be angle between the electric field Eem and the projected rotational axis on the sky. The Stokes parameters Qi and U i at each ra- diation is represented by Qi = ΠsynI i cos 2χi and U i = ΠsynI i sin 2χi. After collecting the photons from the pos- sible points for each rotation phase Φ and a viewing angle ξ, the expected p.d. and p.a. are, respectively, obtained from P (ξ, Φ) = ΠsynpQ2(ξ, Φ) + U 2(ξ, Φ)/I(ξ, Φ) and χ(ξ, Φ) = 0.5atan [U (ξ, Φ)/Q(ξ, Φ)]. The inclination angle α and the viewing angles ξ mea- sured from the rotational axis are the model parameters. The radial distance ri of the inner boundary in equation (2) is also a model parameter, because the distance ri is determined by the current through the gap (Takata et al. 2004). Because the last-open line must be modified from the traditional magnetic surface, which is tangent to the light cylinder for the vacuum case, by the plasma effects (Romani 1996), the altitude of the upper surface of the outer gap, above which the pairs are created and emit the synchrotron photons, is used as a model parameter using a fractional polar angle a = θu/θlc, where θu is the po- lar angle of the footpoints of the magnetic field lines of the gap upper surface and θlc is the polar angle of the field lines which are tangent to the light cylinder for the vacuum case. 3. Results Figure 1 compares the polarization characteristics at 1 eV predicted by the traditional (left column) and the present (right column) outer gap models. The traditional model considers the emissions from the secondary pairs with the outer gap starting from the null charge surface of the Goldreich-Julian charge density, ri = rn. The present 1 0.8 0.6 0.4 0.2 0 80 60 40 20 0 -20 -40 -60 -80 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 y t i s n e t n I ] . g e d [ . A P . ] % 0 0 1 x [ . D P . secondary pairs, ri=rn secondary+tertiary pairs, ri=0.68rn 1eV secondary pairs, ri=rn secondary+tertiary pairs, ri=0.68rn secondary pairs, ri=rn secondary+tertiary pairs, ri=0.68rn -0.4 -0.2 0 0.2 Φ 0.4 0.6 0.8 1 -0.4 -0.2 0 0.2 Φ 0.4 0.6 0.8 1 Fig. 1. Polarization characteristics for the traditional model (left column) and the present model (right column). The upper, middle and lower panels in each column show, respectively, the light curve, the position angle and the polarization degree for α = 40◦ and a = 0.93. model takes into account the emissions from inside null charge surface and the tertiary pairs. We assume the outer gap starts from the radial distance of 68% of the distance to null charge surface, ri = 0.68rn. The model parameters are α = 40◦, a = 0.93 and ξ ∼ 101◦, where the viewing angle is chosen so that the predicted phase separation be- tween the two peaks is consistent with the observed value δΦ ∼ 0.4 phase. From the pulse profiles, we find that the traditional model can not produce the outer-wing and the off-pulse emissions. On the other hand, the present model produces the outer-wing and the off-pulse emissions with the emis- sions from inside of the null charge surface. As seen in the polarization characteristics by the tradi- tional model, we find that the secondary emissions beyond the null charge surface make the polarization characteris- tics such that the polarization degree takes a lower value at the bridge phase and a larger value near the peaks. In the synchrotron case, the cone like beam is radiated at each point, and an overlap of the radiations from the different particles on the gyration phase causes the depolarization. For the viewing angle ξ ∼ 101◦, the radiations from all gyration phases contribute to the observed radiation at the bridge phase, because the line of sight passes through middle part of the emission regions. In such a case, the depolarization is strong, and as a result, the emerging ra- diation from the secondary pairs polarizes with a very low p.d. (< 10%). Near the peaks, the radiations from the some range of gyration phases are not observed because the line of sight passes through the edge of the emission re- gions at peaks. In such a case, the depolarization is weaker and the emerging radiation highly polarizes. In the polarization angle swing in the present model, the large swings appear at the both peaks, and the differ- ence of the position angle between the off-pulse and the bridge phases is about 90 degree. We can see the effects of the tertiary pairs on the polarization characteristics at the bridge phase. By comparing the p.d. at the bridge J. Takata, H.-K. Chang, & K.S. Cheng: Study on polarization of high-energy photons from the Crab pulsar 123 y ] t s t i i n u s . n b r e a [ t n x u I l f 5 4 3 2 1 0 ) o ( . A . P ] ° [ ψ 200 150 100 50 50 40 ] ) % ( % . P D . P [ 30 20 10 0 -0.5 total pulsed 0.0 Φ 0.5 -0.5 0.0 0.5 1.0 Φ 10 ] s COMPTEL BeppoSAX PDS BeppoSAX MECS BeppoSAX LECS OSSE GRIS 2 m c / g r e 0 1 - 0 1 [ x u F x l 2 E 1 0.1 10-6 10-5 10-4 10-3 10-2 Energy [MeV] 10-1 100 101 Fig. 2. Optical polarization for the Crab pulsar. Left:Polarization characteristics for the total emissions. Right:Polarization characteristics of the emissions after subtraction of the DC level (Kanbach et al. 2005). The figures was transcribed from Dyks et al. (2004). Fig. 3. The optical-X ray spectrum for the Crab pulsar. The calculation is for α = 40◦, a = 0.93, ri = 0.68 and ξ ∼ 101◦. The X-ray data are taken from Kuiper et al. (2002) and reference therein, and the optical data from Sollerman et al. (2000). phase, we find that tertiary pairs produce the radiations with ∼ 10% of the p.d. at the bridge phase. Figure 2 summarizes the Crab optical data. Left col- umn and the right column show, respectively, the Crab optical data for the total emissions and for the emissions after subtraction of the DC level, which has the constant intensity at the level of 1.24% of the main pulse intensity. In the total emissions (left column), the impressive polar- ization feature is that the off-pulse and bridge phases have the fixed value of the p.a. These polarization features of the observation are not predicted by the present model. The present model are more consistent with the Crab op- tical data after the subtraction of the DC level. Especially, the model reproduces the most striking feature in the ob- served p.a. that the large swing at both peaks, and the observed low p.d. at bridge phase ∼ 10%. Also, the pat- tern of the p.d. are reproduced by the present model. Figure 3 compares the model spectrum with the Crab data in optical-MeV bands. The model parameters are same with that in the right column in Figure 2. In this case, we assume that the pairs escape from the light cylin- der with the Lorentz factor Γ ∼ 20 due to the synchrotron cooling effect, which makes a spectral break around 10 eV in Figure 3. The model spectrum also explains the general features of the data. The outer gap model can explain the general features of the observed light curve, the spectrum and the polarization characteristics in optical band for the Crab pulsar, simultaneously. It is worth to note that we can distinguish the two viewing angles mutually symmetric with respect to the rotational equator using the polarization swing. For such symmetric viewing angles, the light curves, the spectra and the p.d. curves are identical. However, the p.a. curves are mirror symmetry with respect to the rotational equa- tor because of the difference directions of the projected magnetic field on the sky. The pattern of the position an- gle for the viewing angle smaller than 90◦ swings to op- posite directions from the Crab data at the both peaks. Therefore, the present model predicts that the viewing angle larger than 90◦ are preferred for the Crab pulsar. 4. Conclusions We have considered the light curve, the spectrum and the polarization characteristics for the Crab pulsar predicted by the outer gap model which takes into account the emis- sions from inside the null charge surface and from the ter- tiary pairs. We have shown that the expected polarization characteristics are consistent with the Crab optical data after subtraction of the DC level. The outer gap model explains the spectrum, light curve and the polarization characteristics, simultaneously. Acknowledgements. This work was supported by the Theoret- ical Institute for Advanced Research in Astrophysics (TIARA) operated under Academia Sinica and the National Science Council Excellence Projects program in Taiwan administered through grant number NSC 94-2112-M-007-002, NSC 94- 2752-M-007-002-PAE and NSC 95-2752-M-007-001-PAE. And the author, KSC, was supported by a RGC grant number HKU7015/05P. References Blaskiewicz, M. et al. 1991, ApJ, 370, 643 Cheng, K.S., Ho, C. & Ruderman, M. 1986, ApJ, 300, 500 Cheng, K.S., Ruderman, M. & Zhang, L. 2000, ApJ, 537, 964 Daugherty, J.K. & Harding, A.K. 1996, ApJ, 458, 278 Dyks J., Harding A.K. & Rudak B. 2004, ApJ, 606, 1125 Hirotani, K. 2006, Mod. Phys. Lett. A 21, 1319 Kanbach, G. et al. 2005, AIP Conference Proceeding, 801, 306 Kuiper, L., et al. 2001, A&A, 378, 918 Romani R.W. 1996, ApJ, 470, 469 Sollerman, J. et al. 2000, ApJ, 537, 86 Takata, J., Shibata, S. & Hirotani K. 2004, MNRAS, 354, 1120 Takata, J. et al. 2006, MNRAS, 366, 1310 Zhang, L. & Cheng K.S. 1997, ApJ, 487. 370
astro-ph/9805144
1
9805
1998-05-12T10:48:02
Element Ratios and the Formation of the Stellar Halo
[ "astro-ph" ]
It is well-established that the vast majority of metal-poor Galactic halo stars shows evidence for enrichment by solely massive stars. Recent observations have identified more varied behavior in the pattern of elemental abundances measured for metal-rich, [Fe/H]=-1, halo stars, with both high and low values of the ratio of [alpha/Fe]. The low values are most naturally due to the incorporation of iron from Type Ia supernovae. These `low-alpha' halo stars have been interpreted as being accreted from dwarf galaxies. However, these stars are on very high-energy radial orbits, that plunge into the Galactic center, from the outer regions of the stellar halo. We demonstrate here that known dwarf galaxies could not reproduce the observations. Rather, the observations are consistent with fragmentation of a gaseous proto-halo into transient star-forming regions, some of which are sufficiently dense to survive close to the Galactic center, and self-enrich to relatively high metallicities. Those that probe the outer halo sustain star formation long enough to incorporate the ejecta from Type Ia supernovae. The values of the `Type II plateau' in the element ratios for disk stars and for halo stars are equal. This implies that the (massive) stars that enriched the early disk and halo had the same IMF. However, there is a discontinuity in the chemical enrichment history between the halo and disk, as probed by these samples, consistent with previous inferences based on angular momentum considerations.
astro-ph
astro-ph
ELEMENT RATIOS AND THE FORMATION OF THE STELLAR HALO Gerard Gilmore Institute of Astronomy, University of Cambridge, Madingley Road, Cambridge, UK CB3 0HA Department of Physics and Astronomy, The Johns Hopkins University, Baltimore, MD 21218 Rosemary F. G. Wyse ABSTRACT It is well-established that the vast majority of metal-poor Galactic halo stars shows evidence for enrichment by solely massive stars. Recent observations have identified more varied behavior in the pattern of elemental abundances measured for metal-rich, [Fe/H]∼ −1 dex, halo stars, with both high and low values of the ratio of [α/Fe]. The low values are most naturally due to the incorporation of iron from Type Ia supernovae. These 'low-alpha' halo stars have been interpreted as being accreted from dwarf galax- ies. However, these stars are on very high-energy radial orbits, that plunge into the Galactic center, from the outer regions of the stellar halo. We demonstrate here that known dwarf galaxies could not reproduce the observations, which require the proposed parent dwarf to have a very high mean density to provide 'low-alpha' stars on orbits of such low values of periGalactic distance. Rather, the observations are consistent with fragmentation of a gaseous proto-halo into transient star-forming regions, some of which are sufficiently dense to survive close to the Galactic center, and self-enrich to relatively high metallicities. Those that probe the outer halo sustain star formation long enough to incorporate the ejecta from Type Ia supernovae. Further, we point out that the values of the 'Type II plateau' in the element ratios for disk stars and for halo stars are equal. This implies that the (massive) stars that enriched the early disk and halo had the same IMF. However, the bulk of the halo stars with [Fe/H]∼ −1 dex that have been observed have significantly lower values of [α/Fe] than does a typical disk star of that metallicity. Thus there is a discontinuity in the chemical enrichment history between the halo and disk, as probed by these samples, consistent with previous inferences based on angular momentum considerations. Subject headings: Galaxy: halo -- Galaxy: formation -- Galaxy : evolution -- Galaxy -- abundances 1 1. Introduction Elemental abundances contain significantly more information than just 'metallicity', since differ- ent elements are synthesised in stars of different masses, and on different timescales. In particu- lar, oxygen and the alpha-elements are produced predominantly by massive stars, ejected into the interstellar medium by core-collapse, Type II, supernovae. Iron has important contributions by both Type II supernovae and Type Ia su- pernovae, the latter believed to result from the explosive nucleosynthesis of degenerate material, associated with a white dwarf in a binary sys- tem. The vast majority of local stars belonging to the Galactic halo population have enhanced values, compared to that of the Sun, of the ratio of the abundances of oxygen and of the alpha- elements to that of iron (e.g. Wheeler, Sneden & Truran 1989; Nissen et al. 1994; Gratton et al. 1997a). Further, these halo stars have very similar values of this 'over-abundance' of alpha-elements, [α/Fe]halo ∼ +0.3. This 'halo' value is, within nucleosynthetic yield uncertain- ties, just that predicted for chemical enrichment by a mean, IMF-averaged, Type II (core col- lapse) supernova, assuming a normal (i.e. solar- neighborhood) massive-star mass function (Wyse & Gilmore 1992; Nissen et al. 1994). Thus a typical halo star shows no evidence for nucleosynthetic products from Type Ia super- novae. This lack restricts the duration of star for- mation that created these stars to be shorter than the time it takes for there to have occurred signif- icant Type Ia explosions. This timescale is rather model-dependent. In the commonly-discussed double-degenerate models (Iben & Tutukov 1984; see Iben et al. 1997 for a recent discussion, in- cluding the expected statistics of close WD bina- ries) it is plausibly of order 1 Gyr after the forma- tion of the progenitor stars (e.g. Smecker-Hane & Wyse 1992; Matteucci & Fran¸cois 1992). A sim- ilar timescale for significant supernovae follows from models in which the white dwarf is accret- 2 ing due to Roche-Lobe overflow from a low-mass, evolved companion (Yungelson & Livio 1998). This restriction on the duration of star formation is not necessarily the same as a restriction on the global age range of the halo, since different re- gions of the halo could initiate (short-lived) star formation at different times, resulting in a signif- icantly larger total age range in the halo (see also Gilroy et al. 1988 for an analogous argument for rapid local enrichment, based on the appearance of s−process elements in metal-poor halo stars). It is very implausible that the duration of star formation across the entire halo was as short as 1 Gyr, no matter one's preconceptions of how the halo formed -- if 'monolithic', this requires remarkable synchronisation over ∼ 50 kpc, while if 'chaotic', the regions that formed the halo had remarkably uniform star formation properties. Star formation is likely to occur in transient overdense regions, perhaps of the mass of present- day Giant Molecular Cloud complexes in the Galactic disk, even in models that envisage a well-defined 'monolithic' Galactic potential well during the formation of the stellar halo (e.g. Fall & Rees 1985). Was there pre-enrichment of one region by another? Did some regions self-enrich sufficiently long to incorporate the products of Type Ia supernovae? Isochrone-based determi- nations of the ages of field halo stars are of lim- ited accuracy, and indeed the only technically- feasible means of detecting an extended duration of star formation that may be as short as a Gyr is to identify halo stars with evidence for pre- enrichment by Type Ia supernovae. The manifes- tation of this is lower values of [α/Fe] than can be understood from Type II supernovae alone. Given the expectation of some increase in metal- licity with time as star formation proceeds, the most natural place to look for such an effect is among the most chemically-enriched halo stars. We note that a variation in the massive-star IMF can also provide for lower values of element ratios, still with enrichment by only Type II su- pernovae. However, there is firm observational evidence for universality of the IMF, surprising as this may be (see papers in Gilmore & Howell 1998), and indeed any invoked IMF variation to explain lower values of [α/Fe] would be extremely ad hoc. Thus we favor interpretations within the framework of a fixed IMF. Several recent analyses, most notably that by Nissen & Schuster (1997), have discovered ex- actly the element ratio behavior expected for some enrichment by Type Ia supernovae. The authors of these analyses have favored, as the ex- planation for 'low-alpha' halo stars, accretion of stars from distinct objects such as satellite dwarf galaxies. Here we argue rather that the observa- tions are more naturally explained by the varia- tion in the duration of star formation in regions of the proto-halo. Fragmentation into cool, dense self-gravitating regions is fairly generic to models of the collapse of galaxy-scale perturbations; as- similation and mixing of these transient regions is not to be thought of as accretion. Of course, accretion, in the sense of a distinct other proto- galaxy merging into our proto-galaxy, remains a very attractive explanation for the creation of the thick disk (e.g. Gilmore & Wyse 1985; see Ve- lazquez & White 1998, for a recent simulation), some ∼ 12 Gyr ago (e.g. Gilmore, Wyse & Jones 1995). 2. The Observations of Elemental Abun- dances Halo stars with low values of the [α/Fe] ele- ment ratios were identified in a study of metal- rich halo stars, those with [Fe/H] ∼ −1 dex, by Nissen & Schuster (1997), and serendipitously in samples of local, low-metallicity halo stars by King (1997) and by Carney et al. (1997). Nissen & Schuster studied approximately equal numbers of 'disk' and 'halo' dwarfs (16 disk, 13 halo; so designated based on kinematics) in the metallic- ity range −1.3 ∼< [Fe/H] ∼< −0.4. Two giant stars in each of the 'younger' (e.g. Richer et al. 1996) globular clusters Ruprecht 106 and Pal 12 were analysed by Brown, Wallerstein & Zucker (1997) 3 and found to have approximately solar values of alpha-element-to-iron ratios.1 Figure 1 shows the [Mg/Fe] data, as a func- tion of [Fe/H], for the Nissen & Schuster (1997) sample of disk and halo stars, plus the anoma- lous metal-poor halo stars from King (1997) and Carney et al. (1997); the shaded region indicates the locus of the majority of metal-poor halo stars (e.g. Nissen et al. 1994; Gratton et al. 1997a; note that there are sometimes zero-point offsets in the value of [Mg/Fe] between different observa- tional data sets, which can lead to apparent gra- dients in [α/Fe] plotted against [Fe/H] if all pub- lished data are naively combined. Each data set shows similar behavior in their regions of over- lap in [Fe/H]. Since it is probable that these off- sets reflect differences in analysis methods rather than real systematic variations in [α/Fe] as a function of [Fe/H], we do not consider them fur- ther here). As may be seen in the Figure, the major- ity of the metal-rich halo stars observed shows significantly lower values of [α/Fe] than the en- hanced, super-solar values of the alpha-element ratios derived for the more metal-poor halo stars. Thus the halo stars with enhanced element ra- tios are 'anomalous' in this metallicity range. This contrasts with the typical (metal-poor) halo where the stars with low values of [α/Fe] are the anomalous ones. The halo stars of all metallic- ity with low values of [α/Fe] share one property -- they tend to be on orbits with large values of apoGalactic distances, Rapo ∼> 10 kpc, whereas the 'normal' halo stars are not to so biased to the outer halo. This trend suggests a real phys- ical meaning to the present orbital parameters. Further, all the halo stars of the Nissen & Schus- ter (1997) sample are on derived orbits with very 1Note that the relative ages of globular clusters are gener- ally derived with fixed -- enhanced -- values of the alpha- element ratios, so this result, if a property of all stars in the cluster, increases the age assigned to these clusters and thus reduces the age spread within the system of Galactic globular clusters. low periGalactic distances, Rperi ∼< 1 kpc; this is consistent with the apparent correlation be- tween Rperi and metallicity established for such kinematically-selected samples by Ryan & Nor- ris (1991)2, such that the more metal-rich halo stars tend to be on orbits of smaller periGalac- tic distances. Thus all the metal-rich halo stars observed are on orbits with small values of Rperi, while of these, the stars with low values of [α/Fe] are on orbits with large values of Rapo. Note that it is not the case that all metal- poor halo stars with large Rapo necessarily have low values of [α/Fe], although such a trend ex- ists for the metal-rich stars with low values of [α/Fe]. As demonstrated in Table 10 of Carney et al. (1997) there are many metal-poor stars that have orbits associating them with the outer halo, but that have normal, enhanced, values of the alpha-element ratios. The metal-poor halo stars with low values of [α/Fe] are on retrograde orbits (King 1997; Carney et al. 1997), but there is no such signature in the kinematics of the metal-rich stars with low values of [α/Fe]. 3. What Does It All Mean? We wish to understand why a small fraction of stars in the stellar halo, biased to metal-rich stars on high-energy, high-eccentricity orbits, have low values of [α/Fe]. Consider the pattern of element ratios in self- enriching systems (of fixed IMF) with given star formation rates. The solid lines in Figure 1 illus- trate schematically the trends produced in sim- 2Carney et al. (1996) argue that this correlation arises only when the azimuthal velocity is used to isolate halo stars and interpret it as reflecting the existence of a separate population of very metal-poor 'far' halo stars on retro- grade orbits. While this is intriguing, the significance of a distinct population on retrograde orbits is difficult to establish given that the angular momentum sign of the orbits of fully 38% of halo stars will be retrograde, given a mean azimuthal streaming velocity of +30 km/s, and a V-dispersion of 100 km/s, as found by Ryan & Norris (1991). 4 ple (homogeneous) models of chemical evolution with differing star formation rates. The impor- tant feature is the location of the turndown due to iron from Type Ia supernovae -- since the short- est pre-explosion lifetime of Type Ia supernovae is presumed to be a fixed value, the location of this turndown occurs at a higher value of [Fe/H] for a higher star formation rate, reflecting the more rapid metallicity increase with time (note that the effect of gas flows on the location of the turndown from the 'Type II plateau' at early times/lower metallicities can be incorporated by a suitably-modified star-formation rate; inflow of unenriched material will reduce [Fe/H] keeping [α/Fe] fixed. Outflow driven by Type II super- novae will behave similarly). Such an effect has been proposed to explain the element ratios seen in local disk stars, born at different locations in the disk (Edvardsson et al. 1993). Figure 1 shows that the observed differences in element ratio patterns are inevitable conse- quences of a variation of the star formation rate. The amplitude of the invoked variation in star formation rates is required only to be such as to allow up to ∼< 1 Gyr longer duration of star for- mation to produce lower values of [α/Fe]. This difference in age ranges is undetectable with current isochrone-based techniques. Note that there is no requirement from the element ra- tio data that the progenitor stars of the Type Ia supernovae be part of the same star form- ing region as the gas they will eventually enrich. However, the observed trend for low values of [α/Fe] to be observed for higher-metallicity halo stars favors rather self-enrichment scenarios, in (sub)structure of relatively deep potential wells. Clearly, the data are consistent with the lower [α/Fe] stars forming over longer timescales, al- lowing incorporation of iron from Type Ia super- novae. But where did these halo stars form? In a distinct separate galaxy, or essentially where we observe them now? 3.1. Accretion Models Nissen & Schuster (1997) considered two pos- sibilities, either that the outer regions of the stel- lar halo of the Galaxy had a longer duration of star formation than did the inner halo, or that the halo stars with low values of [α/Fe] were ac- creted from a dwarf galaxy within which star for- mation is presumed to have been sustained over a sufficiently long period, longer than in the field halo. The authors favor the latter explanation. Indeed, 'distinct accretion events' are the pre- ferred explanation of all the authors for the data in Figure 1, and also for the 'anomalous' glob- ular clusters (it has been suggested that these two clusters have been captured from the Large Magellanic Cloud; Lin & Richer 1992). Motivation for the interpretation in terms of 'accretion' came partially from the prediction of low values of the alpha-element to iron ratios in the stars of dwarf galaxies in which there is ex- tended star formation, perhaps even successive bursts of star formation, as inferred for many of the extant satellite galaxies associated with the Milky Way (Gilmore & Wyse 1991; Una- vane, Wyse & Gilmore 1996). It is clear from observations that only the more massive dwarf systems can self-enrich to a metallicity of 1/10 the solar value (see, for example the metallicity -- luminosity (mass) relation of Lee et al. 1993). Of the local dwarf galaxy companions to the Milky Way, only the Fornax dSph, the Sagittar- ius dSph and the Magellanic Clouds have a sig- nificant population of stars as metal-rich as the observed low-alpha halo stars. However, these systems also contain stars of a wide range of ages (e.g. Smecker-Hane 1996), and late accretion of stars from systems like these satellite galaxies would produce stars in the halo which are many Gyr younger than the bulk of the halo, reflecting the star-formation history of the satellite galaxy parent. The present sample of 'low-alpha' halo stars shows no obvious signature of having a rel- ative youth of this amplitude, which would be manifest in their colors (e.g. Unavane, Wyse & 5 Fig. 1. -- Iron abundance against element ra- tio [Mg/Fe] for the halo stars (open circles) and disk stars (filled circles) from Nissen & Schus- ter (1997; their Table 4). The disk stars show a well-defined trend, contrasted with the appar- ent scatter of the halo stars. The shaded region marks the locus of normal metal-poor halo stars in this plane. The star symbol represents the anomalous metal-poor halo subgiant of Carney et al. (1997), and the open triangle the anoma- lous common proper-motion pair of halo dwarfs from King (1997). The solid lines schematically illustrate the expected pattern of element ratios in self-enriching systems of constant IMF, but varying star formation rates. Gilmore 1996). Thus late accretion from systems like the ex- tant satellite galaxies is not favored. What other kind of accretion is possible? Models of the formation of the stellar halo generically invoke early star formation in tran- sient ∼ 107 M⊙ local potential wells, perhaps re- flecting the initial primordial fluctuation power spectrum, or formed by thermal and/or gravi- tational instability within a larger system. The turn-around, collapse and relaxation to dynam- ical equilibrium of the present-day Galactic po- tential takes at least a couple of global dynamical times, or a few Gyr. Thus disruption and as- similation of transient systems during the relax- ation of the large-scale Galactic potential should not be considered as 'accretion', since these sys- tems never really had a separate identity. Sys- tems that do retain coherence through this pe- riod can be said to be 'accreted' if assimilated subsequently, and it is these we consider next. Since we are discussing essentially old stars, we shall focus on the stellar components of this sub- structure; in any case a satellite galaxy would be most likely stripped of its gas after a few passages through the gaseous disk of the Milky Way with periGalactic distances like those observed for the halo stars. 'Substructure' orbiting within the (dark halo) potential of a larger 'parent' galaxy is subject to tidal forces, and to dynamical friction. The mass and the (mean) density of the substructure will to a large extent determine its fate. Dynami- cal friction will cause the orbit of the substruc- ture to decay in radius, on a timescale inversely proportional to the mass of the substructure (for given orbit and large galaxy); relatively massive substructure will sink to the center of the larger galaxy on a few orbital times. Tidal stripping will cause the substructure to lose material when its internal gravity is overcome by the gravita- tional field of the larger galaxy; the simple Jacobi criterion implies that a system will lose material unless (or until) its internal density is of order 6 three times the mean density of the larger galaxy interior to its orbit; simulations have shown that this simple criterion works remarkably well (e.g. Johnston, Hernquist & Bolte 1996). Diffuse sub- structure will lose material while orbiting even in the far outer regions of the larger galaxy. The orbits of stars that have been accreted into the Galaxy by tidal effects on a satellite galaxy are only slightly different from that of the satellite at the time the stars were captured; this produces a kinematic signature of moving groups of stripped stars, that can persist for many orbital times (cf. Tremaine 1993; Johnston 1998). Thus one may consider four broad classes of dissipationless (stars plus dark matter) substruc- ture: massive, dense systems will survive fairly intact while sinking to the center of the larger galaxy; massive, diffuse systems will be largely disrupted by tides during their orbital decay; low mass, dense systems will be largely unaffected, while low mass, diffuse systems will be disrupted with no orbital decay. The stars under consideration here are on rather eccentric orbits, but with even Rapo much less than that of typical extant satellite galax- ies. As discussed by Velazquez & White (1998), for example, in their study of the interaction between satellite galaxies and large disk galax- ies, local application of the standard (Chan- drasekhar) dynamical friction formula provides reliable results in the case of non-circular orbits. Thus due to the higher densities in the central regions of the bigger galaxy, the frictional effects are higher there. Further, most tidal stripping occurs at pericenter passage (e.g. Velazquez & White 1998 and references therein). Thus if the observed halo stars were removed from a satel- lite galaxy by tidal stripping, their very low peri- Galactic distances would require that the satel- lite galaxy be able to penetrate deep inside the Milky Way. Their low values of Rapo also would, in this scenario, require a rather massive satel- lite, so that dynamical fraction can operate on less than a Hubble time. The periGalactic distances provide the more stringent constraint. As discussed above, the low [α/Fe] stars are on radial orbits, with periGalac- tic distances ∼< 5 kpc, and indeed the majority of the stars having Rperi ∼< 1 kpc. A lower limit to the mean density of the Milky Way Galaxy inte- rior to 1 kpc is given by adopting just the mass of the central bulge, ∼ 1 × 1010M⊙ (e.g. Kent 1992). The predominant stellar populations in the Galactic bulge and inner disk are probably at least ∼ 8 Gyr old (Ortolani et al. 1995; Holtz- man et al. 1993), consistent with theoretical ex- pectations for the timescale of assembly of the 'core' of the Milky Way in hierarchical-clustering models (e.g. Lacey & Cole 1993). Thus one may use inferences from the mass distribution of the inner Galaxy to constrain the fate of substruc- ture back to that time. The mean density interior to the periGalac- tic approach (∼ 1 kpc) of the 'low-alpha' stars is then ∼> 2.5 M⊙ pc−3. To be robust against tides, a satellite galaxy on an orbit with this small a periGalactic distance would need an mean in- ternal density of around three times this, or ∼> 7 M⊙ pc−3. This value is significantly larger (by orders of magnitude) than the typical mean densities derived for dwarf galaxies, both gas- poor and gas-rich (e.g. Puche & Carignan 1991; Ibata et al. 1997). Thus essentially all dwarf galaxies as we see them now (presumably the survivors) would be disrupted after only a couple of such pericenter passages (see Johnston, Hern- quist & Bolte 1996). Indeed, they would likely be disrupted significantly earlier in their orbital decay, when the pericenter distances are of or- der the present derived apocenter distances of the anomalous halo stars (e.g. Ibata et al. 1997). Thus, if the halo stars under discussion repre- sent stars captured from a satellite galaxy, that galaxy had to be sufficiently massive to spiral into the center of the Milky Way in less than a Hubble time -- this implies a mass perhaps 10% that of the older stars in the central Milky Way, or equal to the entire mass of the stellar halo (e.g. Carney, Latham & Laird 1990). This satellite had also to be significantly more robust than any of the extant satellite galaxies (the survivors). If this satellite contributed to the formation of the thick disk, it did so without donating any of its own stars to the disk. Each of these is rather surprising. 3.2. Halo Star-Formation Regions The requirements for the progenitor structures for the metal-rich, 'low-alpha' halo stars may be summarised as being dense enough to survive to a periGalactic distance of ∼ 1 kpc, and have a deep enough local potential well to self-enrich to −1 dex, while sustaining star formation for ∼ 1 Gyr, being robust not only to internal feed- back from star formation, but also external ef- fects. Plausibly the first two of these require- ments are satisfied by a density criterion for the substructure, given that the escape energy per unit mass scales as ρ r2. Sustaining star for- mation requires gas retention, and the further requirement that the effects of collisions with other star-forming clouds, or passages through the (forming) gas disk of the Galaxy, be min- imised. As discussed by many authors (see e.g. Gunn 1980; Jones, Palmer & Wyse 1981) gravitational instability in the cooling layers behind shocked gas in the haloes of protogalaxies produces sys- tems of density ∼> 50M⊙pc−3 and of characteris- tic (Jeans) mass of a few times 106M⊙. A more sophisticated analysis by Fall & Rees (1985) in- cluded thermal instability in the hot, gaseous halo of a proto-galaxy, like the Milky Way, to produce cool, dense regions, that are seeds for gravitational instability. Fall & Rees (1985) ar- gue that a fairly massive proto-galaxy, with cir- cular velocity greater than ∼ 150 km/s, is re- quired for this scenario (essentially to provide the hot phase of the gaseous halo). In flat Cold- Dark-Matter dominated cosmologies (e.g. White & Frenk 1991), dark haloes of this circular ve- locity have peak abundance at redshifts ∼> 3, or 7 lookback times of ∼> 12 Gyr. Thus this picture of fragmentation inside dark haloes is reasonable even in hierarchical clustering scenarios. The fragments undergo star formation and are subsequently dispersed, either by the inter- nal effects of star formation, or the external ef- fects of cloud-cloud collisions or tidal disrup- tion by the Galactic potential. It is plausi- ble that the fragments probing smaller Galac- tocentric distance have larger internal densities (indeed, as predicted by the model of Fall & Rees 1985) and larger escape velocities, allowing greater self-enrichment and thus providing a rela- tion between Rperi and [Fe/H] as observed in the remnant stars, now the field halo (higher metal- licity for lower Rperi; Ryan & Norris 1991). The most dense regions could plausibly form globu- lar clusters, as suggested by Fall & Rees (1985), even some of which are subsequently destroyed (e.g. Gnedin & Ostriker 1997). The bias to large apoGalactic distances for the orbits of the 'low-alpha' stars follows natu- rally from the expected longer survival times of large, self-enriching clouds in the outer regions of proto-Galaxy, due to the longer collision times there, and the larger orbital times. Addition- ally, a large-scale density gradient could provide slower local star formation rates in the outer re- gions, provided the properties of density pertur- bations correlate with their local background. Thus fragmentation would appear to provide the kind of substructure required by the obser- vations. 4. Disk -- Halo (Dis)connection A striking aspect of the data in Figure 1 is that the value of the 'Type II plateau' of element ratios is the same for the halo stars as for the disk stars of the same metallicity. Following the same logic as Wyse & Gilmore (1992), this im- plies that the IMF of the (massive) stars that enriched these stars was the same in the early disk and halo. Inversion of the value of the ele- ment ratios to determine the actual slope of the massive star IMF requires adoption of theoreti- cal supernova yields and is hence uncertain, but the robust statement that the IMF was invariant in the enrichment of disk and halo may be made immediately. It is also clear from Figure 1 that most of the metal-rich halo stars do not have element ra- tios equal to this 'Type II plateau', and that the typical halo and disk stars of equal metallicity ([Fe/H] ∼ −1 dex) do not have the same values of the element ratios. There is not the smooth continuity in chemical enrichment from halo to disk often derived in models of chemical evolu- tion. Of course, these halo stars are on orbits of very different angular momentum than those of the disk stars, further emphasizing the dis- tinction between local samples of disk and halo, and the low likelihood that the halo, as sampled, pre-enriched the disk (see Wyse & Gilmore 1992; Gilmore 1995). There is also a spread in [Fe/H] at fixed values of [Mg/Fe]; as mentioned earlier in section 3, this may be achieved by different gas flows/mixing and star formation histories. The onset of star formation in the local disk, as measured by the white dwarf luminosity func- tion, has most recently been estimated to be 10 -- 12 Gyr ago (an increase over earlier estimates, in part due to the inclusion of phase separation and crystallisation energy in the white dwarf cool- ing curves; Hernanz et al. 1994; Chabrier 1998). This older age is in agreement with the ages of the oldest open clusters (e.g. Phelps 1997) and the age of the thick disk (e.g. Gilmore, Wyse & Jones 1995) and in better agreement with the ages of individual stars from Edvardsson et al. (1993; see also Ng & Bertelli 1998 for post- Hipparcos age estimates for stars from this sam- ple). The recent downward revision in the ages of the halo globular clusters to ∼ 12 Gyr, based on Hipparcos calibrations of the luminosities of field subdwarfs (e.g. Gratton et al. 1997b; Reid 1998), leaves little room for a significant hiatus between the epoch of halo star formation and the 8 little information about the wings of the metal- licity distribution of the stellar halo. The present data for disk and halo stars of 'overlapping' metallicities [Fe/H]∼ −1 dex show a disk -- halo dichotomy, in that the elemental abundances are different for typical disk stars and for typical halo stars in this metallicity range, with the halo stars having lower values of [α/Fe] than the disk stars. This complicates the requirements for chemical evolution models for the disk that appeal to gas inflow from the halo; indeed if the halo had ejected gas that pre-enriched the disk, one might have expected the opposite situation, with the disk stars having lower values of element ratios than the halo stars. A simplification of chemical evolution models is that the elemental abundance data favor a fixed massive-star IMF for both disk and halo. We are grateful to NATO Scientific Affairs for a grant to aid our collaboration. RFGW thanks all at the Center for Particle Astrophysics (UC Berkeley) for hospitality during the early stages of writing of this paper. She acknowledges partial support from NASA, through ATP grant NAG5- 3928. onset of star formation in the disk (however, this calibration depends on many corrections and is not unique -- see e.g. Pont et al. 1998). It would follow that disk and halo were independent enti- ties from the very early stages of Galaxy forma- tion. 5. Conclusions The recent observations of halo stars with elemental abundances reflecting pre-enrichment from intermediate-mass stars and Type Ia su- pernovae can be understood in many models of halo formation. The establishment of trends with kinematics, suggested at in the present data, will be important in distinguishing models, as will be age determinations. The fact that the 'low- alpha' halo stars tend to be metal-rich favors self- enrichment scenarios, and the survival of indi- vidual star-forming complexes for approximately a Gyr. That the observed low-alpha stars are on orbits that plunge close to the Galactic cen- ter also implies that a reasonably robust parent system is required. We argue for an interpreta- tion that simply appeals to a small variation in the duration of star formation within regions of the stellar halo, rather than 'distinct accretion events'. The present data implicates a significant frac- tion of the metal-rich, outer halo. One may normalise to the stellar halo as a whole, using the metallicity distribution for local halo stars of Carney et al. (1996) in which ∼ 15% of the stellar halo has [Fe/H]∼> −1 dex, combined with the halo structural parameters of Unavane, Wyse & Gilmore (1996) in which ∼ 10% of the halo is located beyond Z = 5 kpc, to conclude that at most a few percent (10% times 15%) of the stellar halo is being discussed. More data -- not restricted to detailed elemental abundances for selected halo stars -- will clearly help to substan- tiate these estimates. For example, the metal- licity distribution of the halo is derived from lo- cal, kinematically-selected samples, and the inner halo is poorly studied. Further, there is currently 9 REFERENCES Gnedin, O.Y. & Ostriker, J.P. 1997, ApJ, 474, Brown, J., Wallerstein, G. & Zucker, D. 1997, AJ, 114, 180 Carney, B., Latham, D. & Laird, J. 1990, AJ, 99, 572. Carney, B., Wright, E., Sneden, C., Laird, J., Aguilar, L. & Latham, D. 1997, AJ, 114, 363 Carney, B., Laird, J., Latham, D. & Aguilar, L. 1996, AJ, 112, 668 Carney, B., Latham, D., Laird, J. & Aguilar, L. 1994, AJ, 107, 2240 Chabrier, G. 1998, in IAU Symposium 189, 'Fundamental Stellar Properties: The Inter- action between theory and observation', eds T.R. Benning, A.J. Booth & J. Davis (Kluwer: Dordrecht) in press Edvardsson, B., Andersen, J., Gustafsson, B., Lambert, D., Nissen, P. & Tomkin, J. 1993 A&A, 275, 101 Fall, S.M. & Rees, M.J. 1985, ApJ, 298, 18 Gilmore, G. 1995, in IAU Symposium 164, 'Stel- lar Populations', eds G. Gilmore & P. van der Kruit (Kluwer, Dordrecht) p99 Gilmore, G. & Wyse, R.F.G. 1985, AJ, 90, 2015 Gilmore, G. & Wyse, R.F.G. 1991, ApJL, 367, L55 Gilmore, G. & Howell, D. 1998, eds., 'The Stellar IMF', ASP Conference series 142. Gilmore, G., Wyse, R.F.G. & Jones, J.B. 1995, AJ, 109, 1095. 223 Gratton, R., Carretta, E., Clementini, G. & Sne- den, C. 1997a, in proc. 'Hipparcos Venice 97 Symposium', ESA SP-402. Gratton, R., Fusi Pecci, F., Carretta, E., Clementini, G., Corsi, C & Lattanzi, M. 1997b, ApJ, 491, 749 Gunn, J. 1980, in 'Globular Clusters', eds D. Hanes & B. Madore (Cambridge Univer- sity Press: Cambridge) p301 Hernanz, M., Garcia-Berro, E., Isern, J., Mochkovitch, R., Segretain, L. & Chabrier, G. 1994, ApJ, 434, 652 Holtzman, J., Light, R.M., Baum, W.A., Worthey, G., Faber, S.M., Hunter, D.A., O'Neil, E.J., Kreidl, T.J., Groth, E.J. & West- phal, J.A. 1993, AJ, 106, 1826 Ibata, R., Wyse, R.F.G., Gilmore, G., Irwin, M. & Suntzeff, N. 1997, AJ, 113, 634. Iben, I. & Tutukov, A.V. 1984, ApJS, 54, 335 Iben, I., Tutukov, A.V. & Yungelson, L. 1997, ApJ, 475, 291 Johnston, K., Hernquist, L. & Bolte, M. 1996, ApJ, 465, 278 Johnston, K. 1998, ApJ, 495, 297 Jones, B.J.T., Palmer, P.L. & Wyse, R.F.G. 1981, MNRAS, 197, 967 Kent, S. 1992, ApJ, 387, 181 King, J. 1997, AJ, 113, 2302 Gilmore, G., Wyse, R.F.G. & Kuijken, K. 1989, Lacey, C. & Cole, S. 1993, MNRAS, 262, 267 ARAA, 27, 555 Gilroy, K.K., Sneden, C., Pilachowski, C. & Cowan, J.J. 1988, ApJ, 327, 298. Lee, M.G., Freedman, W., Mateo, M., Thomp- son, I., Roth, M. & Ruiz, M.-T., 1993, AJ, 106, 1420 10 Lin, D.N.C. & Richer, H. 1992, ApJL, 358, L57 White, S.D.M. & Frenk, C. 1991, ApJ, 379, 52 Matteucci, F. & Fran¸cois, P. 1992, A&A, 262, L1 Wyse, R.F.G. & Gilmore, G. 1992, AJ, 104, 144 Nissen, P.E. & Schuster, W.J. 1997, A&A, 326, Yungelson, L. & Livio, M. 1998, ApJ, 497, 168 751 Nissen, P., Gustafsson, B., Edvardsson, B. & Gilmore, G. 1994, A&A, 285, 440 Ng, Y.K. & Bertelli, G. 1998, A&A, 329, 943 Ortolani, S., Renzini, A., Gilmozzi, R., Marconi, G., Barbuy, B., Bica, E. & Rich, R.M. 1995, Nature, 377, 701 Phelps, R. 1997, ApJ, 483, 826 Pont, F., Mayor, M., Turon, C. & vandenBerg, D.A. 1998, A&A, 329, 87 Puche, D. & Carignan, C. 1991, ApJ, 378, 487 Reid, I.N. 1998, AJ, 115, 204 Richer, H., Harris, W.E., Fahlman, G., Bell, R., Bond, H.E., Hesser, J.E., Holland, S., Pryor, C., Stetson, P., Vandenberg, D.A. & van den Bergh, S. 1996, ApJ, 463, 602 Ryan, S. & Norris, J. 1991, AJ, 101, 1835 Smecker-Hane, T. A. 1996, in 'Star Forma- tion Near and Far' AIP Conf Proc vol 393, eds. S. Holt & L. G. Mundy (AIP Press: New York), p571 Smecker-Hane, T. & Wyse, R.F.G. 1992, AJ, 103, 1621 Tremaine, S., 1993, in 'Back to the Galaxy', eds S. Holt & F. Verter (AIP: New York) p599 Unavane, M., Wyse, R.F.G. & Gilmore, G. 1996, MNRAS, 278, 727 Velazquez, H. & White, S.D.M. 1998, MNRAS, in press Wheeler, J., Sneden, C. & Truran, J. 1989, ARAA, 27, 279 This 2-column preprint was prepared with the AAS LATEX macros v4.0. 11
astro-ph/0606014
1
0606
2006-06-01T00:01:48
Coherent synchrotron emission from cosmic ray air showers
[ "astro-ph" ]
Coherent synchrotron emission by particles moving along semi-infinite tracks is discussed, with a specific application to radio emission from air showers induced by high-energy cosmic rays. It is shown that in general, radiation from a particle moving along a semi-infinite orbit consists of usual synchrotron emission and modified impulsive bremsstrahlung. The latter component is due to the instantaneous onset of the curved trajectory of the emitting particle at its creation. Inclusion of the bremsstrahlung leads to broadening of the radiation pattern and a slower decay of the spectrum at the cut-off frequency than the conventional synchrotron emission. Possible implications of these features for air shower radio emission are discussed.
astro-ph
astro-ph
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (2006) Printed 27 October 2018 (MN LATEX style file v2.2) Coherent synchrotron emission from cosmic ray air showers Qinghuan Luo School of Physics, The University of Sydney, NSW 2006, Australia -- Received in original form February, 2006 ABSTRACT Coherent synchrotron emission by particles moving along semi-infinite tracks is dis- cussed, with a specific application to radio emission from air showers induced by high- energy cosmic rays. It is shown that in general, radiation from a particle moving along a semi-infinite orbit consists of usual synchrotron emission and modified impulsive bremsstrahlung. The latter component is due to the instantaneous onset of the curved trajectory of the emitting particle at its creation. Inclusion of the bremsstrahlung leads to broadening of the radiation pattern and a slower decay of the spectrum at the cut-off frequency than the conventional synchrotron emission. Possible implications of these features for air shower radio emission are discussed. Key words: Plasmas -- radiation mechanisms: nonthermal -- cosmic rays 1 INTRODUCTION high energy growing attention recently is detection of Gorham, et al. 2004; Falcke et al. (Dagkesamanskii & Zheleznykh Alvarez-Muniz, V´azquez, & Zas on There cosmic radio rays 1989; Zas, Halzen & Stanev 1992; Hankins, Ekers, & O'Sullivan 2000; 1996; 2004; Falcke & Gorham 2003; Falcke, Gorham, & Protheroe 2005). High energy cosmic rays can initiate an electromagnetic cascade in a medium where relativistic electrons and positrons can be produced in a volume with a longitudinal (along the line of sight) dimension being smaller than the relevant radio wavelength. So, particles form a coherent bunch, acting like a single charged particle that emits a short burst of coherent radio emission. Two radiation processes have been considered: coherent Cerenkov radi- ation (Arskar'yan 1962, 1965) and coherent synchrotron radiation (Falcke & Gorham 2003; Huege & Falcke 2003). (Coherent radio emission from air showers was first consid- ered by Kahn & Lerche (1966), also Colgate (1967), but their theory was not explicitly based on geosynchrotron emission.) The former requires charge asymmetry, say an excess of electrons, and a relatively dense medium for coherent Cerenkov emission to be at radio frequency. For example, the Moon is a good target for high energy neutri- nos that can lead to a cascade in the lunar rocks. Excess electrons can develop leading to coherent Cerenkov emission at radio frequency with wavelength comparable with or larger than the longitudinal size of the cascade (Arskar'yan 1962, 1965). For air showers, coherent synchrotron emission is generally considered more important than Cerenkov radiation (in the radio band) (Falcke & Gorham 2003; Gorham, et al. 2004). The shower produces a bunch of relativistic electrons and positrons emitting coherent syn- chrotron radiation in geomagnetic fields. Unlike coherent Cerenkov radiation, coherent synchrotron emission does not need charge asymmetry (Falcke & Gorham 2003). (Suprun, Gorham & Rosner simulation based on the So far, the relevant spectra of coherent synchrotron showers were commonly calculated emission from air retarded- using numerical potential method 2003; Huege & Falcke 2005a,b). In this method, the radiation is calculated from the retarded potential (Jackson 1998). Although coherent synchrotron emission has been consid- ered analytically (Aloisio & Blasi 2002; Falcke & Gorham 2003; Huege & Falcke 2003), their calculation is based on the standard synchrotron radiation formula, which does not include the effect due to the particle's finite track. For air showers, effective coherent emission occurs in the core of the shower where most of radiating particles are created. Thus, it is of interest to consider radiation by a particle moving along a semi-infinite trajectory. Apart from the usual synchrotron emission there is emission due to the onset of the particle's curved trajectory. The latter component is referred to as the modified impulsive bremsstrahlung (MIB) as it is due to the combined effect of the usual impulsive (or prompt) bremsstrahlung due to particle (e±) creation, which is modeled as an abrupt jump in the particle's velocity from zero to c, and curvature of the particle's trajectory. It is worth noting that the finite track effect was considered for Cerenkov radiation (Tamm 1939) and was taken into account explicitly in calculation of cosmic-ray induced showers in a dense medium (Zas, Halzen & Stanev 1992). In the case of Cerenkov radiation, the finite track leads to a reduction in radiation intensity and modification 2 Luo of angular distribution of the emission. Since the main objective of studying radio emission from air showers is to infer the properties of the cosmic rays that induce the showers, one needs to determine the radio spectrum accurately. In this paper, we present a general formalism for coherent synchrotron radiation from a nonstationary many-particle system, which takes account of MIB. The formalism developed here is based on the single particle treatment (Melrose & McPhedron 1991), in which radia- tion is due to a current associated with particle's motion in a medium. Here, the current is regarded as extraneous as it is different from that due to the plasma response to waves. The spectrum of radiation from a many-particle system is derived from the total current that is obtained by adding all the currents due to individual particles. In the relativistic limit as in the usual synchrotron radiation, the spectrum can be expressed in terms of the Airy functions and as a result, the radiation is highly beamed. In Sec.2, a general formalism for synchrotron emission from a many-particle system is derived by including the ef- fect of MIB due to the effect of a particle's semi-infinite track. Coherent synchrotron emission is considered in Sec. 3 and the application to air showers is discussed in Sec. 4. 2 ENERGY SPECTRUM In cosmic ray air showers, electrons and positrons are created with a relativistic velocity, emitting synchrotron radiation in the geomagnetic field. Effective coherent emission by these secondary particles occurs at the core of the shower located very close to where most particles are produced. Thus, the finite track effect, in particular the initial position of the par- ticle's orbit, can be important and needs to be included ex- plicitly in the calculation of the synchrotron spectral power. In the following, we start with the single particle formalism. 2.1 Single particle treatment Consider a charged (q) particle created at t = 0 moving along a trajectory x(t) with a flight time T . The current associated with the particle is j(ω, k) = qZ T 0 dt v(t) exp [i (ωt − k · x(t))] , (1) where v = dx/dt is the particle's velocity. The current (1) can be regarded as an extraneous current due to a single particle's motion (as compared to the induced current due to plasma response). In the usual application to radiation in astrophysical plasmas, the time integration is taken from −∞ to ∞ (Melrose 1986). A finite T , particularly the initial point at t = 0, introduces a boundary (finite track) effect into (1). The energy spectrum can be found from the ex- pression (Melrose & McPhedron 1991) UM (k) = M · j2, 1 2ε0 e∗ where e∗ M is the complex conjugate of the polarization eM of the wave emitted in the mode M . The single particle's spectral power can be derived from PM = UM /T . (2) The orbit of a charged particle spiraling in a magnetic field B oriented along the z-axis can be written as x(t) = x0c + v⊥ Ω (cid:20)sin(cid:0)Ωt − ψ0(cid:1)ex + cos(cid:0)Ωt − ψ0(cid:1)ey(cid:21) (3) +vkt ez, where t > 0, x0c is the initial position of the particle's gy- rocenter, ψ0 is the initial gyrophase, which is defined here as the the azimuthal angle of the particle's initial velocity relative to the magnetic field, Ω = ηΩe/γ, q = ηe, η is the charge sign, γ = 1/(1 − v2/c2)1/2 is the Lorentz factor, and Ωe = eB/me is the gyrofrequency. The standard method to calculate the current is to expand the exponential term in terms of Bessel functions (Melrose & McPhedron 1991). In the relativistic limit as in the case relevant here, instant emission can only be seen during a very short time interval ∆t ∼ Rg/cγ3 = β⊥/Ωeγ2, where Rg = v⊥/Ω is the gyrora- dius. Thus, the orbit can be expanded on tΩe ∼ β⊥/γ2 ≪ 1 and the exponential in (1) can be expressed into the form ωt − k · x ≈ d + aτ + bτ 3, where τ = t + t0 with t0 = Ω−1 tan(φ − ψ0) and a = ω − kkvk (4) 2 tan(φ − ψ0) sin(φ − ψ0)(cid:21), b = 1 d = −k · x0c − −k⊥v⊥(cid:20)cos(φ − ψ0) + 1 6 k⊥v⊥Ω2 cos(φ − ψ0), Ω(cid:20)ω − kkvk cos(φ − ψ0)(cid:21) tan(φ − ψ0). sin2(φ − ψ0) 3 k⊥v⊥ − 2 1 (5) We assume the observation direction is k = k/k = (sin θ cos φ, sin θ sin φ, cos θ) and define spherical coordinates k = er. The projection of the current to the plane perpen- dicular to k can be written as j⊥ ≡ j − kk · j, which has the following components: j⊥θ ≈ eη"v⊥(cid:18) cos(φ − ψ0) F +iΩF ′ sin(φ − ψ0)(cid:19) cos θ − vkF sin θ# eid, j⊥φ ≈ ieηv⊥ΩF ′ cos(φ − ψ0) eid, where the relevant integrals are defined as dτ ei(aτ +bτ 3), F = Z T +t0 F ′ ≡ ∂F ∂a t0 = iZ T +t0 t0 dτ τ ei(aτ +bτ 3). (6) (7) The relativistic beaming (γ ≫ 1) implies that φ − ψ0 ≪ 1 and α−θ ≪ 1, where α = arctan(v⊥/vk) is the pitch angle. In these approximations, one has a ≈ ω[1− nβ cos(θ− α)] ≈ (ω/2)[2(1 − n) + γ−2 + (θ − α)2] and b ≈ (nωΩ2/6) cos(φ − ψ0) sin θ sin α, where the refraction index is assumed to sat- isfy 1 − n ≪ 1. Coherent synchrotron emission from cosmic ray air showers 3 2.2 Spectrum One may write the energy spectrum as the energy radi- ated per unit frequency per unit solid angle, U (ω, k) = (ω2/8π3c3)PM UM (k), where UM (k) is given by (2) and the summation is made over two orthogonal modes. One finds U (ω, k) = ≈ ω2 16π3ε0c3(cid:16)j⊥θ2 + j⊥φ2(cid:17) 16π3ε0c(cid:20)(α − θ)2F2 + e2ω2 Ω2 e γ2 sin2 α F ′2(cid:21). (8) The integrals in (7) can be expressed in terms of the Airy functions provided that the flight time T is much longer than the duration of the synchrotron pulse (∆t ∼ 1/Ωeγ2). In this approximation, one may take the limit T + t0 → ∞ and the integrals reduce to the form 0.25 0.2 0.15 0.1 0.05 0 -0.05 0 2 4 6 8 10 Figure 1. Plots of Gi(ξ) (solid) and Gi′(ξ) (dashed). F ≈ F ′ ≈ ξ = π π (3b)1/3" Ai(ξ) + iGi(ξ)# − Φ(a, b), (3b)2/3" Ai′ (ξ) + iGi′(ξ)# − (3b)1/3 ≈(cid:16) ω a ∂a 2Ω(cid:17)2/3 2(1 − n) + γ−2 + (θ − α)2 [cos(φ − ψ0) sin θ sin α]1/3 , ∂Φ(a, b) (9) , (10) 0 0 (12) dτ ei(aτ +bτ 3), Φ(a, b) =Z t0 where Gi(ξ) = (1/3)Bi(ξ)−R ξ [Ai(ξ′)Bi(ξ)− Ai(ξ)Bi(ξ′)]dξ′, Ai(ξ) and Bi(ξ) are the Airy functions (see Abramowitz & Stegun 1970). Similar to Ai(ξ) and −Ai′(ξ), both Gi(ξ) and −Gi′(ξ) are a decaying function for ξ > 0 (as shown in figure 1). The first terms in both square brackets on the right-hand sides of (9) and (10) correspond to synchrotron radiation for a particle moving along a semi-infinite trajec- tory. The terms Gi and Gi′ describe MIB arising from the combined effect of the semi-infinite track's initial point and curvature. One can show that MIB has features of usual im- pulsive bremsstrahlung (cf. Eq 16), i.e. emission due to an in- stantaneous change in the velocity from zero to v ∼ c at t = 0 (cf. Eq. 1) (Landau & Lifshitz 1971; Grichine 2003). One should emphasize here that the impulsive bremsstrahlung considered here is different from the usual bremsstrahlung by a charged particle interacting with the Coulomb field of nuclei in matter. It can be shown that the last terms (Φ and ∂Φ/∂a) on the right-hand side of both (9) and (10) can be ignored provided that φ − ψ0γ ≪ π(2Ωeγ2/ω)1/3. The characteristic frequency can be estimated from a3 ∼ b, which leads to ω ∼ γ2Ωe. The spectrum can be written as sum of usual syn- chrotron emission (Usyn) from a semi-finite track and MIB emission (Ub) due to the velocity jump at t = 0 and the trajectory's curvature, that is, U = Usyn + Ub, (13) with Usyn ≈ e2 16π3ε0c(cid:16) ω a(cid:17)2"(θ − α)2(cid:16)πξAi(ξ)(cid:17)2 Ub ≈ (11) e2 (cid:16)πξ2Ai′(ξ)(cid:17)2#, aγ (cid:19)2 +(cid:18) Ωe sin α a(cid:17)2"(θ − α)2(cid:16)πξGi(ξ)(cid:17)2 16π3ε0c(cid:16) ω aγ (cid:19)2 +(cid:18) Ωe sin α (cid:16)πξ2Gi′(ξ)(cid:17)2#, (14) (15) where we neglect both the term (12) and its derivative. It is often convenient to separate the radiation into two components, with one polarized perpendicular to the orbit plane, corresponding to the first terms in (14) and (15), and the other polarized in the plane, corresponding to the sec- ond terms involving derivatives. Specifically, one may write Usyn = Usyn⊥ + Usynk and Ub = Ub⊥ + Ubk. Notice that Usyn is smaller by a factor of 4 than that for normal syn- chrotron emission from a full infinite track (cf. Eq. 17). Tak- ing the limit Ωe → 0 one can easily verify that Ub reduces to the familar form for the prompt bremsstrahlung in the case where a particle's velocity abruptly changes from zero to a constant velocity (Landau & Lifshitz 1971). In this limit, the angle α is re-interpreted as an angle that the velocity makes with respect to the z axis. Since one has ξ → ∞ and hence πξGi(ξ) → 1 (Abramowitz & Stegun 1970), one finds Ub(ω, k) ≈ e2 4π3ε0c(cid:20) θ − α γ−2 + (θ − α)2(cid:21)2 , (16) which is the same as that for the case where a particle in- staneously acquires a constant velocity (Landau & Lifshitz 1971). Apart the usual from its property of impulsive bremsstrahlung, MIB also has a feature of synchrotron ra- diation (cf. Sec. 2.3), i.e. Ub increases with frequency as a power-law, similar to the usual synchrotron spectrum at low frequencies. Such similarity can be understood as due to that the emitting particle's trajectory is curved, while for the usual impulsive bresstrahlung the particle's trajectory is a straight line. Because of this similar feature to synchrotron radiation MIB is nonzero when one applies it to pair cre- ation. 4 Luo 7 6 5 4 3 2 1 0 -10 -5 0 (θ−α)γ 5 10 2.5 2 1.5 1 0.5 0 -10 -5 0 (θ−α)γ 5 10 Figure 2. Plot Uk (arbitrary scale) as a function of (θ − α)γ with φ = ψ0 for a single particle. The dashed line corresponds to the usual synchrotron emission with Uk = 4Usynk, in the absence of MIB emission. We assume γ = 80, n = 1, the gyrofrequency Ωe = 5 × 106 s−1, ω/2π = 100 MHz and the pitch angle α = π/4. Figure 3. Plot U⊥ (arbitrary scale) as a function of (θ − α)γ with φ = ψ0. The parameters are as in figure 2. Apart from a reduction in intensity, the modified synchrotron emission (usual synchrotron plus MIB) has a much wider profile than the usual synchrotron emission. 2.3 Usual synchrotron radiation The conventional synchrotron emission can be reproduced by adding the other semi-infinite orbit from −∞ to 0 in (1). Specifically, one first drops the terms Φ in (9) and ∂Φ/∂a in (10), both of which are cancelled out by the corresponding terms from the negative semi-infinite trajec- tory, and then replaces F and F ′ in (8) with 2Re(F ) and 2Re(F ′), respectively. Thus, the bremsstrahlung terms can- cel out. One then reproduces the usual synchrotron for- mula (Melrose & McPhedron 1991) 1.75 1.5 1.25 1 0.75 0.5 0.25 U (ω, k) = 4Usyn e2 ≈ 6π3ε0c(cid:16) ω ×" (θ − α)2 γ−2 + (θ − α)2 K2 Ω sin α(cid:17)2(cid:20)γ−2 + (θ − α)2(cid:21)2 2/3(ρ)#, 1/3(ρ) + K2 -7.5 -5 -2.5 2.5 5 7.5 0 (θ−α)γ Figure 4. Angular profile (arbitrary scale) as in figure 2. The solid and dashed lines correspond respectively to Ubk and Usynk. (17) where n = 1 and ρ = (2/3)ξ3/2. We rewrite the Airy func- tions in terms of the modified Bessel functions and the ap- proximations θ − α ≪ 1 and φ − ψ0 ≪ 1 are used. The angular distribution of single particle's spectrum (13) is shown in figures 2-5. Here, the radiation is separated into the parallel component, Uk = Usynk + Ubk and perpen- dicular component, U⊥ = Usyn⊥ + Ub⊥. The inclusion of the boundary effect leads to an overall reduction in intensity and a significant broadening of the angular profile. Figures 6 and 7 show a comparison of the two components Ub and Usyn as a function of frequency. These two components are similar to each other at low frequencies. However, Ubk has two cutoffs, with the lower one determined from the zero of Gi′(ξ) (cf. figure 1). For the perpendicular polarization, Ub⊥ levels out at high frequencies and behaves much like the usual impul- sive bremsstrahlung. The energy spectrum for the parallel polarization is shown in figure 8. The synchrotron emission drops off exponentially above the critical frequency. Since Gi(ξ) ∼ 1/ξ and Gi′(ξ) ∼ 1/ξ2, which drop off much slower than the exponential decay of Ai(ξ) and Ai′(ξ) at a large ξ, the bremsstrahlung component decays much slower than the usual synchrotron emission. 1.75 1.5 1.25 1 0.75 0.5 0.25 0 -10 -5 0 (θ−α)γ 5 10 Figure 5. Angular profile (arbitrary scale) as in figure 2. The solid and dashed lines correspond respectively to Ub⊥ and Usyn⊥. Coherent synchrotron emission from cosmic ray air showers 5 U2π b , 2π Usyn -33 10 -34 10 -35 10 U2π -33 10 -34 10 -35 10 8 10 9 10 ω/2π (Hz) 10 10 8 10 9 10 10 10 ω/2π (Hz) Figure 6. Energy spectrum (in J Hz−1 sr−1) as a function of frequency ω/2π (Hz) for φ = ψ0 for a single particle. The solid and dashed lines represent 2πUbk and 2πUsynk, respectively. We assume (θ − α)γ = 0.5, N = 107, γ = 80, n = 1, Ωe = 5 × 106 s−1, (θ − α)γ = 0.5, and ψ = ψ0. At low frequencies, the spectrum is very similar to that for the usual synchrotron, i.e. U ∝ ω2/3 for ω ≪ ωc ∼ Ωeγ2. Figure 8. Energy spectrum 2πUk = 2π(Ubk + Usynk) (in J Hz−1 sr−1). The dashed line corresponds to synchrotron emis- sion 8πUsynk. The parameters are as in figure 6. In comparison with the normal synchrotron emission (dashed line), due to the MIB the spectrum (solid line) drops off much slower than the exponential decay at ω ≫ ωc. U2π b 2πUsyn , -33 10 -34 10 -35 10 8 10 9 10 10 10 ω/2π (Hz) Figure 7. Energy spectrum as in figure 6 for the ⊥-polarized components. Notice that the spectral component 2πUb⊥ (solid line) tends to level out at high frequencies where the approximate form (16) applies. 2.4 Many-particle system The single particle formula can be extended to a many- particle system by adding all the currents from individ- ual particles. Let the sth particle be created at time ts. A nonzero ts adds a phase term ωts to (5). The total current can be obtained summing up all individual currents given by (6) with all relevant quantities labelled by s. Then, the total energy spectrum is derived as Utot(ω, k) ≈ e2ω2 16π3ε0c + Ω2 e γsγs′ N Xs,s′=1(cid:20)ηsηs′ (αs − θ)(αs′ − θ)FsF ∗ s′ sin αs sin αs′ F ′ sF ′∗ ′ , (18) s′(cid:21) eiϕss where N is the total number of charged particles. The co- herence is determined by the phase ϕss′ = ds − ds′ , given by ϕss′ ≈ ω(ts − ts′ ) − k · (x0s − x0s′ ) − + ω Ωs ω Ωs′ [1 − nβs cos(θ − αs)] (φ − ψ0s) [1 − nβs′ cos(θ − αs′ )] (φ − ψ0s′ ). (19) We use the following expression for the initial position: x0 = x0c − (v⊥/Ω)(sin ψ0 ex − cos ψ0 ey). The usual sponta- neous emission corresponds to that the phase is ϕss′ ≫ 1 for s 6= s′ and that only terms of s = s′ contribute to the total spectrum. Therefore in the case of spontaneous syn- chrotron radiation, the total spectrum can be written as Utot = N ¯U , where ¯U is the single particle's spectrum aver- aged over the particles' momentum distribution. Since the initial gyrophase does not enter the final form of the energy spectrum, spontaneous synchrotron radiation is axially sym- metric with respect to the magnetic field line direction, i.e. the angular pattern of emission depends only on θ not φ. In the case of coherent synchrotron emission (cf. Sec. 3), such symmetry is broken since the maximum coherence depends explicitly on the initial gyrophases (cf. Eq. 19). 3 COHERENT SYNCHROTRON EMISSION Coherent emission occurs provided that the majority of emitting particles satisfy the condition ϕs − ϕs′ ≪ 1. In general, one can calculate the total spectrum numerically using (18) for a given distribution of the particle injection time, position and initial velocity. In some special cases, one may write down its analytical form. The simplest case is where all particles have the same initial velocity, which is considered here. One may write the total spectrum into the form 6 Luo (20) Utot = NhS⊥(ω)U⊥ + Sk(ω)Uki, where 1 6 Sk,⊥(ω) 6 N is called the coherence factor. The value Sk,⊥(ω) = N corresponds to completely coherent emission and Sk,⊥(ω) = 1 to spontaneous emission. When the number densities of electrons and positrons are equal, j⊥θ does not contribute to the coherent power as contribu- tions from electrons and positrons cancel out. This leads to S⊥ = 1. The total spectrum is e2 Utot = NhU⊥ + Sk(ω)Uki, 16πε0c sin α(cid:16) 4ω Uk ≈ ×(cid:20)Ai′2(ξ) + Gi′2(ξ)(cid:21), Ωe sin α(cid:17)2/3 (21) (22) where ξ ≈ (ω/2Ω sin α)2/3[γ−2 + (θ − α)2] and n = 1 is assumed. In contrast to coherent Cerenkov emission, which requires a charge asymmetry (net charge) (Askar'yan 1962, 1965), coherent synchrotron emission can occur for a neutral plasma. In the case of charge symmetry, the polarization is linear, perpendicular to the plane of the magnetic field and wave vector. The coherence factor is then given by Sk(ω) = 1 N N Xs,s′=1 cos [ω(ts − ts′ ) − k · (x0s − x0s′ )] . (23) Eq. (23) is sum of phasors, which can be modeled as (e.g. Hartemann 2000) Sk(ω) ≈ 1 − hcos Θi2 + Nhcos Θi2 +2hcos ΘihN (1 − hcos Θi2)i1/2 , (24) where N ≫ 1, Θ = ω(ts − n · x0s/c), n = kc/ω, and the average is made over a distribution P (Θ), hcos Θi ≡Z 2π 0 dΘ P (Θ) cos Θ. (25) For a uniform distribution P (Θ) = 1, corresponding to spon- taneous emission, one has Sk(ω) = 1, and for P (Θ) = δ(Θ), one has Sk(ω) = N , corresponding to completely coherent emission. As an example, we consider the case in which par- ticles are injected at the same time, say at ts = 0, with a gaussian profile in the longitudinal (along k) spatial dis- tribution with a width ∆l. The probability can be written as P (Θ) = 1 √π∆Θ e−(Θ/∆Θ)2 , (26) where ∆Θ = ω∆l/c. Then, one finds hcos Θi = exp(−∆Θ2/2) = exp[−(ω∆l/c)2/2]. A plot of Sk as a func- tion of ∆l is shown in figure 9. Similarly, if all particles are injected at the same location with a gaussian profile with a width ∆t in time, one has hcos Θi = exp[−(ω∆t)2/2]. This shows that to attain effective coherence, the spread in time of particle injection must be ∆t ∼ √2/ω. The spectrum from a gaussian bunch is calculated using (20) and is shown in figure 10. The spectrum consists of two regions, separated by a critical frequency ωcoh ∼ (2 ln N )1/2c/∆l: the coherent (Sk > 1) emission region ω < ωcoh, where the polarization is predominantly linear, and spontaneous (Sk = 1) emission S 7 10 6 10 5 10 4 10 3 10 2 10 10 1 0 2 4 6 8 10 l /λ∆ Figure 9. Coherence factor vs bunch size. The solid and dashed lines correspond respectively to a gaussian bunch (26) and a bunch of Γ-form (27) with A = 5. One assumes the total number of emitting particles N = 107. The bunch size is in units of the wavelength λ. region ω > ωcoh, where the polarization is elliptical as both Uk and U⊥ are present. The calculation of the coherence factor can easily be extended to other types of distribution. In particular, the Γ-probability distribution is thought to be the more relevant for air showers (Agnetta et al. 1997), which is discussed in detail in Sec. 4. As the secondary particles from air showers have a dis- tribution in momentum, one needs to use the full expression (18) to calculate the spectrum. When the particle's momen- tum distribution is included, an analytical form for the spec- trum can be derived only in some special cases. For example, when the bunch size is smaller than the relevant wavelength, one may ignore the phase term (19) and expresses the spec- trum as a integration of the single particle formula over the particle's distribution. The total intensity is about N 2 times larger than that by single particle. 4 APPLICATION TO AIR SHOWERS The formalism developed in the previous sections can be applied to radio emission from extensive air show- ers (EAS). Detailed modeling requires numerical model- ing of EAS, which was already considered by several au- thors (Suprun, Gorham & Rosner 2003; Huege & Falcke 2005a,b). Here we only discuss qualitatively implications of the modified synchrotron emission for radio emission from EAS and further detailed modeling will be considered else- where. A notable feature of the modified synchrotron emis- sion is that the corresponding intensity derived from a semi- inifinite orbit is smaller than the conventional synchrotron emission by about a factor of 4. In principle, this feature can be tested against observations provided accurate radio spectra with well calibrated intensities become available, for example with the planned Low Frequency Array (LO- FAR) (Falcke & Gorham 2003). Other features include an angular broadening of the radiation pattern and a slower de- cay of the spectrum above the critical frequency (ω ∼ Ωeγ2). The broadening due to prompt bremstrahlung is more pro- Coherent synchrotron emission from cosmic ray air showers 7 U2π tot -20 10 -22 10 -24 10 -26 10 -28 10 8 10 9 10 ω/2π (Hz) 10 10 Figure 10. Spectrum 2πUtot (in J Hz−1 sr−1) from a gaus- sian bunch (solid line) with a longitudinal size ∆l = 1 m and the Γ-pdf bunch (long-dashed line) with A = 3. The short- dashed line corresponds to the spectrum derived by ignoring the prompt bremstrahlung. The parameters are as in figure 6. There is a critical frequency ωcoh that separates the spectrum into the coherent region ω < ωcoh and spontaneous emission re- gion ω > ωcoh. For the gaussian bunch, the critical frequency is ωcoh/2π ∼ (2 ln N )1/2c/2π∆l ∼ 270 MHz and for the Γ-pdf bunch it is ωcoh/2π ∼ (c/∆l)N 1/(1+A) ≈ 1 GHz. nounced for the perpendicular polarization than for the par- allel polarization. Since the perpendicularly polarized com- ponent depends on the charge excess, such broadening is important when there is significant charge asymmetry. The spectral hardening occurs near the synchrotron cut-off fre- quency, which is much higher than the transition frequency ωcoh. So, this spectral feature may not be observable for air showers. In the following we derive the coherence factor using the procedure in Sec. 3. As an example, one assumes that an air shower occurs at a few km height and that all emit- ting particles are located in the shower maximum. As the zeroth order approximation, the near field effect can be ig- nored and the energy spectrum (Eq. 18) derived in Sec. 2 is applicable. The near field effect may need to be included if the maximum of an air shower develops very near the ground level. There are extensive discussions of air showers in the literature (Gaisser 1990) and in principle, one can obtain a distribution of particle's injection time (ts), initial momen- tum (γv0s) and position (x0s). Assuming the primary cos- mic ray energy to be Ep, the number of secondary electrons and positrons can be estimated as N ∼ Ep/γmec2 ∼ 107 for Ep ∼ 1015 eV and γ = 80. In the practical situation these particles are injected over an extended range rather than at a single fixed height. For EAS, the distribution in ts for secondary electrons/positrons can be inferred from mea- surements of arrival times of charged particles (muons plus electrons/positrons). One should note that these measured times are not actual arrival times of electrons and positrons since muons arrive earlier than electrons/positrons. The ar- rival times can be modeled by a Γ-probability distribution function (Γ-pdf) (Agnetta et al. 1997; Antoni, et al. 2001). Thus, the probability can be written as (Huege & Falcke 2003): P (Θ) = 1 ∆ΘΓ(1 + A) (cid:16) Θ ∆Θ(cid:17)A e−Θ/∆Θ, (27) for Θ > 0 and P (Θ) = 0 for Θ 6 0, where Γ(x) is the Gamma function and the power index A can be estimated from the distribution of electrons arrival times. The standard devia- tion is σΓ = (1 + A)1/2∆Θ. From (27), one obtains hcos Θi =(cid:0)1 + ∆Θ2(cid:1)−(1+A)/2 cos(cid:20)(1 + A) arctan ∆Θ(cid:21), (28) hΘi = (1 + A)∆Θ. The corresponding critical frequency is given by ωcoh ∼ (c/∆l)N 1/(1+A). This frequency defines a transition from coherent to incoherent emission. For ∆l = 1 m, N = 107 and A = 5, one has ωcoh/2π ∼ 1.2 GHz. The coherence fac- tor can be obtained by substituting (28) for (24). Here we write down the two limiting cases: (29) Sk ≈  N (1 + ∆Θ2)−1−A × cos2(cid:20)(1 + A) arctan ∆Θ(cid:21), ω < ωcoh, ω ≫ ωcoh. 1, (30) The dashed line in figure 9 shows Sk as a function of the bunch size ∆l, featuring a higher transition frequency ωcoh than the gaussian bunch. The low-frequency approx- imation in (30) corresponds to the nonsquared form given by Huege & Falcke (2003) except that the cosine factor is retained here. The coherent emission can be elliptically polarized if there is a charge excess as it is the case for air showers. Assuming the excess number of electrons is Nc, one has . (31) Nc N S⊥ Sk ∼ The right-hand side is sensitive to the cut-off energy of the excess electron energy. For a cut-off near the MeV energy, the excess of electrons can reach Nc/N ∼ 10 − 15%. (Note that cascades in a dense medium such as rocks can give rise to about Nc/N ∼ 20% excess negative charge.) As a result, the coherent emission can be elliptically polarized with an ellipticity ∼ (S⊥/Sk)1/2δc = (Nc/N )2δc ∼ 0.3δc, where δc is the typical ellipticity of single particle's radiation. 5 CONCLUSIONS Synchrotron emission by a particle moving along a semi- infinite trajectory is considered. Since effective coherent syn- chrotron emission by secondary particles in an air shower occurs in the core of the shower where most emitting parti- cles are created, the initial point of the particle's trajectory need be included explicitly. It is shown that radiation from a particle moving along a semi-infinite track can be separated into the usual synchrotron emission and the bremsstrahlung- like emission (MIB). The latter is due to emission as the result of onset of the particle's curved trajectory. The spec- tral intensity of the modified synchrotron emission (usual synchrotron plus MIB) is lower than that for normal syn- chrotron emission, roughly by a factor of 4. It is interest- ing to note that such reduction is consistent with the re- cent result from numerical simulation by Huege & Falcke Dagkesamanskii, R. D. & Zheleznykh, M., 1989, JETP Lett. 50, 259. Falcke, H. et al. 2005, Nature, 435, 313. Falcke, H. & Gorham, P., Astrop. Phys. 19, 477. Falcke, H., Gorham, P., & Protheroe, R. J., 2004, New Astron Rev. 48, 1487. Gaisser, T. K., 1990, Cosmic Rays and Particle Physics (Cambridge: University Press). Gorham, P. W., et al., 2004, Phys. Rev. Lett. 93, 041101. Grichine, V. M., 2003, Radiation Phy. & Chem. 67, 93. Hankins, T., Ekers, R., & O'Sullivan, J. D., 1996, MNRAS, 283, 1027. Hartemann, F., 2000, Phys. Rev. E61, 972. Huege, T. & Falcke, H., 2005, Astroparticle Phys. 24, 116. Huege, T. & Falcke, H., 2005, A&A, 430, 779. Huege, T. & Falcke, H., 2003, A&A, 412, 19. Jackson, J. D., 1998, Classical Electrodynamics (New York: Wiley). Kahn, F. D. & Lerche, I., 1966, Proc. R. Soc. London, A289, 206. Landau, L. D. & Lifshitz, E. M., 1971, The Classical Theory of Fields (Oxford: Pergamon Press). Melrose, D. B., 1986, Instabilities in Space and Laboratory Plasmas, Cambridge University Press. Melrose, D. B., McPhedron, R. C., 1991, Electromagnetic processes in dispersive media, Cambridge U. press. Suprun, D. A., Gorham, P. W., & Rosner, J. L., 2003, As- troparticle Phys. 20, 157. Tamm, I. E., 1939, J. Phys. (Moscow) 1, 439. White, S. M. & Melrose, D. B., 1982, PASA 4, 362. Zas, E., Halzen, F., & Stanev, T., 1992, Phys. Rev. D45, 362. 8 Luo (2005b). The radiation pattern has a broader angular dis- tribution than the usual synchrotron emission. This feature is especially pronounced for the perpendicularly polarized component. The spectrum has a much slower decay above the critical frequency ω ∼ γ2Ωe, while the usual synchrotron spectrum has an exponential cutoff above the critical fre- quency. In the application to radio emission from air show- ers, the reduced intensity can be verified in principle pro- vided accurate observations of the radio spectrum are avail- able. Although there exist early observations of air shower radio emission, there are some uncertainties in determining the calibration factor (Allan 1971; Atrashkevich et al 1978). The current LOFAR Prototype Station (LOPES) and the future LOFAR may provide a better opportunity to test the predicted spectrum. Since the transition frequency (to spon- taneous emission) is much lower than the cut-off frequency, change near the cut-off frequency may not be observable. The broadening of the radiation pattern occurs mainly for the perpendicularly polarized component, which may be ob- servable provided that there is significant charge asymmetry in the emitting plasma. A major advantage of the formalism presented here is that the initial conditions including the time of parti- cle creation, initial velocity and gyrophase all appear in the phase (19) and these quantities can be modeled sta- tistically. Although the single-particle formalism was used by White & Melrose (1982) to treat coherent gyromagnetic emission, in their calculation, the finite track effect was not considered. It is shown here that the distribution of the par- ticle injection time is important in determining the coher- ence. The semi-infinite track approximation adopted here is valid provided that the emitting particles are highly rela- tivistic with γ ≫ 1. In the relativistic limit, the synchrotron pulse duration (1/Ωeγ2 ∼ 10−8 s) is much shorter than the typical flight time T ∼ 10−6 s for a particle's free path 500 m, and therefore the orbit can be approximately regarded as semi-infinite. If electrons and positrons from an air shower have an energy cutoff extending to MeV energies with γ ∼ 1, the synchrotron approximation is no longer valid and one has cyclotron emission instead. In this case, a finite orbit needs to be considered. One may extend the calculation in Sec. 3 and 4 to include the particle distribution in momentum and this requires a numerical approach, which is not considered here. REFERENCES Abramowitz, M., Stegun, I., 1991, Handbook of Mathemat- ical Functions (New York: Dover Publications). Agnetta, G. et al. 1997, Astroparticle Phys. 6, 301. Allan, H. R., 1971, in Progress in Elementary Particles and Cosmic Ray Physics, Vol. 10, p. 171. Aloisio, R. & Blasi, P. 2002, Astroparticle Phys. 18, 183. Alvarez-Muniz, J., V´azquez, R. A., & Zas, E., 2000, Phys. Rev. D62, 063001. Antoni, T. et al. 2001, Astroparticle Phys. 14, 245. Askar'yan, G. A. 1962, Sov. Phys. JETP, 14, 441. Askar'yan, G. A. 1965, Sov. Phys. JETP, 21, 658. Atrashkevich, V. B. et al., 1978, Sov. J. Nucl. Phys. 28, 366. Colgate, S. A., 1967, J. Geophys. Res. 72, 4869.
astro-ph/0110477
1
0110
2001-10-22T08:46:43
RR Lyrae variables in the dwarf spheroidal galaxy Leo I
[ "astro-ph" ]
We report the discovery of a significant population of RR Lyrae variables in the dwarf spheroidal galaxy Leo I. Based on 40 V and 22 B images of the galaxy taken using the ESO Wide Field Imager we have identified so far 74 candidate RR Lyrae's in two CCD's hosting the main body of the galaxy. Full coverage of the light variations and pulsation periods have been obtained for 54 of them, 47 of which are Bailey {\it ab}-type RR Lyrae's (RRab's) and 7 are {\it c}-type (RRc's). The period distribution of the presently confirmed sample of RRab's peaks at P=0\fd60, with a minimum period of 0\fd54. The pulsational properties indicate for Leo I an intermediate Oosterhoff type, similar to other dwarf galaxies in the Local Group and the LMC. However, the rather long minimum period of the {\it ab}-type variables, and the significant number of RRab's with long period and large amplitude, suggest that the bulk of the old population in Leo I is more like the Oosterhoff type II globular clusters. The most straightforward interpretation is that a range in metallicity is present among the RR Lyrae's of Leo I, with a significant population of very metal-poor stars. Alternatively, these OoII variables could be more evolved. The average apparent magnitude of the RR Lyrae's across the full cycle is $<V(RR)>= 22.60 \pm 0.12$ mag, yielding a distance modulus $(m-M)_{V,0}= 22.04\pm 0.14$ mag for Leo I on the ``long'' distance scale.
astro-ph
astro-ph
Preprint typeset using LATEX style emulateapj v. 19/02/01 RR LYRAE VARIABLES IN THE DWARF SPHEROIDAL GALAXY LEO I 1 Osservatorio Astronomico di Padova, Vicolo dell'Osservatorio 5, I-35122 Padova, Italy [email protected] ENRICO V. HELD GISELLA CLEMENTINI Osservatorio Astronomico di Bologna, Via Ranzani 1, I-40127 Bologna, Italy [email protected] LUCA RIZZI2 Osservatorio Astronomico di Padova, Vicolo dell'Osservatorio 5, I-35122 Padova, Italy [email protected] YAZAN MOMANY Dipartimento di Astronomia, Università di Padova, Vicolo dell'Osservatorio 2, I-35122 Padova, Italy [email protected] IVO SAVIANE European Southern Observatory, Casilla 19001, Santiago 19, Chile [email protected] AND LUCA DI FABRIZIO Centro Galileo Galilei & Telescopio Nazionale Galileo, PO Box 565, 38700 S.Cruz de La Palma, Spain [email protected] ABSTRACT We report the discovery of a significant population of RR Lyrae variables in the dwarf spheroidal galaxy Leo I. Based on 40 V and 22 B images of the galaxy taken using the ESO Wide Field Imager we have identified so far 74 candidate RR Lyrae's in two CCD's hosting the main body of the galaxy. Full coverage of the light variations and pulsation periods have been obtained for 54 of them, 47 of which are Bailey ab-type RR Lyrae's (RRab's) and 7 are c-type (RRc's). The period distribution of the presently confirmed sample of RRab's peaks at P=0.d60, with a minimum period of 0.d54. The pulsational properties indicate for Leo I an intermediate Oosterhoff type, similar to other dwarf galaxies in the Local Group and the LMC. However, the rather long minimum period of the ab-type variables, and the significant number of RRab's with long period and large amplitude, suggest that the bulk of the old population in Leo I is more like the Oosterhoff type II globular clusters. The most straightforward interpretation is that a range in metallicity is present among the RR Lyrae's of Leo I, with a significant population of very metal-poor stars. Alternatively, these OoII variables could be more evolved. The average apparent magnitude of the RR Lyrae's across the full cycle is hV (RR)i = 22.60 ± 0.12 mag, yielding a distance modulus (m - M)V,0 = 22.04 ± 0.14 mag for Leo I on the "long" distance scale. Subject headings: Galaxies: individual (Leo I) -- galaxies: dwarf -- galaxies: stellar content -- Local Group -- stars: horizontal-branch -- stars: variables: other 1. INTRODUCTION RR Lyrae have long been recognized to be excellent tracers of old stellar populations, as well as good distance indicators for Population II systems. RR Lyrae variables and/or extended horizontal branches (HB's) are now known to exist in all Local Group dwarf spheroidal (dSph) galaxies, despite the different and complex star formation histories (Mateo 1998; Da Costa 1998). The Leo I dSph was believed to represent the only ex- ception to this general trend, with its predominantly "young" population, and a wealth of intermediate age He-burning stars giving rise to a conspicuous red clump (Lee et al. 1993; Ca- puto et al. 1998; Gallart et al. 1999a, 1999b). Recently, color- magnitude diagrams (CMD's) extended to the outer regions have revealed a horizontal branch well populated from blue to red (Held et al. 2000). Given the presence of an extended HB and the low metallicity of the system ([Fe/H]∼ - 2, Lee et al. 1993) one would naturally expect to find RR Lyrae variables in Leo I. Indeed, a few stars near the limit of photographic photometry were mentioned by Hodge & Wright (1978) as probable RR Lyrae's caught at maximum light. Unpublished work of Keane et al. (1993) also reports the presence of these variables. To confirm these early suggestions, a search for RR Lyrae and other short period variables in Leo I was undertaken using the CCD mosaic imager at the 2.2m ESO-MPI telescope. In this letter we report on the successful detection of a conspicu- ous number (& 70) of RR Lyrae variables in this galaxy. The identification of RR Lyrae's in Leo I and the determination of their average luminosity and pulsational properties provide us with important tools for studying the metallicity and age of the oldest stellar population and the early star formation history of Leo I, by allowing at the same time a new independent mea- surement of the distance to this galaxy. 1 Based on data collected at E.S.O. La Silla, Chile, Prop. No. 65.N-0530 2 also Dipartimento di Astronomia, Università di Padova 1 2 2. OBSERVATIONS AND REDUCTION Images of Leo I (α2000=10 08 27, δ2000=12 18 30) through the B, V , I filters were obtained on the nights of April 20-25, 2000 with the Wide Field Imager (WFI) at the 2.2 m ESO-MPI telescope at La Silla, Chile. The WFI camera consists of 8 "ccd 44"-type EEV CCD's with a total field of view 34′ × 33′. This coverage allowed us to observe the body of the galaxy and its outer regions in just one pointing. The main body of the galaxy was centered on the lower four CCD's of the mosaic to reduce light contamination from Regulus, which is about 20 ar- cmin south. The full data-set comprises 22 B, 40 V , and 4 I 15-min exposures totaling ∼ 20 hours of observing time spread over 6 consecutive half nights. Weather conditions were gen- erally photometric, while the seeing varied considerably during the observations. Standard stars from Landolt (1992) were ob- served during all photometric nights of the run in all CCD's. Basic reduction (flat fielding, registration, co-adding) of the CCD mosaic data were performed using the IRAF 3 package MSCRED (Valdes 1998) and the pipeline script package WF- PRED developed at the Padua Observatory. Reductions were done, and data calibrated, for the individual CCD's, yielding calibration uncertainties of 0.03 mag in B and 0.04 mag in V (night-to-night rms scatter of the zero points). For variable stars, the calibration were applied only to average magnitudes and colors given the relatively large color terms of the WFI camera (about 0.28 and - 0.075 in b and v, respectively). The ALLFRAME program (Stetson 1994) was used to measure all images and combine the individual photometric measurements into a master catalog for each CCD. The combined photometry reaches V ≃ 25.5 (about 3 mag fainter than the HB) with errors σV = σB = 0.2 mag (standard error of the mean). The limiting magnitude for the individual frames varies between V = 23.0 and 24.5 depending on the seeing and photometric conditions. Measurement errors vary accordingly between σV = 0.12 -- 0.35 at V ∼ 23.5. The measurement errors for stars at the HB level, estimated from the standard deviations of the measurements for non-variable stars, are about σV (HB) = 0.13 and σB(HB) = 0.11 on the individual frames. 3. RR LYRAE VARIABLE STARS IN LEO I The time-series photometric measurements in the CCD cat- alogs were analyzed to detect variable objects and derive their light curves. Here we report on the first successful results for two CCD's (named #6 and #7 in the ESO WFI Manual) com- prising the body of the galaxy (Leo I was centered on CCD #7). All results for the remaining CCD's, together with photometric measurements and light curves, will be presented in a separate paper (G. Clementini et al. 2001, in preparation). Identifica- tion of the variables was performed independently on the instru- mental v and b data sets, running the program VARFIND (devel- oped by Dr. P. Montegriffo) on the photometric catalogs (see Clementini et al. 2001 for details). At this stage, the search was optimized to find candidate variables near the HB. Stars whose standard deviations of the v and b measurements are larger than 3σ, where σ is the rms of non-variable stars at the HB level, were flagged as candidate variables. As a result, 96 candidates were found, some of which are in the innermost area of Leo I. RR Lyrae's allow us to probe the old population all the way to the center of the galaxy. Period search and definition of the light curve of the sus- pected variables were done using differential photometry in the instrumental magnitude system, with respect to some carefully chosen, photometrically stable reference star (in general, red giants with B - V ≥ 0.6). Period search was performed us- ing GRATIS (GRaphical Analyzer of TIme Series), a software developed at the Bologna Observatory (see Clementini et al. 2000, 2001). FIG. 1. -- Differential v and b light curves of ab- and c-type RR Lyrae's in Leo I from data in CCD #6 (left panel) and #7 (right panel). Different symbols are used for data taken in different nights. Figure 1 presents the first light curves of RR Lyrae stars in Leo I. We have plotted the differential v and b light variations for a selection of RR Lyrae stars that sample quite well the typ- ical periods. Out of 96 candidate variables, the majority were found to be ab-type RR Lyrae's (63 stars), with a few exam- ples of c-type RR Lyrae's (11 objects). We also identified one probable anomalous Cepheid and 6 candidate short period bi- naries, while the remaining 15 objects could not be classified unambiguously. Some low amplitude variables, in particular c-type RR Lyrae's, may actually have escaped detection. A deeper search is in progress using the Alard's (2000) Optimal Image Subtraction method (G. Clementini et al. 2001, in prep.). Periods, amplitudes, and epochs of maximum light were de- rived for a subset of 54 RR Lyrae's (47 RRab's and 7 RRc's) with complete light curves, using the GRATIS χ2 Fourier fit- ting routine (one harmonic was used for the c-type RR Lyrae's and two to five harmonics for the RRab's). The accuracy of the derived periods depends on the sampling of the light curves and is in general of the order 0.001 d. The amplitude uncer- tainties, estimated from the rms of the fit residuals, are ∼0.10 mag. Intensity-average differential magnitudes were obtained by integration over the entire pulsation cycle using the best fit- ting model curves. By adding the instrumental magnitudes of the reference stars, we obtained the mean b, v magnitudes of the RR Lyrae's, from which the mean B, V magnitudes were calculated via the calibration equations. 3 IRAF is distributed by the National Optical Astronomy Observatories, which are operated by the Association of Universities for Research in Astronomy, Inc., under cooperative agreement with the National Science Foundation. 3 FIG. 3. -- Period -- V amplitude relation for the RR Lyrae's in Leo I. Periods are in days. Filled and open symbols are RRab and RRc variables, respectively. The regression lines show the AV - log P relations for the ab-type variables in the globular clusters M 3 and M 15 (solid lines), M 2 (dotted line), ω Cen (dashed line), and NGC 6441 (long dashed line). Dots represent the amplitude- period distribution of RR Lyrae variables in Leo II (Siegel & Majewski 2000). Figure 3 also shows a comparison with the AV - logP dis- tribution of the RR Lyrae variables in the dSph galaxy Leo II (Siegel & Majewski 2000). The pulsational properties of the RR Lyrae stars in the two dwarf spheroidals appear quite sim- ilar. The characteristics of the ab-type RR Lyrae's qualify Leo I as a system intermediate between the OoI and OoII clus- ters, similar in this respect to other dwarf galaxies in the Local Group (Siegel & Majewski 2000; Cseresnjes 2001; and refer- ences therein) and to the LMC (Bono, Caputo, & Stellingw- erf 1994). However, both the rather long minimum period of RRab's and the presence of RRab's with long periods -- large amplitudes, indicate a larger similarity with the OoII GCs. The implications of such Oosterhoff intermediate properties for the old stellar populations in Leo I (and in other dSph's as well) rests on the usual interpretation of metallicity as the dom- inant parameter in determining the Oosterhoff dichotomy. In- deed, OoII clusters are generally very metal-poor while OoI clusters are of intermediate metallicity (e.g., Smith 1995). If this interpretation is correct, then the distribution of periods and amplitudes of Leo I RR Lyrae's implies a metallicity distribu- tion for the old population extending from values more metal- poor than [Fe/H]=- 2.15 dex (the metallicity of the OoII cluster M 15), to as metal-rich as [Fe/H]∼ - 1.6 (the metallicity of the OoI cluster M 3). Alternative interpretations may exist, though. There are clus- ters like M 2, ω Cen, NGC 6441, and NGC 6388 whose po- sitions in the period-amplitude diagram imply that their RR Lyrae's are brighter than expected for their metallicity (see Fig- ure 3). This observational evidence is interpreted as the Oost- erhoff dichotomy being caused by evolution off the zero-age horizontal branch (ZAHB), with the OoI variables being on the ZAHB and the OoII variables being more evolved (Clement & Shelton 1999, Clement & Rowe 2000, Lee & Carney 1999b). FIG. 2. -- HR diagram of Leo I from a master ALLFRAME combination of all data for CCD #6. This CMD is based on the mean calibrated B, V magnitudes of stars detected in both passbands. The RR Lyrae variables (filled circles) and a Cepheid (filled square) identified in this chip are plotted according to their intensity-averaged magnitudes and colors. Lines show the blue and red edges of the instability strip of the globular cluster M 3, from Corwin & Carney's (2001) photometry, reddened to the value appropriate for Leo I. The red edge has been corrected to account for the difference in metallicity with respect to Leo I (Walker 1998). The location of the detected variables in the HR diagram of the outer region of Leo I is shown in Figure 2. The vast majority of the data points for the RR Lyrae's are found on the HB in cor- respondence to the instability strip, and their mean magnitude defines the mean V apparent luminosity of the HB of Leo I. 4. PULSATIONAL PROPERTIES: CLUES TO THE OLD STELLAR POPULATION IN LEO I Figure 3 shows the relation between the period and the V amplitude for the RR Lyrae's in the present sample. Open symbols represent the c-type pulsators (RRc), filled symbols are RRab stars. The average period of the Leo I RRab vari- ables is hPabi=0.d602 (σ=0.d059), and the minimum period is 0.d539. The solid lines in Figure 3 represent the AV - logP relations defined by the ab-type variables with clean light curves in the globular clusters M 3 and M 15 ([Fe/H]=- 1.66 and [Fe/H]=- 2.15, respectively, on the Zinn & West's 1984 scale). These clusters can be considered the prototypes of Oost- erhoff type I (OoI) and II (OoII) globular clusters (GCs) in the Milky Way (Sandage, Katem, & Sandage 1981; Sandage 1993). Also shown are the AV - logP relations followed by the ab vari- ables in M2, ω Cen, and NGC 6441 ([Fe/H]=- 1.62, - 1.60, and - 0.5 dex). The data are from Clement & Shelton (1999) for M3, Bingham et al. (1984) and Silbermann & Smith (1995) for M15, Lee & Carney (1999a) for M 2, Clement & Shelton (1999) for ω Cen, and Pritzl et al. (2000) for NGC 6441. 4 Then, a difference in age would exist between the two Oost- erhoff groups. Whether this can be a general interpretation of the Oosterhoff phenomenon or these are just peculiar clusters is still a matter of debate. In this regard, it is interesting to note the presence in Leo I of a variable with very long period (star 201064 in Figure 3, P = 0.d889). Although its large ampli- tude may be in doubt since the variable falls near the center of Leo I where crowding is severe, and is possibly affected from straylight from Regulus, variables with long periods have been found in NGC 6441 and NGC 6388 (Pritzl et al. 2000). A good indication on the mean metallicity of RR Lyrae's in Leo I can be derived using the relation derived by Sandage (1993) between the average period of the RRab variables in the Milky Way GCs and their metallicity, loghPabi = - 0.092[Fe/H]- 0.389. By applying this relation to the mean pe- riod of RRab variables in Leo I, we obtain [Fe/H]hPabi = - 1.82 for the average metallicity of the old population. Previous val- ues of the metallicity of Leo I are quite uncertain, ranging from [Fe/H]=- 2 to - 1 (see Lee et al. 1993 and references therein). A recent comparison of the (V - I) color of the red giant branch (RGB) of Leo I with the fiducial sequences of Galactic GCs yields a mean metallicity [Fe/H]≃ - 1.8 (Y. Momany et al. 2001, in preparation). Since the bulk of the red giant stars in Leo I have intermediate age (between 2 and 7 Gyr: Caputo et al. 1998; Gallart et al. 1999b), this value should be corrected (made more metal-rich) by ∼ 0.2 dex (cf. Held, Saviane, & Mo- many 1999). The fact that the RR Lyrae's are only slightly more metal-poor than RGB stars seems to imply that metal enrich- ment (from the primordial gas composition to [Fe/H]∼ - 1.8) occurred early in the life of the galaxy, and modest enrichment took place in between the early burst that gave rise to the oldest population, and the main star formation episode occurred 4 ± 3 Gyr ago. 5. THE DISTANCE TO LEO I The mean magnitude of the RR Lyrae's in Leo I provides an independent method to estimate the distance to this galaxy with some degree of confidence. The average apparent luminosity of the RR Lyrae's with full coverage of the B and V light curves is hV (RR)i = 22.60 ± 0.12 mag (standard deviation of the mean, 48 stars). This dispersion is fully accounted for by the com- bined effects of photometric errors, metallicity distribution in the RR Lyrae's sample, and evolution off the ZAHB. Given the large distance of Leo I, the spread caused by the intrinsic depth of the galaxy is insignificant. The absolute magnitude of the RR Lyrae's is MV (RR) ≃ 0.50 at [Fe/H]= - 1.5 in the "long" distance scale, consis- tent with (m - M)RR 0 = 18.53 for the LMC (see Walker 1999, and references therein), and 0.75 mag in the "short" scale. To correct the absolute magnitude of the RR Lyrae's to the metallicity of the Leo I variables, we adopted the RR Lyrae luminosity -- metallicity dependence given by the rela- tion ∆MV (RR)/∆[Fe/H] = 0.2 mag/dex, which is supported by both the Baade-Wesselink analysis of Fernley et al. (1998) and recent results on the M 31 globular clusters (Corsi et al. 2000). For the metallicity of Leo I we assumed the value de- rived in this paper from the mean period of ab-type RR Lyrae's, [Fe/H]=- 1.82. A reddening E(B - V ) = 0.04 ± 0.02 mag was adopted following Schlegel, Finkbeiner, & Davis (1998). These assumptions yield a distance modulus (m- M)RR 0 = 22.04 ± 0.14 mag (d=256 kpc) in the long scale (and 21.79 mag or 228 kpc in the short scale). The quoted error includes the errors on the mean RR Lyrae magnitude, the calibration zero point error, and the adopted reddening uncertainty. This result is in good agree- ment with our recent distance determination from the RGB tip method on a different data set (Y. Momany et al. 2001, in prep.), yielding (m- M)tip 0 = 21.93 ± 0.16. The RGB tip method is based on the RR Lyrae luminosity-metallicity relation of Lee, Demarque, & Zinn (1990), yielding MV (RR) = 0.565 at [Fe/H]= - 1.5. Our new result is intermediate between previous values in the literature (see Lee et al. 1993). We thank P. Stetson for providing us his set of photometric programs, including ALLFRAME. We are grateful to M. Cate- lan for helpful discussions regarding the nature of the Oost- erhoff dichotomy, to C. Cacciari for useful comments on the manuscript, and to an anonymous referee for helpful remarks. This study was partially supported by the National projects "Processing of large-format astronomical data" (COFIN-98) and "Stellar Observables of Cosmological relevance" (COFIN- 00). REFERENCES Alard, C. 2000, A&AS, 144, 363 Bertelli, G., Bressan, A., Chiosi, C., Fagotto, F., & Nasi, E. 1994, A&AS, 106, Bingham, E. A., Cacciari, C., Dickens, R. J., & Fusi Pecci, F. 1984, MNRAS, Bono, G., Caputo, F., & Stellingwerf, R. F. 1994, ApJ, 423, 294 Caputo, F., Cassisi, S., Castellani, M., Marconi, G., & Santolamazza, P. 1998, Clement, C. M., & Shelton, I. 1999, ApJL, 515, 85 Clement, C. M., & Rowe, J. 2000, AJ, 120, 2579 Clementini, G. et al. 2000, AJ, 120, 2054 Clementini, G., Gratton, R. G., Bragaglia, A., Carretta, E., Di Fabrizio, L., & Maio, M. 2001, AJ, submitted (astro-ph/0007471) Corsi, C. E., Rich, M., Cacciari, C., Federici, L., & Fusi Pecci, F. 2000, in A Decade of HST Science, ed. M. Livio, STScI Publications, in press Corwin, T.M., Carney, B.W. 2001, AJ, in press Cseresnjes P. 2001, A&A, 375, 909 Da Costa, G.S. 1998, In Stellar Astrophysics for the Local Group, ed. A. Aparicio, A. Herrero, & F. Sanchez (Cambridge: Cambridge Univ. Press), 351 Fernley, J., Carney, B. W., Skillen, I., Cacciari, C., & Janes, K. 1998, MNRAS, 275 209, 765 AJ, 117, 2199 Gallart, C., et al. 1999a, ApJ, 514, 665 Gallart, C., Freedman, W. L., Aparicio, A., Bertelli, G., & Chiosi, C. 1999b, 293, L61 AJ, 118, 2245 Held, E. V., Saviane, I., & Momany, Y. 1999, A&A, 345, 747 Held, E. V., Saviane, I., Momany, Y., & Carraro, G. 2000, ApJL, 530, L85 Hodge, P. W., & Wright, F. W. 1978, AJ, 83, 228 Keane, M., Olszewski, E., Suntzeff, N., & Saha, A. 1993, AAS, 182, #32.03 Landolt, A. U. 1992, AJ, 104, 340 Lee, J.-W., & Carney, B. W. 1999a, AJ, 117, 2868 Lee, J.-W., & Carney, B. W. 1999b, AJ, 118, 1373 Lee, Y. W., Demarque, P., & Zinn, R. 1990, ApJ, 350, 155 Lee, M. G., Freedman, W., Mateo, M., Thompson, I., Roth, M., & Ruiz, M.-T. 1993, AJ, 106, 1420 Mateo, M. 1998, ARA&A, 36, 435 Pritzl, B., Smith, H. A., Catelan, M., & Sweigart, A. 2000, ApJL, 530, 41 Sandage, A. 1993, AJ, 106, 687 Sandage, A., Katem, B., & Sandage, M. 1981, ApJS, 46, 41 Schlegel, D. J., Finkbeiner, D. P., & Davis, M. 1998, ApJ, 500, 525 Siegel, M. H., & Majewski, S. R. 2000, AJ, 120, 284 Silbermann, N. A., & Smith, H. A. 1995, AJ, 110, 704 Smith, H. A. 1995, RR Lyrae stars, Cambridge Astrophysics Series, (Cambridge Univ. Press: Cambridge) Stetson, P. B. 1994, PASP, 106, 250 Valdes, F. G. 1998, in ASP Conf. Ser. 145: Astronomical Data Analysis Software and Systems VII, 7, ed. R. Albrecht, R. N. Hook, and H. A. Bushouse, (ASP, San Francisco), 53 Walker, A.R. 1998, AJ, 116, 220 Walker, A.R. 1999, In Post-Hipparcos Cosmic Candles, ed. A. Heck, & F. Caputo, (Dordrecht: Kluwer), p. 125 Zinn, R., & West, M. J. 1984, ApJS, 55, 45 5
astro-ph/0703757
1
0703
2007-03-29T13:23:00
Gamma-Rays from Dark Matter Mini-Spikes in M31
[ "astro-ph" ]
The existence of a population of wandering Intermediate Mass Black Holes (IMBHs) is a generic prediction of scenarios that seek to explain the formation of Supermassive Black Holes in terms of growth from massive seeds. The growth of IMBHs may lead to the formation of DM overdensities called "mini-spikes", recently proposed as ideal targets for indirect DM searches. Current ground-based gamma-ray experiments, however, cannot search for these objects due to their limited field of view, and it might be challenging to discriminate mini-spikes in the Milky Way from the many astrophysical sources that GLAST is expected to observe. We show here that gamma-ray experiments can effectively search for IMBHs in the nearby Andromeda galaxy (also known as M31), where mini-spikes would appear as a distribution of point-sources, isotropically distributed in a \thickapprox 3^{\circ} circle around the galactic center. For a neutralino-like DM candidate with a mass m_{\chi}=150 GeV, up to 20 sources would be detected with GLAST (at 5\sigma, in 2 months). With Air Cherenkov Telescopes such as MAGIC and VERITAS, up to 10 sources might be detected, provided that the mass of neutralino is in the TeV range or above.
astro-ph
astro-ph
Gamma-Rays from Dark Matter Mini-Spikes in M31 Mattia Fornasa1, Marco Taoso1, and Gianfranco Bertone1,2 1 INFN, Sezione di Padova, via Marzolo 8, Padova, 35131, Italy and 2 Institut d Astrophysique de Paris, UMR 7095-CNRS, Universit´e Pierre et Marie Curie, 98bis boulevard Arago, 75014 Paris, France ∗ The existence of a population of wandering Intermediate Mass Black Holes (IMBHs) is a generic prediction of scenarios that seek to explain the formation of Supermassive Black Holes in terms of growth from massive seeds. The growth of IMBHs may lead to the formation of DM overdensities called "mini-spikes", recently proposed as ideal targets for indirect DM searches. Current ground- based gamma-ray experiments, however, cannot search for these objects due to their limited field of view, and it might be challenging to discriminate mini-spikes in the Milky Way from the many as- trophysical sources that GLAST is expected to observe. We show here that gamma-ray experiments can effectively search for IMBHs in the nearby Andromeda galaxy (also known as M31), where mini-spikes would appear as a distribution of point-sources, isotropically distributed in a ≈ 3◦ circle around the galactic center. For a neutralino-like DM candidate with a mass mχ = 150 GeV, up to 20 sources would be detected with GLAST (at 5σ, in 2 months). With Air Cherenkov Telescopes such as MAGIC and VERITAS, up to 10 sources might be detected, provided that the mass of neutralino is in the TeV range or above. I. INTRODUCTION The nature of Dark Matter (DM) is, more than 70 years after its discovery, still an open problem. It is com- monly assumed that DM is made of Weakly Interact- ing Massive Particles (WIMPs), arising in theories be- yond the Standard Model (see Refs. [1, 2] for recent re- views), the most widely discussed DM candidates being the supersymmetric neutralino and the lightest Kaluza- Klein particle (LKP) in theories with Unified Extra- Dimensions [3, 4, 5]. These particles will be actively searched for in upcoming high energy physics experi- ments such as the Large Hadron Collider (LHC, see e.g. Refs.[5, 6, 7, 8] for recent discussions in the context of DM searches), while hints on the nature of DM may already come from direct detection experiments aiming at detect- ing the nuclear recoils due to DM scattering off nuclei in large detectors (see e.g. [9] for a recent update on the sta- tus of direct searches). Alternatively, one could search for DM indirectly, i.e. through the detection of its annihila- tion products such as photons, neutrinos, positrons and antiprotons. The annihilation rate being proportional to the square of the DM density, ideal targets of indirect searches include all those regions where the DM density is strongly enhanced, due to gravitational clustering, as in the case of the Galactic center [10, 11, 12, 13, 14, 15, 16] and halo substructures [17, 18, 19, 20, 21, 22, 23, 24, 25], or because of energy losses capture in large celestial bod- ies, as in the case of the Sun and the Earth (see Ref. [2] and references therein). Large DM overdensities can also form as a consequence of astrophysical processes, such as the adiabatic growth of Supermassive [26, 27, 28] or Intermediate Mass Black Holes [29, 30]. In fact, DM halos inevitably react to the growth of black holes, leading, in the case of adiabatic growth, to the formation of large DM overdensities called spikes [26]. A DM cusp with a power-law density profile ρ ∝ r−γ, gets redistributed after the BH growth into a steeper profile ρsp ∝ r−γsp , with γsp = (9 − 2γ)/(4 − γ), within the radius of gravitational influence of the Black Hole (BH) (see below for further details). BHs can thus be thought as annihilation boosters, because the annihila- tion rate after their growth is boosted by several orders of magnitude, making these objects ideal targets for indirect DM searches. Even in absence of mergers [31], and ignor- ing a possible off-center formation [32], a spike around the Supermassive BH at the Galactic center would inevitably be destroyed by the combined effect of gravitational scat- tering off the observed stellar cusp at the GC, and DM annihilations [14]. The very same gravitational processes can still lead to the formation of moderate enhancements called crests (Collisionally REgenerated STructures), but these structures do not lead to significant enhancements of the annihilation signal [33]. Mini-spikes around In- termediate Mass Black Holes (IMBHs) are more promis- ing targets of indirect detection, since they would not be affected by these dynamical processes, and they should appear as bright point-like sources, which could be easily detected by large field of view gamma-ray experiments as GLAST [34] and further studied with ground-based Air Cherenkov telescopes (ACTs) [30] such as CANGAROO [35], HESS [36], MAGIC [37] and VERITAS [38]. Here, we further explore the mini-spikes scenario, and focus on the population of IMBHs in the Andromeda Galaxy (also known as M31), a spiral galaxy very sim- ilar to the Milky Way (MW), whose center is located 784 kpc away from us. We compute gamma-ray fluxes from DM annihilations around IMBHs in M31, and show that the prospects for detection with GLAST are very promising: in an optimistic case (a neutralino with a mass mχ = 150 GeV and annihilation cross section σv = 3 × 10−26 cm3s−1), GLAST may detect up to 20 point- ∗Electronic address: [email protected],[email protected],[email protected] like sources (at 5σ and with a 2 months exposure), within 3◦ from the center of Andromeda. The proposed observa- tional strategy appears particularly suited for ACTs like MAGIC and VERITAS, (M31 is in a region of the sky not accessible to HESS), since the main difficulty in the search for Galactic mini-spikes is that they cannot per- form deep full-sky searches, due to their limited field of view. In the case of mini-spikes in M31, ACTs can search for them by scanning a small region of ≈ 3◦ around the center of M31, and an effective exposure of ≈ 100 hours in this region would be sufficient to probe the proposed scenario, at least for DM mass in the TeV range. The next-generation Cherenkov Telescopes Array (CTA)[39], is expected to significantly improve the sensitivity, in- crease the field of view and decrease the energy threshold with respect to existing ACTs, thus representing an ideal experiment for the proposed scenario. The paper is organized as follows: next section (Sec- tion II) is devoted to describe the formation scenario of IMBHs. We then (Section III) present the IMBHs cata- logue and how it is adapted to the Andromeda Galaxy. We compute the gamma-ray flux emitted by each point- like spike around an IMBH, considering a particular en- ergy annihilation spectrum for a DM particle. In section IV A we estimate the prospects for detection for a generic ACT. Then in section IV B we turn to GLAST. Finally our results are discussed in Section V. II. INTERMEDIATE MASS BLACK HOLES A. IMBHs formation scenario IMBHs are compact objects with mass larger than ≈ 20M⊙, the heaviest remnant of a stellar collapse [40], and smaller than ≈ 106M⊙, the lower end of the mass range of SuperMassive Black Holes (SMBH) [41]. The theoretical and observational motivations for IMBHs were recently reviewed in Ref. [42]. For instance, Ultra-Luminous X- ray point sources (ULXs) could be interpreted as accret- ing IMBHs, since alternative explanations in terms of AGNs, neutron stars or SMBHs appear to be problem- atic or even ruled out [42, 43]. From a theoretical point of view, a population of mas- sive seed black holes could help to explain the origin of SMBHs. In fact, observations of quasars at redshift z ≈ 6 in the Sloan Digital survey [44, 45, 46] suggest that SMBHs were already in place when the Universe was only ∼ 1 Gyr old, a circumstance that can be un- derstood in terms of rapid growth starting from massive seeds (see e.g. Ref. [47]). In fact, a generic prediction of scenarios that seek to ex- plain the properties of the observed SuperMassive Black Hole population, is that a large number of "wandering" IMBHs should exist in DM halos [48, 49, 50]. Despite their theoretical interest, it is difficult to obtain con- clusive evidence for the existence of IMBHs. A viable detection strategy could be the search for gravitational 2 waves produced in the mergers of the IMBH popula- tion [51, 52, 53, 54, 55, 56], with space-based interfer- ometers such as LISA [57]. In Ref. [30], two scenarios for IMBHs formation have been considered. The first posits IMBHs as remnants of the collapse of Population III stars. Zero-metallicity Population III stars are more massive than more recent metal-enriched stars and, if heavier than 260M⊙, they would collapse directly into black holes ([42] and Refs. therein). Here, we will focus only on the second scenario, based on Ref. [49], where IMBHs form at high redshift from gas collapsing in mini-halos. If the latter are massive enough, proto-galactic disks form at the center of each halo. Gravitational instabilities introduce an effective viscosity that causes an inward mass and an outward angular momentum flow. The process goes on till it is interrupted by feedback from star formation (1-10 Myrs) that heats the disk. Then the so-formed object undergoes gravitational collapse into a black hole. A characteristic mass scale of 107M⊙ is imprinted to the mini-halo by the requirements that it is heavy enough to form a gravita- tional unstable disc and that the black hole formation timescale is shorter than the typical major mergers one. The resulting black holes have a mass log-normally scat- tered, with a σ• = 0.9, around the mean value of [49]: M• = 3.8 × 104M⊙(cid:16) κ 18 (cid:19)3/2(cid:18) ×(cid:18) 1 + z 0.5(cid:17)(cid:18) f 10 Myr(cid:19) , 0.03(cid:19)3/2(cid:18) Mvir 107M⊙(cid:19) (1) t where κ is that fraction of the baryonic mass which loses its angular moment that remains in the remnant black hole. f is the fraction of the total baryonic mass in the halo that has fallen into the disc, Mvir is the halo virial mass, z is the redshift of formation and t the timescale for the evolution of the first generation of stars. In Ref. [30] Bertone et al. have studied the population of IMBHs in the MW, following the evolution of mini- halos hosting IMBHs at high redshift (populated with the prescriptions of Ref. [49]), down to redshift z = 0 (see Ref. [30] for further details). As a result, they obtained 200 realizations of the IMBHs population in the Galaxy, that were used to produce 200 mock catalogs of DM mini- spikes, and to study the prospects for detection of these objects in the Galaxy. The average number of unmerged IMBHs was found to be N = 101± 22, and each of these objects is characterized by its mass, distance from the center of the galaxy, and surrounding DM distribution. B. DM distribution around IMBHs Following earlier work on the dynamics of stars and DM around compact objects (see Ref. [58] and references therein), Gondolo and Silk have shown that the adiabatic growth of a massive black hole in the center of a dark halo modifies the distribution of the surrounding DM, induc- ing an enhancement of the density called "spike" [26]. They focused their attention on the SMBH at the center of our Galaxy, but the same formalism can be applied also to IMBHs. The initial DM distribution in all mini- halos can be adequately parametrized with a Navarro, Frenk and White (NFW) profile [59]: ρ(r) = ρ0(cid:16) rs r (cid:17)(cid:18)1 + r rs(cid:19)−2 , (2) where rs, called the scale radius, sets the radius at which the profile slope changes. The new profile after the adi- abatic growth, will be [26]: gdN dx 310 210 10 1 -110 -210 3 FPS Eq.10 BBEG ρsp(r) = ρ(rsp)(cid:18) r rsp(cid:19)−7/3 , (3) -310 -210 -110 x 1 where ρ is the density function of the initial NFW profile. rsp gives the upper limit inside which Eq. 3 is considered valid and is related to the radius of gravitational influ- ence of the black hole rh: rsp ≈ 0.2rh [60], where rh is implicitly defined as: M (r < rh) ≡ Z rh 0 ρ(r)r2dr = 2M• with M• is the mass of the black hole. The spike profile diverges at low radii but annihila- tions set an upper limit to the physical density. Solving the evolution of the DM particles number density, one finds that the upper limit depends on the microphysical properties of the DM particles (mass and annihilation cross section) and on the evolution timescale of the black hole. We denote the distance where the density equals this upper limit rlim, and following Ref. [30] we define a cut-radius for our density profiles: rcut = Max[4RSchw, rlim], (4) where RSchw = 2.95 km M/M⊙ is the BH Schwarzschild radius. The density between RSchw and rcut is assumed to be constant to ρsp(rcut). Figure 1: Differential photon spectrum per annihilation. Dif- ferent parametrizations and annihilation channels are shown. Solid line (FPS) is an analytic fit relative to the b¯b channel, as obtained in Eq. 7. Dashed line (Eq. 10) is relative to the same annihilation channel b¯b, but with a different parametrization of the FFs (see Eqs. 9 and 10). Dotted line (BBEG) is rel- ative to B 1 annihilations and includes final state radiation from annihilation to charged leptons [65] (see text for more details) between the host halo masses, since the number of un- merged IMBHs scales linearly with the host halo mass, and the galactocentric distance by the ratio of virial radii. We obtain for M31 an average number of IMBHs per re- alization NM31 = 65.2 ± 14.5. The mass spectrum re- mains unchanged, with an average mass around 105 M⊙, while the average distance from the center of the galaxy is 32.31 kpc. We have verified that our rescaling proce- dure reproduces in a satisfactory way the properties of the IMBHs population in Andromeda, by comparing our results with a limited number of mock catalogs obtained as an exploratory study in Ref. [30]. III. GAMMA-RAYS FROM IMBHS IN M31 A. IMBHs in M31 Although similar, the Milky Way and Andromeda do not have exactly the same properties. The mock catalogs of IMBHs built for our Galaxy, thus have to be modified to account for the different average number and different spatial distribution in the host halo. A comparison be- tween the properties of Andromeda and of the Galaxy is shown in Table I. We start from the mock catalogs obtained in Ref. [30] and we rescale the total number of objects by the ratio Distance to the center [kpc] Virial Radius [kpc] Virial Mass [M⊙] rs [kpc] ρ0 [ M⊙ kpc3 ] Milky Way Andromeda 8.5 205 784.0 180 1.0 × 1012 6.8 × 1011 21.75 8.18 5.376 × 106 3.780 × 107 Table I: Distance from the Sun (in kpc), virial radius (defined as the radius within which the density reaches 200 times the critical density, in kpc), virial mass (in solar masses) and the two NFW density profile parameters (in kpc and M⊙kpc−3 respectively), both for the MW and the Andromeda Galaxy [61, 62]. ](cid:176) [ [cm ]-1s-2 -12 10 1.5 1 0.5 0 -0.5 -1 4 300 GeV 500 GeV 1000 GeV 1400 1200 1000 800 600 400 200 0 -15 -14.5 -14 -13.5 -13 -12.5 -12 -11.5 log -11 F( 10 ) 0.5 0.4 0.3 0.2 0.1 0 1.5 ](cid:176) [ -1.5 -1.5 -1 -0.5 0 0.5 1 Figure 2: Left: Map of gamma-rays emission (above 100 GeV, in cm−2s−1) by DM annihilation (mχ = 1 TeV) around IMBHs in Andromeda, averaged over all realizations. Bins are 0.1◦ wide, to match the angular resolution of ACTs and GLAST. The circle shows for comparison the M31 scale radius rs. Right: Luminosity function of IMBHs (fluxes are in cm−2s−1), for mχ = 0.3, 0.5 and 1 TeV. The vertical line shows the contribution of the smooth component of the M31 halo, assuming a NFW profile and mχ = 1 TeV. B. Gamma-rays flux from IMBHs in M31 Once the mock catalogs of IMBHs in M31 have been obtained, it is possible to calculate the gamma-ray flux from each IMBH in every realization. The calculation follows Eq. 14 in Ref. [30]: Φ(E) = 1 d2 σv 2m2 χ dNγ(E) = Φ0 rcut σv dNγ(E) ρ2(r)r2dr dE Z rsp 10−26cm3/s(cid:19)(cid:16) mχ 100 GeV/cm3(cid:19)2 10−3 pc(cid:19)−5/3 dE (cid:18) 780 kpc(cid:19)−2(cid:18) 5 pc(cid:19)14/3(cid:18) rcut ρ(rsp) d , ×(cid:18) ×(cid:18) rsp (5) 1 TeV(cid:17)−2 where Φ0 = 2.7×10−14 cm−2s−1, d is the IMBH distance to the observer, σv is the DM annihilation cross section times relative velocity and mχ is the DM particle mass (the letter χ, usually adopted for neutralino, is used here to denote a generic WIMP candidate). rcut and rsp, rep- resent the inner and outer size of the spike, as discussed in the previous section. dNγ(E)/dE is the differential photon yield per annihi- lation, that can be expressed as: dNγ(E) dE = Xa Ba dN a γ (E) dE , (6) where Ba is the branching ratio BR(χχ → a¯a) and dN a γ /dE the secondary photon spectrum relative to the annihilation channel a¯a. The latter term is thus a purely Standard Model calculation, while branching ratios have to be derived in the framework of new theories beyond the Standard Model, such as SUSY or UED. We review here different parametrizations of the pho- ton yield that have been recently proposed in literature. The first parametrization we focus on, has been obtained in Ref. [61], and it is relative to annihilations into b¯b. The authors have parametrized the results obtained with the event generator PYTHIA [63] as follows dN b γ(x) dx = xaeb+cx+dx2+ex3 , (7) where the parameters depend on the neutralino mass, and for the specific case mχ = 1 TeV, (a, b, c, d, e) = (−1.5, 0.37, −16.05, 18.01, −19.50). While for annihila- tion to τ s dN τ γ (x) dx = xa(bx + cx2 + dx3)eex, (8) = = for mχ (a, b, c, d, e) and 1 TeV, (−1.31, 6.94, −4.93, −0.51, −4.53). Alternatively, one can start from the most recent Frag- mentations Functions (FFs) (e.g. [64]), describing the hadronization of partons into the particles of interest. The FF of b quarks hadronizing in neutral pions has been fitted with a simple analytic form that captures in a sat- isfactory way the behavior of the FF at large x finding · j q F the following analytic fit f (x) = 7.53 x0.87e14.62x . (9) Convolving the spectrum pions with their decay spec- trum into photons one finally obtains the differential pho- ton yield dNγ(x) dx = Z 1 x f (x′) 2 x′ dx′. (10) We have also considered an example inspired from the- ories with Unified Extra-Dimensions, where the role of DM is usually played by the first excitation of the hy- percharge gauge boson, and referred to as B(1). Since the B(1) annihilation into fermions does not suffer from chirality suppression, as in MSSM, we also include the contribution from annihilation to l¯lγ, as calculated in Ref. [65], as well as the contribution from τ fragmenta- tion, and usual from annihilations to b¯b, with the appro- priate branching ratio. The final state radiation arising from annihilation to charged leptons has a characteristic, very hard, spectral shape [65, 66] dN l γ(x) dx = Xl=e,µ α π x2 − 2x + 2 x ln(cid:20) m2 B(1) m2 l (1 − x)(cid:21) . (11) The three prescriptions for the annihilation spectrum are plotted in Fig. 1 (for mχ = 1 TeV). As expected, all spectra are very similar up to x ≡ E/mχ ∼ 0.1, but the spectrum relative to B(1) annihilations is harder at large x, and exhibits a distinctive sharp cut-off at x = 1. To show the small effect that the adoption of differ- ent annihilation spectra has on the prospects for indirect detection, we have calculated the DM annihilation flux from the smooth component of the M31 halo, assuming a NFW profile with the parameters described in Tab. I above. The results are displayed in Table II, and as one can see, differences are within a factor of 2. In the re- main of this paper, we will thus work only with the first analytic fit, since the uncertainties associated with other astrophysical and particle physics parameters are signif- icantly larger. By calculating the gamma-ray flux in Eq. 5 for IMBHs in all realizations, we obtain a gamma-ray map, above Ethr = 100 GeV, from mini-spikes in Andromeda. Each M31 realization actually produces a different emission M31 flux [cm−2s−1] FPS [61] Eq. 10 BBEG [65] 1.33 × 10−14 9.79 × 10−15 1.60 × 10−14 Table II: Gamma-ray flux over 100 GeV from Andromeda (in cm−2s−1) for a smooth NFW, and for the different parametrizations discussed in the text. Differences among the predicted fluxes are within a factor of 2. 5 signal Galactic/ExtraGal. back. Hadronic/Electronic back. q(<F ) [cm ]-1s-2 -810 -910 -10 10 -11 10 -12 10 -13 10 10 -14 0 0.5 1 1.5 2 2.5 3 ](cid:176) [ Figure 3: Gamma-ray flux (in cm−2s−1) from DM annihila- tion around IMBHs (solid thick line), integrated over a cone of size θ towards the center of M31, as a function of θ. We show for comparison the hadronic/electron background, assuming ǫh = 10−2 (solid thinner line) and the diffuse extragalactic background (dashed line). map, so in Fig. 2 we show the average of all 200 maps, which clearly exhibits, as expected, a strong enhance- ment of the flux in the innermost regions of the Galaxy. The pixel size matches the angular resolution of ground- based telescopes such as VERITAS and MAGIC, and of GLAST. For the map, a DM particle mass of 1 TeV and an annihilation cross section σv = 3× 10−26cm3s−1 have been adopted. Note that this map alone does not provide any information on the detectability of the fluxes, which will be discussed in detail in the next 2 sections. Note also that the actual distribution of IMBHs, will provide (as we shall see later) a much more 'patchy' emission, far from the average smooth behaviour shown in Fig. 2. We also show in the right panel of Fig. 2 the lumi- nosity function of IMBHs (sum of all realizations), for different values of the DM particle mass. The distribu- tion is approximately gaussian, and the average flux of IMBHs is larger than emission due the smooth compo- nent. The dependence from the mass results in due to a balance between the m−9/7 dependence in Eq. 5, and the mχ dependence of the upper limit in the integral of the energy spectrum. Having set in the figure an energy threshold Ethr = 100 GeV, the luminosity flux towards higher fluxes when the mass increase. We will come back later to this threshold effect, that leads to higher fluxes for higher masses when mχ ∼ Ethr despite the explicit m−9/7 dependence of the annihilation flux. Meanwhile we note that this effect disappears when mχ ≫ Ethr, as can be seen from Table III. χ χ q ](cid:176) [ [cm ]-1s-2 10· -12 ](cid:176) [ 2 1 0 -1 -2 ACT -2 -1 0 1 2 2 1 0 -1 -2 3 2.5 2 1.5 1 0.5 0 ](cid:176) [ 6 [cm ]-1s-2 -910 GLAST 0.35 0.3 0.25 0.2 0.15 0.1 0.05 -2 -1 0 1 2 0 ](cid:176) [ Figure 4: Left (right) panel shows a map of the gamma-ray flux in units of photons cm−2s−1, from DM annihilations around IMBHs in M31, relative to one random realization of IMBHs in M31. The size of the bins is 0.1◦ and the threshold for the left (right) panel is 100 GeV (4 GeV) as appropriate for ACTs (GLAST). The circles highlight IMBHs within the reach of current ACTs for a 5σ detection in 100 hours (within the reach of GLAST for a 5σ detection in 2 months). IV. PROSPECTS FOR DETECTION we compare the number of signal photons, to the fluctu- ations of the background As we shall see the prospects for detection de- pend on the expected or measured experimental per- formances, but also on the atmospheric and astrophys- ical backgrounds. We perform separate analysis for Air Cherenkov Telescopes and the upcoming gamma-ray satellite GLAST. A. Prospects for ACTs The calculations in this section are performed for a generic ACT, but they are particularly relevant for two specific experiments: MAGIC and VERITAS. As for HESS, being located in Namibia, it cannot detect gamma-rays from the direction of Andromeda. To determine the significance of the signal from an in- dividual mini-spike, as calculated in the previous section, Average flux [cm−2s−1] mχ = 50 GeV mχ = 150 GeV mχ = 300 GeV mχ = 500 GeV 5.26 × 10−11 7.65 × 10−11 6.92 × 10−11 5.81 × 10−11 n = nγ √nbk = √T · ∆Ω R Aef f (E, θ) dΦ qR Aef f (E, θ) dΦbk dE dEdθ dE dEdθ , (12) where T is the exposure time, Aef f the effective area, ∆Ω the solid angle, dΦbk/dE is the total background differential flux. For Air Cherenkov Telescopes, the main background is due to hadrons interacting with the atmosphere and producing electromagnetic showers. Following Ref. [67] [24], we consider dΦh dΩdE = 1.5 ×(cid:18) E GeV(cid:19)−2.74 p cm2s GeV sr . (13) The ratio of the number of hadrons misinterpreted as gamma-rays, over the total number of cosmic ray hadrons, ǫh, provides an estimate of the telescope po- tential to discriminate the gamma-ray signal from the hadronic background. We adopt a typical value ǫh = 10−2, following Refs. [24] [68]. The electronic contribu- tion to the background is [24]: dΦe dΩdE = 6.9 × 10−2(cid:18) E GeV(cid:19)−3.3 e cm2s GeV sr (14) Table III: Average flux from IMBHs in all 200 realizations (in cm−2s−1), for different values of DM mass, σv = 3 × 10−26cm3s−1 and Ethr = 4 GeV. and it is typically subdominant at the energies of interest. In Figure 3 we compare the DM annihilation signal with the different sources of background, as a function of j q F · j q F 7 average realization ACTs GLAST GLAST ACTs GLAST ACTs 10 1 -110 )sN(>5 24 22 20 18 16 14 12 10 8 6 4 2 0 0 200 400 600 800 1000 1200 cm 1400 [GeV] 0 1 2 3 4 5 6 7 ](cid:176) [ j Figure 5: Number of detectable mini-spikes in M31 with GLAST (2 months) and ACTs (100 hours) as a function of the DM particle mass (left) and as a function of the angular distance from the center of M31 (right). In the left panel, error bars denote the 1 − σ scatter among different realizations. In the right panel, the total number of objects is shown as an empty histogram, while the vertical lines denote the size of the region that contains 90% of the detectable IMBHs. the field of view. The minimum flux for a 5σ detection with an effective area of Aef f = 3×104 m2 [37] and an ex- posure time of 100 hours, is φmin = 1.6× 10−12cm−2s−1. To produce this estimate we have considered values of effective area and angular resolution similar to MAGIC and the result is consistent with earlier estimates of the MAGIC sensitivity [2]. An actual estimate of the instrument performance suggests that the mini- mum flux can be up to an order of magnitude higher [69]. The fluxes in Fig. 2 are thus found to lie below the min- imum detectable flux so determined, so one may naıvely conclude that there is no hope to detect them with this experimental setting. However, for a more careful assess- ment of the prospects for detection, we need to estimate the detectability in each realization and then average over all the realization, and not viceversa. In fact, as demon- strated in the left panel of Fig. 4, in each random real- ization, the emission is much more patchy, with a large number of high emission peaks, corresponding to indi- vidual mini-spikes, that can thus be resolved with the adopted angular resolution. Black circles highlight the position of objects brighter than the experimental sensi- tivity (indicated in the color scale by the black line). In total, for mχ = 1 TeV, the number of detectable IMBHs is N5σ = 5.2 ± 3.1, where the error is relative to the 1-σ scatter among different realizations. We note that current simulations indicate that the next-generation Cherenkov Telescopes Array (CTA)[39], may significantly improve the sensitivity, down to φmin ≈ 10−13cm−2s−1, thus leading to a substantial improve- ment in the prospects for detection. B. Prospects for GLAST The space satellite GLAST is expected to play a crucial role in indirect DM searches, thanks both to its ability to perform observations at energy scales comparable to the mass of common DM candidates and to its poten- tial of making deep full-sky maps in gamma-rays, thanks to its large (∼ 2.4 sr) field of view [34]. Despite the smaller effective area, it is not affected, being a satellite, by the atmospheric hadronic and electron background. Furthermore, its lower energy threshold (30 MeV) allows to probe lighter DM particles, typically leading to higher fluxes. The angular resolution of GLAST is ≈ 3◦ in the energy range 30 MeV-500 MeV, becomes 0.5◦ from 500 MeV to 4 GeV, and reaches 0.15◦ above 4 GeV [34]. As in the case of ACTs, we compare the expected fluxes with the photon background, which in this case, since GLAST will perform observations above the atmosphere, is mainly due to diffuse gamma-ray emission. The galac- tic and extragalactic background has been measured in [70, 71] by EGRET in the energy range between 30 MeV and 10 GeV and we extrapolate it to higher energies by fitting with a power-law with spectral index of -2.1. The resulting formula is dΦextra/gal dΩdE = 2.3 × 10−6(cid:18) E GeV(cid:19)−2.1 γ cm2s GeV sr . (15) [72, 73, 74, 75], We note here that a large fraction of the observed gamma-ray background might be actually due to DM annihilations in particular if as- trophysical processes can boost the annihilation sig- nal In this case, the smoking-gun for this scenario would be the distinctive shape of the angular power-spectrum of the background, that may allow, al- ready with GLAST, the discrimination against ordinary astrophysical sources [78]. [76, 77]. The sensitivity above 30 MeV, i.e. the minimum de- tectable flux for a 5σ detection with an exposure of 2 months, is found to be φmin = 3.2 × 10−8cm−2s−1. This value is derived from Eq. 12, adopting values of the en- ergy dependent effective area provided by [79], and is consistent with GLAST sensitivity maps obtained in Ref. [80]. The integral flux above threshold from IMBHs, av- eraged among realizations and integrated in a 3◦ cone towards M31, is φ30 = 1.3 × 10−7cm−2s−1. In the right panel of Fig. 4, we show the results of our analysis relative to a random realization, and adopting mχ = 150 GeV, and σv = 3×10−26 cm3 s−1. Mini-spikes appear as high emission peaks, and can be easily resolved by selecting photons above 4 GeV, so that the angular resolution of GLAST approaches 0.1 degrees. Black cir- cles highlight those objects that produce a flux detectable at 5σ with GLAST, with a 2 months exposure. V. DISCUSSION AND CONCLUSIONS Although we have performed the analysis of the prospects for detection with GLAST and ACTs for 2 dif- ferent benchmark scenarios (essentially high DM particle mass for ACTs, low mχ for GLAST), the analysis can be easily extended to any value of the particle physics pa- rameters of the annihilating DM particle. To explore the dependence on mχ, we show in the left panel of Fig. 5 the number of objects that can be detected with the afore- mentioned experiments, as a function of the DM parti- cle mass. Near the experiment threshold, fluxes increase with mass. When mχ ≫ Ethr this threshold effect dis- appears and one recovers the expected behavior (smaller fluxes for higher masses). Similarly, one can plot the number of detectable ob- jects as a function of the angular distance from the cen- 8 ter of M31, to estimate the region where most mini-spikes can be found. This is shown in the right panel of Fig. 5 where the total number of objects is also shown for com- parison. Vertical lines denote the angular size of the re- gion that contains 90% of the detectable IMBHs for the various experiments, which has a characteristic size of θ = 3.3◦. We stress that while in the case of Galactic IMBHs the identification of mini-spikes will require a case-by- case analysis of the spectral properties of unidentified gamma-ray sources, the detection of a cluster of sources around the center of Andromeda would per se provide unmistakable evidence for the proposesd scenario. In conclusion, we have computed gamma-ray fluxes from DM annihilations in mini-spikes around IMBHs in the Andromeda Galaxy. We have studied the prospects for detection with Air Cherenkov telescopes like MAGIC and VERITAS and with the GLAST satellite, and found that a handful of sources might be within the reach of current ACTs, while the prospects for the planned CTA are more encouraging. The obvious advantage of the pro- posed scenario with respect to mini-spikes in the MW, is that they are not randomly distributed over the sky, but they are contained, at 90%, within 3 degrees from the center of Andromeda, and can thus be searched for with ACTs by performing a deep scan of this small region. The prospects for GLAST appear more promising, since an exposure time of 2 months allows the detection of up to of ≈ 20 mini-spikes, that would be resolved as a cluster of point-sources with identical spectra, within a ∼ 3◦ region around the center of Andromeda. Such a distinctive prediction cannot be mimicked by ordinary astrophysical sources. As in the case of IMBHs in the MW, null searches would place very strong constraints on the proposed scenario in a wide portion of the DM parameter space. Acknowledgements We thank Riccardo Rando, of the GLAST collabora- tion, as well as Mos`e Mariotti and Michele Doro of the MAGIC collaboration, for useful information and discus- sions on experimental strategies and sensitivities. We also thank Andrew Zentner for earlier collaboration and for providing the mock catalogs of IMBHs in the MW and Lidia Pieri for comments. GB is supported by the Helmholtz Association of National Research Centres. [1] L. Bergstrom, Rept. Prog. Phys. 63 (2000) 793 [arXiv:hep-ph/0002126]. [2] G. Bertone, D. Hooper and J. Silk, Phys. Rept. 405 (2005) 279 [arXiv:hep-ph/0404175]. [6] http://lhc.web.cern.ch/lhc/. [7] E. A. Baltz, M. Battaglia, M. E. Peskin and T. Wizanski, Phys. Rev. D 74 (2006) 103521 [arXiv:hep-ph/0602187]. [8] A. K. Datta, K. Kong and K. T. Matchev, Phys. Rev. D [3] T. Appelquist, H.-C. Cheng and B. A. Dobrescu, Phys. 72 (2005) 096006 [arXiv:hep-ph/0509246]. Rev. D 64 (2001) 035002 [arXiv:hep-ph/0012100]. [9] C. Munoz, Int. J. Mod. Phys. A 19 (2004) 3093 [4] G. Servant and T. M. P. Tait, Nucl, Phys. B 650 (2003) [arXiv:hep-ph/0309346]. 391 [arXiv:hep-ph/0206071]. [10] F. Aharonian et al. [H.E.S.S. Collaboration], [5] H.-C. Cheng, K. T. Matchev and M. Schmaltz, Phys. Rev. D 66 (2002) 056006 [arXiv:hep-ph/0205314]. Phys. Rev. Lett. 97 (2006) 221102 [Erratum-ibid. 97 (2006) 249901] [arXiv:astro-ph/0610509]. 9 [11] G. Zaharijas and D. Hooper, Phys. Rev. D 73 (2006) [46] C. J. Willott, R. J. McLure and M. J. Jarvis, Astrophys. 103501 [arXiv:astro-ph/0603540]. [12] S. Profumo, Phys. Rev. D 72 (2005) 103521 [arXiv:astro-ph/0508628]. J. 587 (2003) L15 [arXiv:astro-ph/0303062]. [47] Z. Haiman, and A. Loeb, Astrophys. J. 552 (2001) 459. [48] R. Islam, J. Taylor and J. Silk, Mon. Not. Roy. Astron. [13] Y. Mambrini, C. Munoz, E. Nezri and F. Prada, JCAP Soc. 340 (2003) 6471 [arXiv:astro-ph/0208189]. 0601 (2006) 010 [arXiv:hep-ph/0506204]. [14] G. Bertone and D. Merritt, Phys. Rev. D 72 (2005) 103502 [arXiv:astro-ph/0501555]. [15] A. Cesarini, F. Fucito, A. Lionetto, A. Morselli 267 and P. Ullio, Astropart. Phys. 21 (2004) [arXiv:astro-ph/0305075]. [49] S. M. Koushiappas, J. S. Bullock and A. Dekel, 292 Mon. Not. Roy. Astron. Soc. 354 (2004) [arXiv:astro-ph/0311487]. [50] M. Volonteri, F. Haardt, and P. Madau, Astrophys. J. 582 (2003) 559 [arXiv:astro-ph/0207276]. [51] K. S. Thorne and V. B. Braginsky, Astrophys. J. Lett. [16] A. Bouquet, P. Salati and J. Silk, Phys. Rev. D 40 (1989) 204 (1976) L1. 3168. [17] J. Silk and A. Stebbins, Astrophys. J. 411 (1993) 439. [18] L. Bergstrom, J. Edsjo and P. Ullio, Phys. Rev. D 58, 083507 (1998) [arXiv:astro-ph/9804050]. [52] ´E ´E Flanagan and S. A. Hughes, Phys. Rev. D 57 (1998) 4566. [53] ´E ´E Flanagan and S. A. Hughes, Phys. Rev. D 57 (1998) 4535. [19] C. Calcaneo-Roldan and B. Moore, Phys. Rev. D 62, [54] R. Islam, J. Taylor and J. Silk, Mon. Not. Roy. Astron. 123005 (2000) [arXiv:astro-ph/0010056]. Soc. 354 (2004) 629 [arXiv:astro-ph/0309559]. [20] R. Aloisio, P. Blasi and A. V. Olinto, Astrophys. J. 601 [55] T. Matsubayashi, H. Shinkai and T. Ebisuzaki, Astro- (2004) 47 [arXiv:astro-ph/0206036]. [21] S. M. Koushiappas, A. R. Zentner and T. P. Walker, Phys. Rev. D 69 (2004) 043501 [arXiv:astro-ph/0309464]. [22] J. Diemand, B. Moore and J. Stadel, Nature 433 (2005) 389 [arXiv:astro-ph/0501589]. [56] S. M. phys. J.614 (2004), 864. Koushiappas arXiv:astro-ph/0503511. and A. R. Zentner, [57] http://lisa.nasa.gov/. [58] G. D. Quinlan, L. Hernquist, S. Sigurdsson, Astrophys. [23] L. Pieri, E. Branchini and S. Hofmann, Phys. Rev. Lett. J. 440 (1995) 554. 95, 211301 (2005) [arXiv:astro-ph/0505356]. [59] J. F. Navarro, C. S. Frenk, S. D. M. White, Astrophys. [24] L. Pieri and E. Branchini, Phys. Rev. D 69, 043512 J. 490 (1997) 493. (2004) [arXiv:astro-ph/0307209]. [25] E. A. Baltz, J. E. Taylor and L. L. Wai, arXiv:astro-ph/0610731. [26] P. Gondolo, J. Silk, Phys. Rev. Lett. 83 (1999) 1719 [arXiv:astro-ph/9906391]. [60] D. Merritt, Proceedings of Carnegie Observatories Cen- tennial Symposium Coevolution of Black Holes and Galaxies [arXiv:astro-ph/0301257]. [61] N. Fornengo, L. Pieri, S. Scopel, Phys. Rev. D 70 (2004) 103529 [arXiv:hep-ph/0407342]. [27] G. Bertone, G. Sigl and J. Silk, Mon. Not. Roy. Astron. Soc. 337 (2002) 98 [arXiv:astro-ph/0203488]. [28] P. Gondolo, Phys. Lett. B 494, 181 (2000) [62] J. J. Geehan, M. A. Fardal, A. Babul, P. Guhathakurta, 996 Mon. Not. Roy. Astron. Soc. 366 (2006) [arXiv:astro-ph/0501240]. [arXiv:hep-ph/0002226]. [29] H. S. Zhao and J. Silk, Phys. Rev. Lett. 95, 011301 (2005) [arXiv:astro-ph/0501625]. [30] G. Bertone, A. R. Zentner, J. Silk, Phys. Rev. D 72 (2005) 103517 [arXiv:astro-ph/0509565]. [31] D. Merrit, M. Milosavljevic, L. Verde, R. Jimenez, Phys. Rev. Lett. 88 (2002) 191301 [arXiv:astro-ph/0201376]. [32] P. Ullio, H. Zhao, M. Kamionkowski, Phys. Rev. D 64 (2001) 043504 [arXiv:astro-ph/0101481]. [63] http://projects.hepforge.org/pythia6/. D [64] S. Kretzer, 054001 www.pv.infn.it/%7Eradici/FFdatabase. Phys. [arXiv:hep-ph/0003177] Rev. 62 (2000) and [65] L. Bergstrom, T. Bringmann, M. Eriksson, M. 131301 Gustafsson, Phys. Rev. Lett. 94 (2005) [arXiv:astro-ph/0410359]. [66] J. F. Beacom, N. F. Bell and G. Bertone, Phys. Rev. Lett. 94 (2005) 171301 [arXiv:astro-ph/0409403]. [33] D. Merritt, S. Harfst and G. Bertone, Phys. Rev. D 75 [67] T. Gaisser, M. Honda, P. Lipari and T. Stanev, in Proc. (2007) 043517 [arXiv:astro-ph/0610425]. [34] http://www-glast.stanford.edu. [35] http://icrhp9.icrr.u-tokyo.ac.jp/. [36] http://mpi-hd.mpg.de/hfm/HESS/HESS.html. [37] http://hegra1.mppmu.mpg.de/MAGICweb/. [38] http://veritas.sao.arizona.edu/index.html. [39] http://www.mpi-hd.mpg.de/hfm/CTA/CTA home.html. [40] C. L. Fryer and V. Kalogera, Astrophyis. J. (2001), 554 [arXiv:astro-ph/9911312]. of the 27th ICRC, (2001). [68] L. Bergstrom, D. Hooper, Phys. Rev D 73 (2006) 063510 [arXiv:hep-ph/0512317]. [69] http://wwwmagic.mppmu.mpg.de/physics/results/index.html. [70] A. N. Cillis, R. C. Hartman, Astrophys. J. 621 (2005) 291. [71] http://heasarc.gsfc.nasa.gov/FTP/compton/data/egret/diffuse maps. [72] L. Bergstrom, J. Edsjo and P. Ullio, Phys. Rev. Lett. 87 (2001) 251301 [arXiv:astro-ph/0105048]. [41] L. Ferrarese and H. Ford, Space Science Reviews (2005), [73] J. E. Taylor and J. Silk, Mon. Not. Roy. Astron. Soc. 339 116. [42] M. C. Miller, E. J. M. Colbert, Int. J. Mod. Phys. D 13 (2004) 1 [arXiv:astro-ph/0308402]. [43] D. A. Swartz, K. K. Ghosh, A. F. Tennant and K.-W. Wu , arXiv:astro-ph/0405498. (2003) 505 [arXiv:astro-ph/0207299]. [74] P. Ullio, L. Bergstrom, J. Edsjo and C. Lacey, Phys. Rev. D 66 (2002) 123502 [arXiv:astro-ph/0207125]. [75] S. Ando, Phys. Rev. Lett. 94 (2005) 171303 [arXiv:astro-ph/0503006]. [44] X. Fan et al. [SDSS Collaboration], Astron. J. 122 (2001) [76] S. Horiuchi and S. Ando, Phys. Rev. D 74, 103504 (2006) 2833 [arXiv:astro-ph/0108063]. [arXiv:astro-ph/0607042]. [45] A. J. Barth, P. Martini, c. H. Nelson and L.C. Ho, As- [77] E. J. Ahn, G. Bertone and D. Merritt, trophys. Lett. 594 (2003) L95. arXiv:astro-ph/0703236. [80] G. Bertone, T. Bringmann, R. Rando, G. Busetto and [78] S. Ando, E. Komatsu, T. Narumoto and T. Totani, A. Morselli, arXiv:astro-ph/0612387. arXiv:astro-ph/0612467. [79] http://www-glast.slac.stanford.edu/software/IS/glast lat performance.htm 10
astro-ph/0207136
2
0207
2002-12-05T12:56:39
Statistical characteristics of observed Ly-$\alpha$ forest and the shape of initial power spectrum
[ "astro-ph" ]
Properties of $\sim$ 5000 observed Ly-$\alpha$ absorbers are investigated using the model of formation and evolution of DM structure elements based on the Zel'dovich theory. This model is generally consistent with simulations of absorbers formation, accurately describes the Large Scale Structure observed in the galaxy distribution at small redshifts and emphasizes the generic similarity of the LSS and absorbers. The simple physical model of absorbers asserts that they are composed of DM and gaseous matter and it allows us to estimate the column density and overdensity of DM and gaseous components and the entropy of the gas trapped within the DM potential wells. The parameters of DM component are found to be consistent with theoretical expectations for the Gaussian initial perturbations with the WDM--like power spectrum. We demonstrate the influence of the main physical factors responsible for the absorbers evolution. The analysis of redshift distribution of absorbers confirms the self consistence of the assumed physical model and allows to estimate the shape of the initial power spectrum at small scales what in turn restricts the mass of dominant fraction of DM particles to $M_{DM}\approx 0.6 - 2$ keV. Our results verify the redshift variations of intensity of the UV background by about $4 - 6$ times and indicate that, perhaps, the available observations underestimate the intensity of this background.
astro-ph
astro-ph
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000) Printed 30 October 2018 (MN LATEX style file v1.4) Statistical characteristics of observed Ly-α forest and the shape of initial power spectrum M. Demia´nski1,2, A.G. Doroshkevich3,4 & V.Turchaninov4 1Institute of Theoretical Physics, University of Warsaw, 00-681 Warsaw, Poland 2Department of Astronomy, Williams College, Williamstown, MA 01267, USA 3Theoretical Astrophysics Center, Juliane Maries Vej 30, DK-2100 Copenhagen Ø, Denmark 4Keldysh Institute of Applied Math Russian Academy of Sciences, 125047 Moscow, Russia Accepted ..., Received ..., in original form ... . ABSTRACT Properties of about 4 500 observed Ly-α absorbers are investigated using the model of formation and evolution of DM structure elements based on the modified Zel'dovich theory. This model is generally consistent with simulations of absorbers formation, describes reasonably well the Large Scale Structure observed in the galaxy distribution at small redshifts and emphasizes the generic similarity of the LSS and absorbers. The simple physical model of absorbers asserts that they are composed of DM and gaseous matter. It allows us to estimate the column density and overdensity of DM and gaseous components and the entropy of the gas trapped within the DM potential wells. The parameters of DM component are found to be consistent with theoretical expectations for the Gaussian initial perturbations with the WDM -- like power spectrum. The basic physical factors responsible for the absorbers evolution are discussed. The analysis of redshift distribution of absorbers confirms the self consistency of the adopted physical model, Gaussianity of the initial perturbations and allows to estimate the shape of the initial power spectrum at small scales what in turn restricts the mass of the dominant fraction of DM particles to MDM ≥ 1.5 − 5 keV. Our results indicate a possible redshift variations of intensity of the UV background by about a factor of 2 -- 3 at redshifts z ∼ 2 -- 3. Key words: cosmology: large-scale structure of the Universe -- quasars: absorption: general -- surveys. 1 INTRODUCTION One of the most perspective methods to study the processes responsible for the formation and evolution of the structure of the Universe is the analysis of properties of absorbers observed in spectra of the farthest quasars. The great po- tential of such investigations was discussed already in Oort (1981, 1984) just after Sargent et al. (1980) established the intergalactic nature of the Ly-α forest. The available Keck and VLT high resolution observations of the forest provide a reasonable database and allow to apply statistical methods for such investigations. The essential progress achieved recently through nu- merous high resolution simulations of absorbers formation and evolution confirms that this process is closely connected with the initial power spectrum of perturbations. These re- sults allow us to consider the properties of absorbers in the context of nonlinear theory of gravitational instability (Zel'dovich 1970; Shandarin & Zel'dovich 1988) and to apply c(cid:13) 0000 RAS the statistical description of structure formation and evolu- tion (Demia´nski & Doroshkevich 1999, 2002; hereafter DD99 & DD02) to the Ly-α forest. This approach is based on the modified Zel'dovich ap- proximation and describes the formation and evolution of DM structure elements for the CDM and WDM initial power spectra without any smoothing or filtering procedures. It outlines the structure evolution as a random formation and merging of Zel'dovich pancakes, their transverse expansion and/or compression and transformation into high density clouds and filaments. Later on, the hierarchical merging of pancakes, filaments and clouds forms rich galaxy walls ob- served at small redshifts. Main stages of this evolution are driven by the initial power spectrum. At small redshifts the main results of the statistical ap- proach are found to be consistent with characteristics of the observed and simulated Large Scale Structure (Demia´nski et al. 2000; Doroshkevich, Tucker & Allam 2002). Here we use 2 Demia´nski, Doroshkevich & Turchaninov this approach for the analysis and interpretation of the ab- sorbers observed at large redshifts. We show that the basic observed characteristics of Ly-α forest are also successfully described in the framework of this theoretical model. The application of this approach to Ly-α forest was al- ready discussed in Demia´nski, Doroshkevich & Turchaninov (2001 a,b, hereafter Paper I & Paper II). Here we improve this analysis by using a richer and more refined sample of ∼ 4500 observed absorbers and a more refined model of ab- sorbers. Our approach uses more traditional methods of in- vestigation of properties of discrete absorbers rather then associating them with a continuous non-linear line-of-sight density field (see, e.g., Weinberg et al. 1998; McDonald et al. 2000; Croft et al. 2001; Croft et al. 2002). It allows to reach more clarity in the description of formation and evolution of absorbers and to reveal its strong link with the initial power spectrum. It allows also to separate several subpopulations of absorbers and to discuss their evolutionary history. The composition and spatial distribution of observed absorbers is complicated and at low redshifts a significant number of stronger Ly-α lines and metal systems is asso- ciated with galaxies (Bergeron et al. 1992; Lanzetta et al. 1995; Tytler 1995; Le Brune et al. 1996). However as was recently shown by Penton, Stock and Shull (2002), even at small redshifts some absorbers are associated with galaxy filaments while others are found within galaxy voids. These results suggest that the population of weaker absorbers dom- inating at higher redshifts can be associated with weaker structure elements formed by the non luminous baryonic and DM components in extended low density regions. In turn, the relatively homogeneous spatial distribution of ab- sorbers implies a more homogeneous spatial distribution of both DM and baryonic components as compared with the observed distribution of the luminous matter. For the truncated Zel'dovich theory (Coles, Melot & Shandarin 1993; Bond & Wadsley 1997), some statistical characteristics of absorbers were already discussed in Gnedin & Hui (1996); Hui, Gnedin & Zhang (1997); and Hui & Rut- ledge (1999). In particular, in these papers possible fits for the number density of absorbers with a threshold hydrogen column density, NHI , or with a threshold Doppler parame- ter, b, were proposed. However, because of the complex evo- lution of absorbers, these functions depend upon both the initial power spectrum and several random factors discussed below (Secs. 2.3, 4.3). So, the application and interpretation of these fits are limited. Here we use a statistical approach (DD02) based on the modified Zel'dovich theory which describes both linear and nonlinear evolution of the DM structure. For the CDM -- like power spectrum and Gaussian initial perturbations it gives the expected mean 1D number density and the probabil- ity distribution function of DM pancakes. Both functions depend only on the cosmological model and moments of the initial power spectrum. Comparison of these functions with absorbers distributions confirms the self consistency of adopted physical model and Gaussianity of the initial den- sity perturbations, allows to estimate the spectral moments and to restrict the mass of dominant DM particles. These re- sults complement the investigations of the linear power spec- trum (see, e.g., Croft, Weinberg, Katz & Hernquist 1998; Gnedin & Hamilton 2002; Zaldarriaga, Scoccimorro & Hui 2002; Croft et al. 2002). However, this theory does not describe the process of relaxation of compressed matter, the disruption of structure elements due to the gravitational instability of DM pan- cakes and the distribution of neutral hydrogen across DM pancakes. Now we have only limited information about prop- erties of the background gas and UV radiation (Scott et al. 2000; Schaye et al. 2000; McDonald & Miralda -- Escude 2001; McDonald et al. 2000, 2001; Theuns et al. 2002a, b). There- fore, in this paper some numerical factors remain undeter- mined. They could be more precisely determined by special investigations of simulated and observed absorbers. This analysis should be supplemented by application of the discussed here approach to simulations which take simul- taneously into account the impact of many important factors and provide unified picture of the process of absorbers for- mation and evolution (see, e.g., Weinberg et al. 1998; Zhang et al. 1998; Dav´e et al. 1999; Theuns et al. 1999). But so far such simulations can be performed only in small boxes what restricts their representativity, introduces artificial cutoffs in the power spectrum and complicates the quantitative de- scription of structure evolution (see more detailed discussion in Hui & Gnedin 1997 and Paper II). The semianalitic mod- els (see, e.g., Bi & Davidsen 1997; Choudhury, Padmanab- han & Srianand 2001; Choudhury, Srianand, & Padmanab- han 2001; Viel et al. 2002) show other perspective approach to study the structure evolution at high redshifts. Perhaps, some results obtained below can be also useful for this pur- pose. Comparison of results obtained in Paper I and Paper II and in this paper demonstrates that the quality and rep- resentativity of the sample of observed absorbers are very important for the reconstruction of processes of absorbers formation and evolution. In this paper we use the sample of ∼ 4500 absorbers at 1.7 ≤ z ≤ 4 compiled from 18 high resolution spectra. However, the redshift distribution of ab- sorbers is inhomogeneous what increases errors and gener- ates some uncertainties. In particular, the statistics of ab- sorbers at z ≥ 3.5 is very poor. Because of the preliminary character of our investigation we estimate the precision of only the most important parameters. Further progress can be achieved with richer samples covering the range of redshifts at least up to z ∼ 4.5 -- 5. The observational evidence of the reheating of the Universe at z ≈ 6 (Djorgovski et al 2001; Fan et al. 2001) makes it important to perform a complex investigation of the early period of structure evolution at redshifts z ∼ 4 -- 6. This paper is organized as follows. The theoretical model of the structure evolution is discussed in Sec. 2. In Sec. 3 the observational databases used in our analysis are pre- sented. The results of statistical analysis are given in Secs. 4 -- 6. Discussion and conclusion can be found in Sec. 7 . 2 MODEL OF ABSORBERS FORMATION AND EVOLUTION The main observational characteristics of absorption lines are the redshift, zabs, the column density of neutral hydro- gen, NHI , and the Doppler parameter, b, while the theo- retical description of structure formation and evolution is dealing with the mean linear number density of absorbers, nabs(z), temperature, overdensity and entropy of DM and c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 gaseous components, and with the ionization degree of hy- drogen. To connect these theoretical and observed parame- ters a physical model of absorbers formation and evolution is required. Many such models were proposed and discussed during the last twenty years (see references in Rauch 1998, and in Paper I and Paper II). However, to describe the properties of the new wider and refined sample of observed absorbers it is necessary to develop a more detailed physical model of absorbers as well. Our model includes some ideas discussed already in earlier publications. 2.1 Physical model of absorbers. In this paper we assume that: (i) The DM distribution forms an interconnected struc- ture of sheets (Zel'dovich pancakes) and filaments, their main parameters are approximately described by the Zel'dovich theory of gravitational instability applied to the CDM or WDM initial power spectrum (DD99; DD02). The majority of DM pancakes are partly relaxed, long-lived, and their properties vary due to the successive merging and ex- pansion and/or compression in the transverse directions. (ii) Gas is trapped in the gravitational potential wells formed by the DM distribution. The gas temperature and the observed Doppler parameter, b, trace the depth of the DM potential wells. (iii) For a given temperature, the gas density within the wells is determined by the gas entropy created during the previous evolution. The gas entropy is changing, mainly, due to shock heating in the course of merging of pancakes, bulk heating produced by the UV background and local sources and due to radiative cooling. (iv) The gas is ionized by the UV background and for the majority of absorbers ionization equilibrium is assumed. (v) The observed properties of absorbers are changing be- cause of merging, transversal compression and/or expansion and disruption of DM pancakes. The bulk heating and ra- diative cooling leads to slow drift of the entropy and density of the trapped gas. Random variations of the intensity and spectrum of the UV background enhance random scatter of the observed absorbers properties. (vi) In this simple model we identify the velocity disper- sion of DM component compressed within pancakes with the temperature of hydrogen and the Doppler parameter b of absorbers. We consider the possible macroscopic motions within pancakes as subsonic and assume that they cannot essentially distort the measured Doppler parameter. As compared with the model discussed in Paper II, here we take into account more accurately the evolution of pancakes after their formation. The formation of DM pancakes as an inevitable first step of evolution of small perturbations was firmly estab- lished both by theoretical considerations (Zel'dovich 1970; Shandarin & Zel'dovich 1989) and numerical simulations (Shandarin et al. 1995). Both theoretical analysis and simu- lations show the successive transformation of sheet-like ele- ments into filamentary-like ones and, at the same time, the merging of both sheet-like and filamentary elements into richer walls. Such continuous transformation of structure goes on all the time. These processes imply the existence c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Characteristics of Ly-α forest 3 of a complicated time-dependent internal structure of high density elements and, in particular, the unavoidable arbi- trariness in discrimination of such elements into filaments and sheets. Our approximate consideration cannot be applied to objects for which the gravitational potential and the gas temperature along the line of sight strongly depends on the matter distribution across this line. In this paper we use the term 'pancake' to denote structure elements with relatively small gradient of properties along the line of sight. With such a criterion, the anisotropic halo of filaments and clouds can be also considered as 'pancake-like'. The subpopulation of weaker absorbers also contains "artificial" caustics (McGill 1990) and absorbers identified with slowly expanding underdense regions (Bi & Davidsen 1997; Zhang et al. 1998; Dav´e et al. 1999). Such absorbers produce a short lived noise, which is stronger at higher red- shifts z ≥ 3. Fortunately, fraction of such absorbers does not exceed 10 -- 15% of the observed sample and they cannot significantly distort our final estimates. In this paper we consider the spatially flat ΛCDM model of the Universe with the Hubble parameter and mean density given by: H 2(z) = H 2 0 Ωm(1 + z)3[1 + ΩΛ/Ωm(1 + z)−3] , 3H 2 0 8πG Ωm(1 + z)3, H0 = 100h km/s/Mpc . hρm(z)i = Here Ωm = 0.3 & ΩΛ = 0.7 are dimensionless matter density and the cosmological term, and h = 0.65 is the dimensionless Hubble constant. (1) 2.2 Properties of the homogeneously distributed hydrogen Properties of the compressed gas can be suitably related to the parameters of homogeneously distributed gas, which were discussed in many papers (see, e.g., Ikeuchi & Ostriker 1986; Hui & Gnedin 1997; Scott et al. 2000; McDonald et al. 2001; Theuns et al. 2002a, b). Thus, the baryonic density and temperature can be taken as hnb(z)i = 2.4 · 10−7cm−3(1 + z)3(Ωbh2/0.02) , Tbg ≈ 1.6 · 104Kθ2 Here Ωb is the dimensionless mean density of baryons, Tbg & bbg are the temperature and Doppler parameter of the gas, θ4 bg(z) ∼ 1 + z describes a slow decrease of temperature with redshift for low density cosmological models, kB & mH are the Boltzmann's constant and the mass of the hydrogen atom. bbg = r 2kBTbg mH ≈ 16θbg km/s . bg(z), (2) Under the assumption of ionization equilibrium of the gas, αrn2 b ≈ nH Γγ , αr(T ) ≈ 4 · 10−13(cid:18) 104K T (cid:19)3/4 cm3 s , (3) where αr(T ) is the recombination coefficient (Black, 1981) and Γγ characterizes the rate of ionization by the UV back- ground, the fraction of neutral hydrogen is xbg = nH /nb = hnb(z)i αr(Tbg)Γ−1 γ = x0(1 + z)3, (4) 4 Demia´nski, Doroshkevich & Turchaninov x0 ≈ 6.7 · 10−8 Γ12 Ωbh2 0.02 (cid:18) 16km/s bbg(z) (cid:19)3/2 , Γγ = 10−12s−1Γ12 , For the standard power law spectrum of radiation we get J(ν) = J21 · 10−21(cid:16) νH ν (cid:17)αγ Γ12 = 12.6J21(3 + αγ )−1 . erg s−1cm−2sr−1Hz−1 , The gas entropy can be characterized by the function Fbg = Tbg hnbi2/3 = 36keV · cm2 (1 + z)2 (cid:16) 0.02 Ωbh2(cid:17)2/3(cid:18) bbg(z) 16km/s(cid:19)2 2.3 Parameters of absorbers In this section we introduce the main relations between the- oretical and observed characteristics of pancakes based on the physical model discussed in Sec. 2.1. The most impor- tant characteristics of DM pancakes are presented (without proofs) as a basis for further analysis. For more details see DD99 and DD02. 2.3.1 Characteristics of DM component The fundamental characteristic of DM pancakes is the di- mensional, µ, or the dimensionless, q, Lagrangian thickness (the dimensionless DM column density) : µ ≈ hρm(z)ilvq (1 + z) = 3H 2 0 8πG lvΩm(1 + z)2q , (6) 6.6 hΩm 0.65 0.3 Ωm , h−1Mpc = 33.8h−1Mpc lv ≈ where lv is the coherent length of initial velocity field (DD99; DD02). The Lagrangian thickness of a pancake, lvq, is de- fined as the unperturbed distance at redshift z = 0 between DM particles bounding the pancake. The actual thickness of a pancake is h ∆r = µρ−1 m = lvq(1 + z)−1δ−1, δ = ρm/hρm(z)i . (7) Here δ is the mean overdensity of a pancake above the back- ground density. Comparing the pancake surface density, µ(z), with the distance between neighboring absorbers, Dsep = c∆z H(z) = 5.5 · 103 ∆z (1 + z)3/2r 0.3 Ωm h−1Mpc , (8) we can also estimate the fraction of matter accumulated by absorbers as µ(1 + z) (1 + z)3/2 hρmiDsep ≈ 6.2 · 10−3q fabs ≃ The fraction of clustered matter characterized by hfabs(z)i can be compared with theoretical expectations (see Sec. 4.4). For Gaussian initial perturbations, the expected prob- (9) ∆z r 0.13 Ωmh2 . ability distribution function for the DM column density is e−ξ erf(√ξ) 2 √π 1 π ≈ 0.82, √ξ (cid:18) q2 hξ2i ≈ 3 4 + q2 + q2 Nqdq ≈ 1 2 hξi ≈ 0(cid:19)3/4 dξ, ξ = q 8τ 2 , (10) + 2 π , hξ3i ≈ 3 16 + 11 2π , ΘΦ = 2γ √π(γ − 1) Γ(1.5 − 1/γ) Γ(1 − 1/γ) = 4 π . c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 (DD02), where the parameter q0 ≪ 1 characterizes the co- herent length of the initial density field (see Sec. 7.5) and the dimensionless 'time' τ (z) describes the evolution of per- turbations in the Zel'dovich theory. For ΛCDM model (1) and z ≥ 2 we have τ (z) ≈ τ0(cid:16) 1 + 1.2Ωm 2.2Ωm (cid:17)1/3 1.27τ0 1 + z 1 + z ≈ (11) 1 , and for the amplitude of perturbations (Doroshkevich, Tucker & Allam 2002) . (5) τ0 ≈ (0.27 ± 0.04)r Ωmh 0.2 , ξ ≈ 1.06q(1 + z)2 0.2 Ωmh . (12) Strictly speaking, relations (6) and (10) are valid for pancakes formed and observed at the same redshift zobs = zf and after pancake formation the transverse expansion and compression changes its DM column density. However, as was shown in DD02, these processes do not change the statistical characteristics of pancakes observed at redshift zobs ≤ zf . This means that statistically we can consider each pancake as created at the observed redshift. More details are given in DD99 and DD02. For the DM dominated cosmological model the observed Doppler parameter is also closely linked with properties of the DM pancakes. Thus, in Paper II the b -- parameter was identified with the infall velocity of matter into pancakes: b ≈ vinf = u0(z)pξ2 + 2ξ , u0(z) ≈ 2.3lvH0τ 2√1 + z ≈ 310km/s (1 + z)3/2r 0.13 Ωmh2 . (13) This assumption is valid also some time after formation of the pancake but later on the pancake relaxation, compres- sion and/or expansion in transverse directions change the b -- parameter. As is seen from (13), for hξi independent from redshift (10), a systematic variation of the mean Doppler parameter with redshift could be expected. The problem was left open in Paper II (due to limited observational data set) but now, with the more representative sample of absorbers, we see surprisingly weak redshift variations of the mean observed Doppler parameter (Sec. 4.1). This means that only small fraction of observed absorbers is 'young' and is described by (13), whereas 'older' relaxed and gravitationally confined absorbers dominate this sample. The simple model of re- laxed absorbers was already considered in Paper II. Here we substantially improve it. As is well known, for an equilibrium slab of DM the depth of potential well is πGµ2 hρ(z)iδ Θ2 qΘΦ , ∆Φ ≈ where random factors ΘΦ & Θq characterizes the nonhomo- geneity of DM distribution across the slab and the evapora- tion of matter in the course of its relaxation. (14) Analysis of numerical simulations (Demia´nski et al. 2000) indicates that the relaxed distribution of DM com- ponent can be approximately described by the politropic equation of state with the power index γ ≈ 2 . In this case, we can expect that Characteristics of Ly-α forest 5 where Γ(y) is Euler function. The actual distribution of DM component across a slab and the value of ΘΦ depend upon the relaxation process which is essentially accelerated due to the pancake disruption into the system of high density clouds and filaments. This means that the factor ΘΦ ran- domly vary from absorber to absorber. At each pancake merging ∼ 10 -- 15% of matter is evap- orated due to the violent relaxation. This means that Θq ∼ 1 for absorbers formed from the homogeneously distributed matter and Θq decreases for rich absorbers formed due to several successive mergings. The Doppler parameter is defined by the depth of po- tential well and for the isentropic gas with pgas ∝ ρ5/3 trapped within the well, we get gas upon the spectrum of local UV background and it varies ran- domly with time and space. The gas temperature varies be- cause of the pancake expansion and compression and due to the pancake disruption into a system of high density clouds. As is seen from (17), these processes change the baryonic density of pancakes and we can write δb = κb(z)δ . (18) This means that the factor κb is small for absorbers formed due to adiabatic and weak shock compression because of the large difference between entropies of the background DM and the gas, and κb → 1 for richer hot absorbers formed due to strong shock compression when this difference becomes small. qΘΦ = δ0b2 bg q2 δ (1 + z) , (15) 3 4 5 4 5 Θ2 πGµ2 hρ(z)iδ b2 ≈ δ0 = ∆Φ ≈ bbg (cid:19)2 10 (cid:18) H0lv bbg(z) (cid:19)2 Ωmh2(cid:17)(cid:18) 16km/s Θδ = (cid:16) 0.13 ΩmΘΦΘ2 q ≈ 4 · 103Θδ , ΘΦΘ2 q . Variations of the gas entropy across an absorber increase the random variations of Θδ and b. 2.3.2 Characteristics of gaseous component The observed column density of neutral hydrogen can be written as an integral over a pancake along the line of sight . 1 + z 0.5 cosθ NHI = Z dx ρbxH = 2hxHihnb(z)ilvq Here hxHi is the mean fraction of neutral hydrogen and cosθ takes into account the random orientation of absorbers and the line of sight (hcosθi ≈ 0.5). We assume also that both DM and gaseous components are compressed together and, so, the column density of baryons and DM component are proportional to each other. (16) However, the overdensity of the baryonic component, δb = nb/hnbi is not identical to the overdensity of DM com- ponent, δ, (see, e.g., discussion in Matarrese & Mohayaee 2002). Indeed, the gas temperature and the Doppler param- eter are mainly determined by the characteristics of DM component (15) but the gas overdensity is smaller than that of DM component due to larger entropy of the gas. More- over, the bulk heating and cooling change the density and entropy of the gas trapped within the DM potential well. Under the assumption of ionization equilibrium of the gas (3) this process is described by the equation 1 nb dnb dt = nbαr(T )[ε(T ) − Tγ /T ], T = Tbgb2/b2 bg , (17) where the recombination coefficient αr(T ) was given in (4), Tγ ∼ (5 − 10) · 104K characterizes the energy injected at a photoionization and ε(T ) ≈ 2T 1/4 describes the radiative cooling for the bremsstrahlung emis- sion and recombination of hydrogen and helium (Black 1981). [1 − 0.3 ln T4 + 0.13T 1/3 T4 = T /104K , ], 4 4 The energy injected at the photoionization, Tγ , depends c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Under the assumption of ionization equilibrium of the gas (3) and neglecting a possible contribution of macroscopic motions to the b-parameter (T ∝ b2), for the fraction of neutral hydrogen and its column density we get: hxHi = x0κb(z)δβ−3/2(1 + z)3Θx, NHI = N0qδβ−3/2(1 + z)5, N0 = 5 · 1012cm−2ΘH , bbg(z) (cid:19)3/2 0.02 (cid:19)2(cid:18) 16km/s cosθ (cid:18) Ωbh2 Ωmh2 hcosθi κbΘx Γ12 β = b/bbg , ΘH = (19) 0.13 , where Γ12, bbg and x0 were defined in (2) and (4) and the factor Θx describes the nonhomogeneous distribution of ion- ized hydrogen along the line of sight. 2.3.3 Absorbers characteristics Eqs. (15) and (19) link the observed and other physical char- acteristics of absorbers. For the DM column density, q, the average gas entropy, Σ = ln(Fs), and overdensity, δ, we get: q3 = NHI N0δ0 β7/2 (1 + z)6 , δ = δ0 q2 β2 (1 + z) , (20) exp(Σ) = Fs = β2/δ2/3 b = κ−2/3 b β2/δ2/3 . Here we use the standard equation of state T /Tbg = β2(z) = Fs(z)δ2/3(z) . The precision of these estimates is moderate and the main uncertainties are generated by the unknown cosθ and parameters ΘΦ, Θq, Θx, κb and Γγ , which vary -- randomly and systematically -- from absorber to absorber (estimates of q and δ are independent from bbg). These variations distort parameters defined in (20) as follows: 1 Θδ Θ2 H Θδ (cid:19)2/9 (cid:18) Θ2 H , ΘHΘδ δ3 ∝ , Fs ∝ κ−2/3 q3 ∝ Independent estimates of these uncertainties can be ob- tained from the analysis of the redshift distribution of ab- sorbers (Sec. 6). (21) . b However, comparing the average DM column density of pancakes, hqi, with expectations (10) we can restrict the possible redshift variations of hΘH Θδi and estimate the com- bined uncertainty introduced by unknown factors. As is seen from (20), ξ3 ≈ (1 + z)6 q2 = (cid:16) NHI 2 · 1016cm−2(cid:17)(cid:18) b 16km/s(cid:19)7/2 G12 , 6 Demia´nski, Doroshkevich & Turchaninov G12(z) = Γ12 κbΘxΘΦΘ2 q cosθ 0.5 (cid:18) Ωmh2 0.13 Ωbh2 0.02 (cid:19)2 . (22) larger than the coherent length of initial density field, this number density can be approximated by the expression This relation shows that, for statistically homogeneous sam- ple of absorbers with hξ3i = const., hcosθi = 0.5, we can estimate the redshift variations of hG12(z)i as follows: hG12(z)i ∝ (cid:28) q(cid:29) ∝ hξ3i/hNHI b7/2i . κbΘxΘΦΘ2 Γ12 Precision of these estimates is limited because of strong ran- dom scatter of the product NHI b7/2 . 2.3.4 Regular and random variations of absorber characteristics The most fundamental characteristic of absorbers is their DM column density, ξ ≈ q(1 + z)2. It describes the forma- tion and merging of pancakes, is only weakly sensitive to the action of random factors and defines the regular variations of the Doppler parameter, b, overdensity, δ, and entropy, Fs. These parameters include also a random component which integrates the evolutionary history of each pancake and the action of random factors discussed in the previous subsec- tion. If the structure of a relaxed pancake can be described by the politropic equation of state with the effective power index γ then we can discriminate the regular and random variations of b, δ, & Fs and introduce the reduced character- istics of relaxed absorbers, υ, ∆ & S: υ = ln[βξ(1/γ−1)] , ∆ = ln[(1 + z)3δ/ξ2/γ ] = ln(δ0) − 2υ , S = ln[(1 + z)−2Fsξ2(5/3−γ)/γ ] = const. + 10υ/3 . (23) These relations indicate that, in fact, our approximate de- scription use only one random characteristic, namely, υ, which is expressed through observed parameters as follows: υ = 7 − γ 6γ ln β − γ − 1 3γ ln(cid:16) NHI N0δ0(cid:17) . (24) As is shown in Secs. 4.5.2, the choice of a suitable value of γ allows to minimize the correlation between ξ and υ. Of course, this reduction is approximate and it can be achieved only statistically for a given sample of absorbers. 2.4 Mean number density of absorbers Following Paper I we will characterize the 1D mean number density of absorbers by the dimensionless function nabs = c H0 hli−1 = H(z) H0 dN (z) dz . (25) Here dN (z) is the mean number of absorbers between z and z + dz, hl(z)i is the mean free path between absorbers at redshift z and c is the speed of light. When absorbers are identified with the Zel'dovich pancakes, this 1D mean num- ber density can be linked with the fundamental character- istics of the cosmological model and moments of the initial power spectrum (DD02 & Sec. 7.5). As was shown in DD02, for Gaussian initial perturba- tions and for richer DM pancakes with sizes significantly c √3(1 + z)2 16πτ (z)√q0 erf(√ξthr) c Wp(ξthr)(1 + z)2 , (26) H0lv (cid:21) , hq(ξthr)i π exp(ξthr)[1 − erf2(√ξthr)] 4√πξthrerf(√ξthr) + 2 exp(−ξthr) nabs ≈ hq(ξthr)i = 4τ 2(cid:20)1 + Wp(ξthr) ≈ 0.5[1 − erf(pξthr)] , where ξthr(z) = qthr/8τ 2(z), qthr(z) is the minimal (thresh- old) DM column density of pancakes, Wp(ξthr) and hq(ξthr)i are the fraction of matter and the mean DM column density for pancakes with q ≥ qthr and the factor (1 + z)2 describes the cosmological expansion of pancakes population. The expression (26) contains only one fitting parame- ter, ξthr, and it describes quite well the evolution of richer pancakes. However, application of this relation to evolution of weaker pancakes is problematic as for ξthr ≪ 1 we get from (26): nabs ≈ 55(1 + z)4 , and this relation does not contain any fitting parameters. The formation of pancakes with small q is suppressed due to the small scale cutoff in the initial power spectrum. For pancakes with small qthr the redshift variations of the mean number density is described by the expression H0lv √ξthr nabs ≈ (DD02) where again ξthr(z) = qthr(z)/8τ 2 and the value q0 characterizes the coherent length of the initial density field (see Sec. 7.5). exp(−ξthr) , (27) Both expressions, (26) and (27), are derived for the Gaussian initial perturbations. They are valid for DM pan- cakes formed due to both linear and nonlinear compressions and allow for merging of pancakes. Moreover, as was shown in DD02, if the observed redshift is identified with the red- shift of pancake formation then both relations remain the same even when the transverse compression and expansion of pancakes are also taken into account. This means that they partly account for the pancake disruption as well. The relation (27) describes also the impact of the gaseous pressure on the formation of observed absorbers. In this case, the parameter q0 must be calculated with the power spectrum corrected for the Jeans damping of small scale perturbations in the baryonic component (see Sec. 7.5). 2.5 Observational restrictions The completeness of the observed samples of absorbers is re- stricted by the condition NHI ≥ Nthr ≈ 1012cm−2 what in turn distorts the characteristics of observed absorbers and makes it difficult to compare them with the theoretical ex- pectations discussed in Secs. 2.3 and 2.4 . This means that these expectations should be corrected for the impact of the threshold column density of observed absorbers. Thus the investigation of spatial distribution of galaxies in the SDSS EDR (Doroshkevich, Tucker & Allam 2002) results in estimates of typical parameters of galaxy walls at small redshifts as hqi ≈ 0.4, hδi ≈ 3 . c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 hβi ≈ 20, (28) With these data the expected column density of neutral hy- drogen within the typical wall (16) is NHI ≈ 0.02N0 ≈ 0.1Nthr ΘH ≤ Nthr . This means that even so spectacular object as the 'Greet Wall' does not manifest itself as absorbers. Rare absorbers with b ≥ 100 km/s, NHI ∼ 1013cm−2 and ξ ∼ 1.5 observed at z ∼ 1.5 -- 4 can be associated with embryos of wall -- like elements of the Large Scale Structure of the Universe (see Sec. 5 for more detailed discussion). q ≈ hqi ≈ (1 + z)−2, For absorbers associated with 'young' pancakes the Doppler parameter is given by (13) and for the richer 'young' pancakes with the mean DM column density we get β ≈ 35(1 + z)−3/2 , ξ ≈ 1, δ ≈ 3 Θδ, NHI ≈ 10−2(1 + z)21/4N0Θδ , and such pancakes can be seen as absorbers already at red- shifts z ≥ 1. have For more numerous poorer pancakes with q ≤ hqi we (29) Characteristics of Ly-α forest 7 Table 1. QSO spectra used zem zmin zmax 0000 − 2601 0055 − 2592 0014 + 8133 0956 + 1223 0302 − 0033,2 0636 + 6803 1759 + 7544 1946 + 7665 1347 − 2462 1122 − 4412 2217 − 2822 2233 − 6066 1101 − 2642 0515 − 4412 2126 − 1587 1700 + 6428 1225 + 3179 1331 + 17010 4.11 3.66 3.41 3.30 3.29 3.17 3.05 3.02 2.63 2.42 2.41 2.24 2.15 1.72 3.26 2.72 2.20 2.10 3.4 3.0 2.7 2.6 2.6 2.5 2.4 2.4 2.1 1.9 1.9 1.5 1.6 1.5 2.9 2.1 1.7 1.7 4.1 3.6 3.2 3.1 3.1 3.0 3.0 3.0 2.6 2.4 2.3 2.2 2.1 1.7 3.2 2.7 2.2 2.1 No of HI lines 431 534 262 256 356 313 307 461 361 353 262 293 277 76 130 85 159 69 (30) 1. Lu et al. (1996), 2. unpublished, courtesy of Dr. Kim 3. Hu et al., (1995), 4. Djorgovski et al. (2001) 5. Kirkman & Tytler (1997), 6. Cristiani & D'Odorico (2000), 7. Giallongo et al. (1993), 8. Rodriguez et al. (1995), 9. Khare et al. (1997), 10. Kulkarni et al. (1996). can be distorted due to the limited representativity of our sample and a nonhomogeneous redshift distribution of ob- served absorbers plotted in Fig. 1 . However, weak redshift variations of the functions hbi, hlg(NHI )i & hG12i (Sec. 4.1) show that the influence of varia- tions of Γ12 on the properties of absorbers is partly compen- sated by variations of κb, Θx, Θq & ΘΦ introduced in Secs. 2.3.1 & 2.3.2 . As is found in Sec. 4.2, a reasonable descrip- tion of the mean absorbers characteristics is achieved when the function hG12i (22) is hG12i ≈ G0(1 − 0.17z), G0 = 50 . (32) Comparison of this function with estimates of Γ12 (Scott et al. 2000; McDonald & Miralda-Escude 2001) allows to estimate the unknown factors as qi ∼ (0.1 − 0.05)(cid:18) Ωmh2 hκbΘxΘΦΘ2 More accurate estimate of these factors can be obtained with numerical simulations. 0.02 (cid:19)2 0.13 Ωbh2 . 3 THE DATABASE. The present analysis is based on the 18 spectra listed in Table 1. The available Ly-α lines were arranged into three samples. The richest sample, S12 14 , includes 4369 absorbers with 1015cm−2 ≥ NHI ≥ 1012cm−2 from the first 14 high resolution spectra. It can be partly incomplete. For compar- ison, we use the most reliable sample S13 18 which includes 2643 absorbers from all 18 spectra for 1013cm−2 ≤ NHI ≤ 1015cm−2 . These lines are more easily identified and they are not so sensitive to outer random influences. The sample 14 contains 2126 weaker lines with NHI ≤ 1013cm−2 from W 12 the first 14 QSOs. It is mainly used to characterize possible δ ≈ 4q(1 + z)2Θδ , β ∼ 30√q(1 + z)−1/2, NHI ≈ 2.5 · 10−2N0q5/4(1 + z)31/4Θδ ≈ (cid:16) q NthrΘδΘH , and such pancakes become visible for z ≥ 1.5 only. sorbers show that: 0.02(cid:17)5/4(cid:16) 1 + z 2.5 (cid:17)31/4 These rough estimates of expected characteristics of ab- (i) At redshifts z ≤ 1.5 we can observe mainly old pan- cakes formed at higher redshifts which kept the measurable column density of neutral hydrogen up to small redshifts. (ii) The expected number density of observed absorbers rapidly increases at redshifts z ∼ 1.5 -- 2 when even rela- tively poor 'young' pancakes can pass over the observational threshold and become seen as weak absorbers. This means that our analysis can be applied to ab- sorbers observed at z ≥ 1.5 -- 2 when the impact of the observational threshold becomes moderate. 2.6 Variations of the UV background Direct estimates of the intensity of UV background through the proximity effect (Scott et al. 2000) give Γ12 ≈ 2 ± 1 at z ∼ 2.5 -- 3 while McDonald & Miralda-Escude (2001) got Γ12 ∼ 0.7 − 0.4 at the same z. Indirect estimates of Γ12 (Songaila 1998) demonstrate a probable sudden drop of the UV intensity of about 3 times at z ∼ 3. This effect can be related to the strong ionization of HeII at these redshifts (Jacobsen et al. 1994; Zeng, Davidsen & Kriss 1998; Theuns at al. 2002a, b). As is shown in Sec. 6.2, possible redshift variations of Γ12 are also seen in the redshift distribution of weaker ab- sorbers at z ≈ 2.5. These variations can be fitted by the expression Γ12 = a0 + 1 2 h1 + a0 − 1 a0 + 1 th(cid:16) z − zγ 0.16 (cid:17)i , (31) where a0 ∼ 1.5 -- 3 and zγ ∼ 2.5 -- 3 characterize the am- plitude and the redshift of these variations. These estimates c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 8 Demia´nski, Doroshkevich & Turchaninov Figure 1. Redshift distribution of the fraction of observed ab- 14 (green points) and S13 sorbers for samples W 12 18 (blue points). 14 (red points), S12 variations of the UV background. To decrease the scatter of absorbers characteristics we exclude from these samples 45 absorbers with b ≥ 90km/s . As is seen from Fig 1, for all samples the redshift distri- bution of absorbers is nonhomogeneous and the majority of absorbers are concentrated at 2 ≤ z ≤ 3 . This means that some of the discussed absorbers characteristics are derived mainly from this range of redshifts. Absorbers at z ≥ 3.5 were identified mainly from the spectrum of QSO 0000-260 (Lu et al. 1996) and here the line statistics is insufficient. 4 STATISTICAL CHARACTERISTICS OF ABSORBERS Γ12 E = 0.1. qi = 0.5, ΘH = D κbΘx In this Section, the functions q, ξ, δ & Fs are found for the ΛCDM cosmological model (1) with bbg = 16km/s, Ωbh2 = 0.02 and the function hG12(z)i given by (32) with G0 = 50, Θδ = hΘΦΘ2 (33) Some quantitative results of this Section depend on the com- pleteness and representativity of the samples and the quite arbitrary choice of Θδ & ΘH (33) through the relations (21). As was noted above, for each absorber these factors change in the course of absorbers evolution and, in fact, the ac- cepted values are averaged over the sample. For the sample S12 14 , ∼ 10-15% of the weaker absorbers could be related to the artificial "noise". For majority of QSO spectra used in our analysis the scatter of observed NHI is less then 10 per cents, while the scatter of the measured Doppler parameters, b, can achieve 20 -- 30 per cents. Non the less, dispersions of absorbers char- acteristics discussed below are defined mainly by their broad distribution functions and by the completeness of the sam- ples. Because of this, we discuss in this Section uncertainties of only the more interesting quantitative characteristics of absorbers. Figure 2. Redshift variations of hG12i (top panel), Doppler pa- rameter, hbi (middle panel) and hlg(NHI )i (bottom panel) for samples S13 14 (green points). Mean values (34) and fit (32) are plotted by straight lines. 18 (blue points) and S12 samples S12 prisingly weakly vary with redshift, 14 and S13 18 , the mean values hbi and hlgNHIi sur- (34) hbi = (31.7 ± 1.6)km/s, hlgNHIi = (13.6 ± 0.06), hbi = (28.6 ± 1.6)km/s, hlgNHIi = (13.1 ± 0.22), respectively. At the same time, the mean Doppler parameter systematically shifts from hbi = (25.6 ± 2.9)km/s for the sample W 14 12 up to hbi = (38.4 ± 2.7)km/s for 660 absorbers with NHI ≥ 1014cm−2 what indicates a weak correlation of observed properties of absorbers. Detailed discussion of observed characteristics of absorbers can be found in Kim, Cristiani & D'Odorico (2002), Kim et al. (2002). The small variations of hbi with the redshift are accom- panied by similar small variations of dispersion σb ∼ 0.5hbi which also do not exceed ∼ 8%. This fact indicates that the broad distribution function of b is weakly dependent upon the redshift. The same is valid for the depth of potential wells, h∆Φi (14), and illustrates a correlation between the DM column density, q, and the overdensity of absorbers, δ described by (20). The nature of such a weak redshift evolu- tion of these statistical characteristics of absorbers remains a mystery. The function hG12i/G0 is also plotted in Fig. 2 together with the rough fit (32). Its weak redshift dependence is a natural consequence of the weak redshift variations of hbi and hlgNHIi. Its deviations from (32) correlate with the ab- sorbers distribution plotted in Fig 1 . 4.1 Redshift variations of the mean observed density, overdensity and entropy of absorbers 4.2 Redshift variations of the mean DM column characteristics Redshift variations of two observed characteristics of ab- sorbers, hbi, and hlgNHIi, are plotted in Fig. 2. For both Other characteristics of absorbers depend upon the physical model. For the model discussed in Sec. 2 (Eq. 20) with the function hG12i (22, 32) the redshift variations of the mean c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Characteristics of Ly-α forest 9 6.3 from the analysis of redshift distribution of absorbers. This disagreement requires more detailed investigation and, perhaps, redefinition of the factors G0, ΘH and Θδ . 14 can be roughly fitted as follows: hδi ≈ 4.7ζ −2.5 , ζ = (1 + z)/4 . As is seen from Fig. 3, the redshift variations of func- tions hδi and hFsi characterizing the overdensity of DM com- ponent and entropy of gaseous component for both samples 18 and S12 S13 hδi ≈ 7.5ζ −3.1, hFsi ≈ 1.93ζ 1.6, These fits emphasize the general tendencies of absorbers evo- lution. Comparison with the background density (2) and en- tropy (5) shows that the mean density of absorbers weakly depends on the redshift and the mean entropy weakly in- creases with time. These results indicate that our sample is dominated by long lived relaxed absorbers. (36) These estimates may be rescaled according to (21). 4.3 The mean size of absorbers 18 , we have Our model of absorbers ( see Sec. 2) allows also to estimate roughly the real size of absorbers along the line of sight, ∆r. Due to the strong correlation between q and δ, this size should be obtained by averaging ∆r as given by (7). For the model parameters given in (33) and for both samples, S12 14 and S13 h∆ri ∼ (80 − 100)h−1kpc , (37) and this size increases up to ∼ 150h−1 kpc for weaker ab- sorbers of the sample W 12 14 . For all samples the PDFs of the sizes are similar to the Gaussian function with a dispersion σr ∼ 0.5h∆ri. sociated with the Doppler parameter, b, as follows: The mean size of absorbers in the redshift space is as- h∆rbi = 2hbi/H(z) ≈ 130h−1kpc(cid:16) 4 1 + z(cid:17)3/2r 0.3 Ωm . As usual, it is larger than that in the real space (37). This difference agrees with the domination of relaxed gravitation- ally bound absorbers in the sample. The expected mean transverse size of absorbers is com- parable with their Lagrangian size along the line of sight and was roughly estimated in DD02 as 0.6lv lvτ 2 1 + z ≈ (1 + z)3 ≈ 300h−1kpc(cid:16) 4 h∆rti ≈ 5.2 Difference between estimates h∆ri and h∆rti agrees with expectations of the Zel'dovich theory and indicates that such absorbers could be unstable with respect to the disruption into a system of denser less massive clouds with ∆rt ∼ ∆r (see, e.g., Doroshkevich 1980; Vishniac 1983). 1 + z(cid:17)3 . (38) 4.4 The mean matter fraction accumulated by absorbers and the mean absorber separation The mean matter fraction accumulated by absorbers, hfabs(z)i, and the mean absorber separation, hDsep(z)i, de- fined by (8 & 9), are plotted in Fig. 4 for the samples W 12 14 , 14 & S13 S12 18 . The estimates of hfabs(z)i are sensitive to small measured line separations, ∆z ∼ 10−4, which are unreli- able. For richer absorbers of the sample S13 18 , the estimates Figure 3. Functions hξi = (1 + z)2hqi (top panel), hFsi (middle panel) and the overdensity hδi, (bottom panel) are plotted vs. redshift, z, for samples S12 18 (blue points). Fits (35, 36) are plotted by lines. 14 (green points), and S13 DM column density, hξi, the mean entropy and overdensity, hFsi and hδi, are plotted in Fig. 3. The mean DM column density of absorbers, hqi & hξi ≈ hqi(1 + z)2, is the most stable parameter. In principle, hξi does not change due to the formation and merging of ab- sorbers and due to their transverse compression and/or ex- pansion (DD02). The differences between the expected and measured hξi characterize, in fact, the influence of disre- garded factors, mainly the completeness and representativ- ity of the samples, the pancake relaxation and the scatter of the function G12. As is seen from comparison of Fig. 1 and Fig. 3, for both 14 variations of measured hξ(z)i around 18 and S12 samples S13 the mean values (35) For the sample S12 hξi ≈ (0.65 ± 0.09) , hξi ≈ (1. ± 0.09), are roughly correlated with the redshift distribution of ab- sorbers plotted in Fig. 1 what emphasizes the limited rep- resentativity of the sample used in the analysis. At small redshifts, z ≤ 2, the influence of the observational restric- tions (Sec. 2.5) increase the measured value of hξi. 14 the measured hξ(z)i ≈ 0.65 is close to the expected value hξ(z)i ≈ 0.82 what verifies the choice of the amplitude of hG12i in (32). However, the estimates of hξ(z)i depend also upon the choice of the amplitude of initial perturbations, τ0 (12), known only with a moderate precision. For the samples S13 18 , hξ(z)i increases due to the rejection of weaker absorbers in this sample. 18 , the minimal DM column density is found to be qmin ∼ 10−2 and it only weakly de- pends upon the redshift. This result is especially sensitive to the random noise, it is based on poor statistic and is not re- liable because our fit of hG12i provides a reasonable descrip- tion for the sample as a whole only. This value is smaller than the estimate hqthri ≈ 3 · 10−2 (54) obtained in Sec. For both samples, S12 14 & S13 c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 10 Demia´nski, Doroshkevich & Turchaninov with richer samples of observed absorbers and by compari- son with simulations. As is well known, at redshifts z ≥ 4 the absorbers be- gin to overlap and their separation becomes problematic. Analysis of the evolution of the mean absorbers separa- tion allows to estimate quantitatively this effect. For this purpose, we can compare the mean separation of absorbers at various redshifts, hDsep(z)i, given by (8) with the mean Doppler parameter which characterizes the observed thick- ness of absorbers. For samples S12 18 these parameters are roughly fitted by power laws: 14 and S13 hDsepi ≈ (cid:16) 4 H0hDsepi 1 + z(cid:17)2.7 ≈ (cid:16) 6.1 hbi 1 + z(cid:17)2.6 , respectively. h−1Mpc, ≈ (cid:16) 5.3 H0hDsepi hbi h−1Mpc , 1 + z(cid:17)2.7 ≈ (cid:16) 8 1 + z(cid:17)2.9 , (39) Figure 4. Mean fractions of matter accumulated by absorbers, hfabsi, (top panel) and the mean absorber separations, hDsepi, (bottom panel), are plotted vs. redshift, z, for samples W 12 14 (red points), S12 18 (blue points). Fits (39) are plotted by solid and dashed lines. 14 (green points), and S13 of fabs(z) based on the expression (9) are also unreliable because they neglect the transverse compression and expan- sion of matter and formation of high density clouds. So, the representativity and precision of measured hfabs(z)i are lim- ited. The estimates of hDsep(z)i are less sensitive to the first factor but are also influenced by the second one. The ran- dom scatter of the measured hfabs(z)i and hDsep(z)i could be caused by the limited representativity of the samples at small and higher redshifts. In spite of uncertainties these results are interesting in some respects. Thus, the regular growth of hfabs(z)i and decrease of hDsep(z)i with redshifts for the sample S13 18 indi- cate the progressive disruption of richer absorbers and their transformation into system of high density clouds. Indeed, due to small cross -- section of such clouds, they are rarely cross the line of sight and their formation increases the sep- aration of richer observed absorber and even can decrease the measured hfabsi for this sample. Both the significant fraction of matter seen as numerous weaker absorbers in the sample W 12 14 and its weak redshift variation point in favor of the domination of absorbers merg- ing as compared with the formation of new absorbers from the background matter. The merging remains unchanged the hfabs(z)i but progressively increases hDsep(z)i with time. These variations can be enhanced by the observational re- strictions, NHI ≥ Nthr, discussed in Sec 2.5 . For all samples the fraction hfabs(z)i ∼ 0.8 exceeds the theoretical expectation hfabs(z)i ∼ 0.5 -- 0.6 for z ≃ 3 (DD02 and Sec. 7.6) and is close to the maximal fraction, fmax ≃ 0.875, which can be compressed in the Zel'dovich' theory. This disagreement indicates the limited applicability of the one dimensional expression (9) and the limited preci- sion of our model of absorbers. These results must be tested These results verify that at z ≥ 4 absorbers effec- tively overlap and the system of individual absorbers is transformed into continuous absorption imitating the Gunn -- Peterson effect. 4.5 Distribution functions of absorber parameters The observed probability distribution function, PDF, of the DM column density of absorbers can be compared with the expected one (10). However, we have no theoretical expres- sion for the PDFs of the Doppler parameter, overdensity and entropy of absorbers because the action of random factors cannot be satisfactory described. Indeed, the relaxation of DM pancakes depends upon unknown internal structure of pancakes, the adiabatic com- pression and/or expansion of an absorber changes its over- density and temperature. On the other hand, the radiative cooling and bulk heating lead to the drift of the gas entropy and overdensity but leaves unchanged the depth of poten- tial well formed by DM distribution. Merging of pancakes increases more strongly the depth of potential well and the gas entropy but the overdensity of the gaseous component increases only moderately. All the time, the temperature and overdensity of trapped gas are rearranged in accordance with the condition of hydrodynamic equilibrium across the pancake. Because of this, our discussion of absorbers evolution has only phenomenological character. The PDF of the DM column density, Nq, is plotted in Fig. 5 for the sample S12 14 together with the fits (10) and (40). The distribution func- tions of the reduced Doppler parameter, υ, are plotted in Fig. 5 for samples S13 14 together with the Gaussian fits. 18 and S12 4.5.1 Distribution functions of the DM column density The PDF for the DM column density of pancakes, Nq(q), plotted in Fig. 5 for samples S12 14 is the most interesting because it can be compared with the theoretically expected PDF (10) which, in particular, is sensitive to the coherent length of initial density field, q0. As is seen from (10), the PDF Nq(ξ), with ξ ≈ q(1 + z)2, does not depend on the redshift and, so, the joint PDF can be obtained for absorbers observed at all redshifts. c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Characteristics of Ly-α forest 11 Table 2. Parameters of three subpopulations of absorbers for the sample S12 14 Nabs hfsi hlgNHI i hbi rbH full sh cl exp 4369 1940 1664 765 1.00 0.44 0.38 0.18 13.1 13.0 13.5 12.3 29 37 23 18 0.30 0.33 0.67 0.42 γ 1.8 1.8 1.6 3.5 hυi 0.8 1.0 0.6 1.3 fs is the fraction of absorbers in a subpopulation, rbH is the linear correlation coefficient of b and logNHI . ∆ = ln[(1 + z)3δ ξ−1.1] = ln δ0 − 2υ , S = ln[(1 + z)−2Fsξ−0.15] = const. + 3.33 υ . (41) These expressions correspond to Eq. (23) with the effec- tive power index of compressed DM component γ = 1.8. For the sample S13 18 we have γ = 2.1. Both values are close to γ = 2 discussed in Sec. 2.3.1 . With such defini- tion, we get the negligible correlations, ≤ 0.01, between ξ and other reduced characteristics and a strong correlation ≈ 1 between υ and ∆ & S. The distribution function Nυ plotted in Fig. 5 is well fitted by the Gauss function with hυi = 0.81 ± 0.04, συ ≈ 0.11hυi what is consistent with the assumption of approximate equilibrium of majority of the observed absorbers. These results mean that, in fact, the joint action of all random factors is characterized by the one random func- tion υ which can be directly expressed through the observed parameters as follows υ ≈ 0.48 ln β − 0.15 ln(NHI /N0/δ0), More detailed analysis requires the discrimination of sub- populations of absorbers with different evolutionary histo- ries. γ = 1.8 . (42) 4.6 Three subpopulations of absorbers As a first step of more detailed investigation of absorbers evolution, we can approximately separate three subpopula- tions of absorbers with different overdensities and entropies. Due to continuous distribution of both δ & Fs for the full samples and because both these values depend upon un- known random factors (21), the separation is quite arbitrary and characteristics of subpopulations depend upon the sam- ple used in the analysis and the parameters (33). Even so, the redshift variation of mean parameters of subpopulations allows to trace roughly their evolutionary history and to de- scribe the main stages of absorbers formation and evolution. The first subpopulation, let us call it "sh", with δ ≥ 1 and Fs ≥ 1 is formed due to the merging and shock compres- sion. The second one, called here "cl", includes low entropy absorbers with δ ≥ 1, Fs ≤ 1 . The third one, called here "exp", contains peculiar absorbers with δ ≤ 1. This choice of the threshold parameters allows to sep- arate absorbers formed and observed at the same redshifts, zf ≈ zobs. However, due to evolution of the background den- sity (2) and entropy (5), the functions δ & Fs defined by (20) are shifted with time even when the overdensity and entropy of a long lived absorber are not changed. This shift leads to some artificial intermixture of subpopulations. For the sample S12 14 the mean parameters of these sub- Figure 5. Top panel: PDFs of the DM column density, Nq, for the sample S12 14 . Fits (10) and (40) are plotted by green and blue lines. Bottom panel: PDF Nυ, for the samples S12 14 (red line) and S13 18 (green line). Gaussian fit is plotted by blue line. For ξ ≥ 0.5hξi, the PDF plotted in Fig. 5 is well fitted by the function (10) for q0 = 0 and hξi = 0.57 what is close to hξi = 0.65 (35). The deficit of observed absorbers with ξ ≤ hξi can be related to incompleteness of the observed sample for small ξ. This result verifies the self consistency of the physical model used here and the Gaussianity of initial perturbations. However, as it is seen from (10), this deficit of weaker absorbers can be related also to the suppression of the forma- tion of poorer pancakes caused by the correlations of small scale initial perturbations and Jeans damping described by the parameter q0 in (10). Indeed, the observed PDF is well fitted also by the function Nq = 0.18 erf(√y) √y exp(−y) (1 + 0.53/y) , y = 1.15 , ξ hξi (40) what is consistent with rough estimate q0 ∼ 0.05. This value depends on the statistics of absorbers at small ξ ≤ hξi and can be considered as an upper limit of q0 only. 4.5.2 Distribution functions of the reduced Doppler parameter As was noted above, the theoretical description for the PDFs of the reduced Doppler parameter, overdensity and entropy of the gas is problematic because they depend upon many random factors. Therefore, here we will restrict our analysis to the fits of observed PDFs, Nυ, taking also into account the correlations of the measured ξ, b, δ & Fs. To adjust the reduced characteristics of pancakes in- troduced in Eq. (23), we will minimize the correlation of ξ with υ, ∆, & S. For the sample S12 14 , the adjusted reduced characteristics are defined as follows: γ = 1.8, υ = ln(βξ−0.44) , c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 12 Demia´nski, Doroshkevich & Turchaninov Figure 6. The mean Doppler parameter, hbi (top panel), the mean entropy, hFsi (middle panel), and the mean overdensity hδi (bottom panel), are plotted vs. redshift, z, together with fits (43, 44, & 45) for the "sh" (red), "cl" (blue) and "exp" (green) subpopulations of the sample S12 14 . Figure 7. The DM column density (top panel) and the relative number fractions of absorbers (bottom panel) for the "sh" (red lines), "cl" (blue lines) and "exp" (green lines) subpopulations of the sample S12 14 are plotted vs. redshift, z . absorbers with populations are listed in Table 2, where rbH is the linear correlation coefficient of b and lg NHI . Redshift variations of hbi, hlgNHIi, hFsi, hδi, hqi, hξi and the relative num- ber fractions of absorbers in these subpopulations, hfmi, are plotted in Figs. 6 & 7 and given in (43, 44 & 45). The subpopulation "sh" is composed of the richest and hottest absorbers with (43) hFsi ≈ 3.0ζ 0.6, hbi ≈ (37 ± 2.7)km/s , hlgNHIi ≈ 13.0 ± 0.4, hξi ≈ 0.8 ± 0.15, where ζ = (1 + z)/4. As is seen from Fig. 7, for this sub- population the decrease of the relative number fraction of absorbers, hfmi, with z is accompanied by the growth of the DM column density, hqi. Such variations are typical for absorbers formed due to pancake merging and strong shock compression of the matter. hδi ≈ 3.75/ζ , The subpopulation "cl" includes absorbers with (44) hδi ≈ 8.7ζ −1.6 . hFsi ≈ 0.6 ± 0.08, hbi ≈ (23.1 ± 1.1)km/s , hlgNHIi ≈ 13.5 ± 0.37, hξi ≈ 0.69 ± 0.1, Such behavior of the mean characteristics indicates that this subpopulation can be dominated by absorbers formed in low entropy regions at zf ∼ zobs due to adiabatic and weak shock compression. Such compression provides higher overdensity than the strong shock compression and does not change the entropy of compressed matter. For these absorbers the ra- diative cooling can also be more important. This subpop- ulation includes also some fraction of long lived absorbers formed at zf > zobs due to merging and strong shock com- pression. This intermixture is caused by redshift variations of background parameters (2) and (5). The peculiar subpopulation "exp" accumulates poorer (45) hFsi ≈ 2.1ζ 2.5, hδi ≈ 0.6 ± 0.08 . hbi ≈ (18 ± 5.4)km/s, hlgNHIi ≈ 12.3 ± 0.3, hqi ≈ 1.1 · 10−2, Such behavior of hqi, hδi &hFsi shows that these absorbers could be related to the unstable poorer pancakes formed within low density regions (see, e.g., Zhang et al. 1998). They are observed mainly at higher redshifts, they expand together with the background and disappear at low redshifts. Large values of the factor γ = 3.5 and of the reduced Doppler parameter hυi = 1.3 (Table 2), verify the peculiar character of this subpopulation. Moreover, the analysis performed in Sec. 6.3 indicates that, for weaker absorbers dominating this subpopulation, at least the DM column density, hqi, can be underestimated. This means that other parameters of the subpopulation can also be estimated unreliably. In particular, if these absorbers are unstable then the random factors become time depen- dent and the estimates (45) must be corrected. Fortunately, the fraction of such absorbers is small and its influence on other estimates is limited. In spite of the quite arbitrary separation of subpop- ulations these results illustrate the action of evolutionary factors mentioned above. The objective character of the dis- crimination is seen from the comparison of observed param- eters, hlgNHIi and hbi, and their linear correlation coeffi- cient, rbH , listed in Table 2 for the full samples and for the separated subpopulations. These results confirm that majority of observed ab- sorbers represent gravitationally bounded pancakes formed in the course of adiabatic and shock compression at zf ≥ zobs. They demonstrate similarity of evolutionary history and generic origin of subpopulations "sh" and "cl". More detailed statistical investigation of absorbers evolution re- c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 quires finer discrimination of subpopulations and therefore richer set of observed absorbers is needed. 4.7 Absorbers as a test of background properties Some absorbers characteristics can be used to estimate the redshift variations of mean properties of homogeneously dis- tributed hydrogen (see, e.g., Hui & Gnedin 1997; Schaye et al. 1999, 2000; McDonald et al. 2001). Thus, weak redshift variations of hFsi (44) suggest also weak redshift variations of Tbg (2). However, such estimates are inevitably approx- imate and their significant scatter is caused by action of many factors discussed above. To obtain more stable results Schaye et al. (1999, 2000) consider a cutoff at small b in the distribution of b(NHI ). Indeed, relations (20) can be rewritten as follows: b = bbgκ3/8 b δ1/8 0 (cid:18) Fs (1 + z)2(cid:19)9/16 N0 (cid:17)1/4 (cid:16) NHI . (46) Parameters N0, δ0, bbg & κb were defined by (2, 15, & 19). Ev- idently, for Fs = const. this relation is identical to the equa- tion of state b2 = const.δ2/3. It demonstrates that, for the subpopulation of absorbers formed at zf ≈ zobs with the same Fs, the observed parameters z, b and NHI are strongly correlated and, for a given NHI , b decreases with Fs. How- ever, statistics of absorbers near the cutoff is small, the ac- tion of factors discussed above erodes the cutoff and makes its interpretation less reliable. (47) For absorbers selected from the sample S12 14 in four red- shift intervals, z ≤ 2 ≤ z ≤ 2.5 ≤ z ≤ 3 ≤ z, the boundary of the observed distribution b(NHI ) can be roughly fitted by the relation b ≈ b0(1 + z)−1(NHI /1012cm−2)γb . For z ≤ 3 we have γb ∼ 0.25 that coincides with (46) but the value of b0 ∼ 22 km/s is ∼ 2 times smaller, what is expected from (46). This means that in the framework of the model under consideration the cutoff can be naturally related to a fraction of low entropy absorbers with Fs ∼ 0.3 rather then with Fs ∼ 1. These absorbers can be formed at zf ∼ zobs due to the adiabatic compression within low entropy regions or due to both adiabatic and shock compressions at redshifts zf ≥ 1.5 zobs. For richer absorbers with lgNHI ≥ 13 the radiative cooling is also more important. This explanation is quite consistent with the estimates (44). At z ≥ 3 we have for the fit parameters γb ∼ 0.15 -- 0.17, b0 ∼ 38km/s, what differs from (46). This difference can be partly related to the poor statistics at higher redshifts and to admixture of long lived absorbers formed at redshifts z ≥ 4 when stronger systematic and random variations of UV background are expected. In particular, at z ∼ 3 HeII becomes ionized and the spectrum of UV background and the effective power -- low index of the temperature -- density re- lation are changed (Songaila 1998; Schaye et al. 2000; The- uns et al. 2002a, b). These results demonstrate that, in the framework of model under consideration, the cutoff in the distribution of b(NHI ) can be explained but its interpretation is not so clear. The problem requires further investigation. c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Characteristics of Ly-α forest 13 5 WALLS AT HIGH REDSHIFTS Majority of the rare absorbers with higher Doppler param- eter, b ≥ (80 -- 90)km/s and moderate column density of neutral hydrogen NHI ≤ 1014.5cm−2, can be identified with the embryos of richer structure elements which are seen now as rich walls in the observed galaxy distribution. The num- ber of such absorbers is small and their statistical char- acteristics cannot be reliably determined. However, for 45 such absorbers with b ≥ 90km/s selected in 9 spectra at 1.8 ≤ z ≤ 3.8 we have hlgNHIi ∼ 13.2, and the mean comoving separation of such absorbers (8) is hFsi ∼ 10 , hξi ∼ 1.45, hδi ∼ 6, (48) (49) hDsepi ≈ (60 ± 20)p0.3/Ωmh−1Mpc . Such large separation suggests that these objects will retain their individuality up to small redshifts and, indeed, this separation is quite consistent with the mean wall sep- aration hDwi ≈ (66 ± 13)h−1 Mpc measured for the SDSS EDR at z = 0 (Doroshkevich, Tucker & Allam 2002). How- ever, for smaller b, the number of absorbers rapidly increases and already for 139 absorbers with b ≥ 70km/s in 15 spectra the mean separation decreases to Dsep ∼ (40 ± 20)h−1Mpc. These results show that the walls begin to form already at z ∼ 3 and continue to form up to the present time. At the same time, the continuous distribution of all parameters of absorbers shows that at high redshifts the discrimination of walls and other structure elements is quite arbitrary. The problem deserves further investigation in a wider range of redshifts with a more representative sample of absorbers es- pecially at high redshifts. Perhaps, such walls can be also observed in the galaxy distribution at high redshifts. 6 REDSHIFT DISTRIBUTION OF ABSORBERS In this section we compare the observed redshift distribution of absorbers with theoretical relations (26, 27) describing the expected redshift evolution of the 1D number density of DM pancakes, nabs, due to their formation and merging. As is seen from these relations, the number density depends upon the cosmological model, moments of initial power spectrum and the threshold DM column density, qthr. Therefore, to compare these expectations with observed absorbers distri- bution we have to select samples with q ≥ qthr rather than to use samples with a minimal NHI discussed in Sec. 4. Such samples can be prepared using estimates of q as given by (20) in spite of their unreliability for smaller q (see Sec. 4.2, 6.3). Both relations (26 & 27) are derived for the Gaussian initial perturbations. They depend on the redshift of ab- sorbers only and, so, the impact of absorbers velocities and Doppler parameters on the resulting estimates is minimal. At z ≤ 4 these factors increase the scatter of measured nabs but only at z ≥ 4 -- 4.5 they lead to a significant overlapping of absorbers (Sec. 4.4). The Jeans damping caused by the temperature of background is well described by (27) through the parameter q0. Four samples, namely, Q50 14, with qthr = 0.05, 0.02, 0.01 & 0.001 were selected from 14 high 14, and Q01 14, Q20 14, Q10 14 Demia´nski, Doroshkevich & Turchaninov resolution spectra. These samples contain 1554, 3299, 3998 and 4475 absorbers, respectively. Comparison of qthr used for the sample preparation with estimates of hqthri obtained from fits (26, 27) allows to test the self consistency of our physical model and the choice of parameters (33). 14, Q10 The richest sample Q01 14 includes almost all lines and probably is incomplete. It can be used for comparison with results obtained for other samples. The sample Q50 14 contains richer absorbers and is weakly sensitive to properties of small scale density field described by the parameter q0 in (27). It is used to test the expression (26) and to estimate hqthri for such absorbers. The samples Q20 14 can be used for independent estimates of both hqthri and q0. Of course, these estimates are statistical and averaged over the sample. To test the sample dependence of our results we use the sample Q10 18 with qthr = 0.01 selected from all 18 spectra. As compared with the sample Q10 14 it contains ∼10% additional absorbers with NHI ≥ 1013cm−2. To demonstrate the pos- sible variations of intensity of UV background we consider also the sample W 12 14 containing 2126 weaker absorbers with NHI ≤ 1012cm−2. Here we also use the sample of HST data (Bahcall et al. 1993, 1996; Jannuzi et al. 1998) which con- tains 1000 absorbers at z ≤ 1.5 . This sample is probably incomplete at least at z ≥ 1 . 6.1 Selection of Poissonian subsamples of absorbers The first problem is the selection of Poissonian subsamples of absorbers and estimation of their mean number density at various redshifts. For this purpose, we use the measured red- shift separations of neighboring absorbers, ∆z, what partly attenuates the influence of selection effects inherent in indi- vidual spectra. Instead of separation we use the dimension- less comoving distance, ∆l = H0∆z/H(z) , and absorbers in the interval z − dz < z < z + dz, taken from the sample under investigation, are organized into an 'equivalent single field' by arranging the separations ∆l one after the other along the line of sight. For richer samples, distribution of absorbers separations obtained in this way is similar to the Poissonian one and the mean number density, hni, can be found by comparing the measured PDF with the differential or cumulative Poissonian PDFs: dN/dx = exp(−hni x)/hni , N = exp(−hni x) . (50) (51) To decrease uncertainties and to estimate the actual scatter of these fits we rejected all points before the maxima of differential PDF and only more representative middle part of PDFs with the fraction of points f ≤ fthr were used. The threshold fraction, fthr, was varied between fthr = 0.7 and fthr = 0.95. Mean values and dispersions of the set of measurements with different fthr were taken as the actual value and scatter of nabs. As a rule, the fit (51) of cumulative PDFs gives more stable estimates and smaller error bars than the fit (50) of differential PDFs. Both estimates of nabs are plotted in Figs. 8 & 9. Figure 8. Top panel: redshift distribution of absorbers, nabs(z), found with fits (50) and (51) (red and green points) for the sample W 12 14 . Fit (27) is plotted by green line. Bottom panel: redshift variations of nabs/nf it for the same sample. Fit (52) is plotted by blue line. 6.2 Variations of intensity of the UV background For the sample W 12 14 the redshift distribution of absorbers is plotted in Fig. 8 together with the fit (27) which charac- terizes the expected redshift distribution of absorbers. The weak absorbers in this sample are especially sensitive to vari- ations of intensity of the UV background radiation what is manifested by the strong variations of nabs in comparison with the smooth fitting curve. This effect can be enhanced by the incompleteness of the sample and the nonhomogene- ity of redshift distribution of observed weaker absorbers. At z ≤ 2 the rapid decrease of nabs is caused by the influence of the threshold of the observed column density of neutral hydrogen, NHI ≥ Nthr ≈ 1012cm−2, discussed in Sec. 2.5. The observed variations of hnabsi are well fitted by the function hnabsi/nf it = 1.6 + 0.65 th(cid:16) 2.5 − z 0.16 (cid:17) , (52) also plotted in Fig. 8. Similar variations with smaller ampli- tude are also seen for absorbers in other samples. The column density of neutral hydrogen is proportional to Γ−1 γ and an increase of Γγ shifts weak absorbers below the observational threshold. This means that the relation (52) can describe quite well the redshift dependence of variations of the UV background but their amplitude depends upon the distribution function of NHI . Direct estimates show that the variations of the intensity by about 2 -- 3 times described by (31) are consistent with the observed variation of hnabsi. As was noted in Sec. 2.6 the influence of these variations on the mean characteristics of absorbers can be partly compensated by variations of random factors ΘH & Θδ. Variations of the UV background at z ∼ 3 were al- ready detected by different methods (see, e.g., Songaila 1998; Scott et al. 2000; McDonald & Miralda -- Escude 2001). This c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Characteristics of Ly-α forest 15 Here the formal errors of the fits are given. q0 ≈ (1.1 ± 0.12) × 10−2 . To test the sample dependence of the results we ana- lyzed in the same manner the sample S10 18 selected with qthr= 0.01 from all 18 spectra (4404 absorbers). In this case, the excess of absorbers with NHI ≥ 1013cm−2 and larger separa- tions decrease the measured nabs and the resulting estimates are hqthri ≈ (3.2 ± 0.3) × 10−2, We see that such perturbations lead to moderate variations of both hqthri and q0 . Real precision of both estimates (54) and (55) depends upon several factors, mostly on the limited representativity of the sample and precision of estimates τ0 (12). As is seen from (27), we really evaluate the product qthrτ −2 and q0τ 2 0 , and, so, the real precision of both estimates (54) and (55) is not better than ∼ 20 -- 30% . As was noted in Sec. 2.4, the redshift distribution of absorbers with larger qthr is weakly sensitive to small scale correlations of perturbations and it can be fitted by the ex- pression (26). Indeed, for samples Q20 14 these dis- tributions are well fitted by expression (26) with hqthri ≈ 0.035 & 0.05, respectively, and the amplitude of the fit is only ∼ 1.2 -- 1.3 times larger then what is expected in (26). However, already for the sample Q10 14 the difference between expected and measured amplitudes of the fit (26) increases by up to ∼ 1.75 times. These results show that the actual reliability and pre- cision of our estimates (55) are limited due to limited repre- sentativity of the samples and limited precision of estimates of the amplitude τ0 (12). None the less, they suggest that: 14 and Q50 0 (i) The redshift distribution of absorbers is quite well fit- ted by relations (26) and (27) derived for the Gaussian initial perturbations. (ii) The distribution of weaker absorbers is actually sen- sitive to the small scale cutoff of power spectrum described by the parameter q0 ≈ (0.9 ± 0.3) · 10−2 . (iii) Comparison of estimates (54) with qthr used for the sample selection indicates that, for weak absorbers with q ∼ 10−2, the expression (20) with parameters (33) un- derestimates the DM column density. However, for richer absorbers the divergence decreases to the quite moderate value (∼ 10%). The model as a whole and estimates of q0 can be im- proved with richer sample of absorbers especially at high redshifts z ≥ 3.5 where the representativity of our samples is insufficient. However, as was noted in Songaila (1998) and Schaye et al. (2000), at z ≥ 3 the random variations of the UV background and the scatter of hnabsi increase. 7 SUMMARY AND DISCUSSION. In this paper we continue the analysis initiated in Pa- per I and Paper II based on the statistical description of Zel'dovich pancakes (DD99, DD02). This approach allows to connect the observed characteristics of absorbers with fundamental properties of the initial perturbations without any smoothing or filtering procedures, to reveal and to illus- trate the main tendencies of structure evolution. It demon- strates also the generic origin of absorbers and the Large Figure 9. Top panel: redshift distribution of absorbers, nabs(z), found with fits (50) and (51) (red and green points) for the sample Q10 14. Fit (27) is plotted by the blue line. Bottom panel: Differences between measured and fitted absorber distributions. approach seems to be sufficiently perspective but quantita- tive results are sample dependent and must be tested with more representative and homogeneous samples of weak ab- sorbers. 6.3 Redshift distribution of DM pancakes 14, Q10 14, Q20 14, and Q01 In this Section we consider the redshift distribution of DM pancakes for the samples Q50 14 prepared with qthr(z) ≈ const. for parameters (33). For the sample Q10 14, the function nabs(z) is plotted in Fig. 9 together with the best fit (27) and the random scatter of observed points around the fit. At redshifts 1.8 ≤ z ≤ 3.3 where the repre- sentativity of samples is better the scatter does not exceed ∼ 20% . mean number density of absorbers At redshifts z ∼ 1.5 -- 2 the rapid growth of observed , (53) 2.7 (cid:17)8 hnabsi ∝ (cid:16) 1 + z coincides with the expected one (30) and is naturally ex- plained by the influence of observational threshold NHI ≥ Nthr ≈ 1012cm−2. Such cutoff is not seen for the subpopula- tion of richer absorbers with NHI ≥ 1014cm−2 what verifies its connection with the observational threshold. The proba- ble incompleteness of the sample of weak absorbers at z ≤ 1.5 enhances this rapid growth of hnabsi. The best estimates of the fit parameters (27) for the samples Q20 14, and Q01 hqthri ≈ (3.3 ± 0.3) · 10−2, hqthri ≈ (3.1 ± 0.3) · 10−2 , and for all three samples q0 ≈ (0.9 ± 0.1) · 10−2 . hqthri ≈ (3.1 ± 0.3) · 10−2 , 14 are, respectively, 14, Q10 (54) (55) c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 16 Demia´nski, Doroshkevich & Turchaninov Scale Structure observed in the spatial galaxy distribution at small redshifts. We investigate the new more representative sample of ∼ 4 500 absorbers what allows us to improve the physical model of absorbers introduced in Paper I and Paper II and to obtain reasonable description of physical characteristics of absorbers. The progress achieved demonstrates again the key role of the representativity of observed samples for the construction of the physical model of absorbers and reveals a close connection between conclusions and observational database. However, the representativity of this sample is not suf- ficient for the more detailed study of structure evolution. Further progress can be achieved with richer sample of ob- served absorbers. 7.1 Main results e.g., Songaila 1998; Scott et al. 2000; McDonald & Miralda- Escude 2001). However, the achieved precision of these esti- mates is limited. The variations of background temperature at these redshifts were also detected by Schaye et al. (2000) and McDonald et al. (2001). The redshift variations of the mean number density of weak absorbers discussed in Sec. 6.2 can be naturally related to such variations caused by the reionization of HeII and pos- sible variations of activity of quasars and other sources of UV radiation. They can be enhanced by the incompleteness and limited representativity of our samples and by the non- homogeneous redshift distribution of observed absorbers. However, these variations do not lead to appreciable variations of the mean absorbers characteristics discussed in Sec. 4 what indicates the complex character of absorbers evolution. The problem requires more detailed investigation using both observations and simulations. Main results of our analysis can be summarized as follows: 7.3 Properties of absorbers (i) The approach used in this paper allows to link the ob- served and other physical characteristics of Ly-α absorbers such as the overdensity and entropy of the gaseous compo- nent and the column density of DM component accumulated by absorbers. (ii) The basic observed properties of absorbers are quite successfully described by the statistical model of DM con- fined structure elements (Zel'dovich pancakes). Comparison of independent estimates of the DM characteristics of pan- cakes confirms the self consistency of the physical model for richer absorbers. However, some characteristics of pancakes formation and evolution remain uncertain. (iii) In the framework of this approach, all characteristics of absorbers can be expressed through two functions, ξ & υ, describing the systematic and random variations of absorber properties. They are directly expressed through the observed parameters, z, NHI & b. (iv) The main stages of structure evolution are illustrated by separation of three subpopulations of absorbers with high and low entropy and low overdensity. (v) The absorbers with high Doppler parameter, b ≥ 90 km/s, can be naturally identified with the embryos of wall -- like structure elements observed in the spatial distribution of galaxies at small redshifts. (vi) The strong suppression of the mean number density of absorbers at z ≤ 1.8 can be naturally explained by the influence of the observational threshold NHI ≥ 1012cm−2. (vii) Redshift distribution of weaker absorbers indicates the probable systematic redshift variations of the UV back- ground. They can be caused by the reionization of HeII. (viii) The PDF of the DM column density and the redshift distribution of absorbers are consistent with Gaussian initial perturbations. (ix) We estimate the spectral moment m0 which in turn is linked with the cutoff of the power spectrum caused by the mass of dominant fraction of DM particles and the Jeans damping. 7.2 Variations of intensity of the UV background The intensity of the UV background and its variations at redshifts z ∼ 2 -- 3 were detected by different methods (see, The physical model of absorbers introduced in Sec.2 links the measured z, b and lgNHI with other physical charac- teristics of both gaseous and DM components forming the observed absorbers. Uncertainties in the available estimates of the background temperature and UV radiation and un- known parameters of the model restrict its applications. For- tunately, actions of these factors partly compensate each other, what allows us to obtain reasonable statistical de- scription for majority of absorbers. The self consistency of this approach is confirmed by two independent estimates of the column density of DM pancakes. Numerical simula- tions can clarify action of unknown factors and improve the model. Analysis of mean absorbers characteristics performed in Sec. 4 shows that the sample of observed absorbers is composed of pancakes with various evolutionary histories. We discuss five main factors that determine absorbers evo- lution after formation. They are: the transverse expansion and compression of pancakes, their merging, the radiative heating and cooling of compressed gas, and the disruption of structure elements into a system of high density clouds. The first two factors change the overdensity of DM and gas but do not change the gas entropy. Next two factors change both the gas entropy and overdensity but do not change the DM characteristics. Using the entropy and overdensity we can roughly discriminate three subpopulations of absorbers, illustrate the influence of these factors and correlations be- tween absorbers characteristics. Introduction of the reduced Doppler parameter, υ, over- density, ∆, and entropy, S, allows to discriminate between the systematic and random variations of absorbers prop- erties. The former ones are naturally related to the pro- gressive growth with time of the DM column density of ab- sorbers, q(z), what can be described theoretically. On the other hand the action of random factors cannot be sat- isfactory described by any theoretical model. However, in the framework of our approach, the reduced parameters are found to be strongly correlated (41) and, in fact, the joint action of all random factors is summarized by one random function, υ, directly expressed through the observed param- eters (42). These results alleviate the problem of absorbers description and, perhaps, the modeling of the Ly-α forest c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 based on the simulated DM distribution (Viel et al. 2002). However, estimates of the matter fraction accumulated by absorbers (Sec. 4.4) demonstrate that the model needs to be improved. Estimates of the size of absorbers (Sec. 4.3) suggest a possible fast disruption of richer absorbers into a system of high density clouds, what is the first step of formation of dwarf galaxies. 7.4 Absorbers as elements of the Large Scale Structure of the Universe The close connection of absorbers with the Large Scale Structure (LSS) of the Universe observed in the spatial galaxy distribution was demonstrated in numerical simu- lations and was recently confirmed by direct observations (Penton, Stocke & Shull 2002). Our results indicate the generic link of absorbers and DM Zel'dovich pancakes and demonstrate that the absorbers characteristics are consis- tent with Gaussian initial perturbations and the Harrison -- Zel'dovich power spectrum. The possible identification of some absorbers with embryos of walls observed in the galaxy surveys at small redshifts shows that, perhaps, also a wider set of richer absorbers could be identified with the LSS ele- ments at high redshifts. 7.5 Characteristics of the initial power spectrum The amplitude and the shape of large scale initial power spectrum are approximately established by investigations of relic radiation and the structure of the Universe at z ≪ 1 detected in large redshift surveys such as the SDSS (Dodel- son et al. 2002) and 2dF (Efstathiou et al. 2001). The shape of small scale initial power spectrum can be tested at high redshifts where it is not so strongly distorted by nonlinear evolution (see, e.g., Croft et al. 2002). Recent discussions on the mass of dominant fraction of DM component and the shape of small scale initial power spectrum are focused on the formation of low mass halos in CDM simulations in comparison with observed low mass satellites (see, e.g., Colin, Avila-Reese & Valenzuela 2001; Bode, Ostriker & Turok 2001), and at the simulations of absorbers formation. Thus, Narayanan et al. (2000) esti- mate the low limit of this mass as MDM ≥ 0.75 keV, while Barkana, Haiman & Ostriker (2001) increase this limit up to MDM ≥ 1 -- 1.25 keV. Here we can compare these estimates with direct mea- surements of redshift distribution of absorbers. In fact, our estimates are related to two spectral moments of the ini- tial power spectrum, m−2 and m0. The first of them de- pends mainly upon the large scale power spectrum and was independently estimated from the measured characteristics of observed galaxy distribution at small redshifts. Together with cosmological parameters, Ωm and h, this moment de- fines the coherent length of initial velocity field, lv, (6), used in Sec. 2.3 for discussion of properties of DM distri- bution. The second moment, m0, depends upon the small scale power spectrum and the shape of transfer function. This means that the estimates of the mass of DM particles are model dependent even though our estimates of q0 and spectral moments are based on the analysis of observations. c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Characteristics of Ly-α forest 17 Figure 10. Spectral moments, m−2 and m0, and the parame- ter q0 are plotted vs. MDM for the CDM (red points) and WDM (blue points) models with the damping factors DW (57) and with both damping factors, DW & DJ (58) for xJ = (2 & 0.5) · 10−2, (green points). Black lines show the observational restrictions (59). As was shown in DD99 and DD02, for the CDM and WDM models with the Harrison -- Zel'dovich asymptotic of power spectrum, p(k) ∝ k, and transfer functions, T (k), given in Bardeen et al. (1986) the coherent length of ini- tial density perturbations, lρ, and the parameter q0 can be written as follows: q0 = 5 m2 −2 m0 , m−2 = Z ∞ 0 x = k k0 , 0.01 lρ = q0lv ≈ 0.34h−1Mpc(cid:16) q0 dx xT 2(x) ≈ 0.023, m0 = Z ∞ k0√m−2 k0 = Ωmh2/Mpc, lv = 1 0 0.2 Ωmh(cid:17) , (56) dx x3T 2(x) , = 6.6 Ωmh2 Mpc , where k is the comoving wave number. For WDM particles the dimensionless damping scale, Rf , and the damping factor, DW , are , Rf = 0.2(cid:18) Ωmh2keV MDM (cid:19)4/3 DW = exp[−xRf − (xRf )2] , (Bardeen et al. 1986). The Jeans wave number, kJ , and the damping factor, DJ , can be taken as k−1 J ≈ 0.7bbg(1 + z)H −1(z), DJ ≈ (1 + x2 xJ = k0/kJ ≈ 2 · 10−2r Ωmh2 bbg 0.13 r 4 J x2)−1 , 16km/s 1 + z (58) (57) , (see, e.g., Matarrese & Mohayaee 2002). Variations of the spectral moments and q0 with the mass of DM particle obtained by integration of the power spectrum with these damping factors are plotted in Fig. 18 Demia´nski, Doroshkevich & Turchaninov 10 for the CDM and WDM transfer functions and for xJ = 2 · 10−2 & 0.5 · 10−2. As is seen from this Fig., the low limit q0 ≥ 0.07√xJ caused by the Jeans damping (58), restricts the range of masses of DM particles detectable with this approach to MDM ≤ (3 -- 5)keV. Our 1σ estimates q0 ≈ (0.6 − 1.2) · 10−2, m0 ≈ 0.15 − 0.5, MDM ≈ (1.5 − 5)keV , are close to those of Narayanan et al. (2000) and Barkana, Haiman & Ostriker (2001). They are based on absorbers with the redshifts z ≤ 3.5 where the scatter of measured nabs is minimal and, so, they weakly depend upon both the Doppler parameter and velocities of absorbers which essen- tially distort nabs at higher redshifts (Sec. 4.4). (59) There are plenty of possible candidates for the WDM particles with the mass ∼ 1 keV. They are, for example, the sterile neutrinos, majorons and even shadow particles. De- tailed review of possible candidates can be found in Sommer- Larsen & Dolgov (2001) and Dolgov (2002). However, our estimates use the spectral moments and therefore do not forbid the existence of multicomponent heavy DM particles with both thermal and non-thermal (Lin et al. 2001) distri- butions with the same spectral moments m−2 and m0. 7.6 Reheating of the Universe Recent observations of high redshift quasars with z ≥ 5 (Djorgovski et al. 2001; Becker et al. 2001; Pentericci et al. 2001; Fan et al. 2001) demonstrate clear evidences in fa- vor of the reionization of the Universe at redshifts z ∼ 6 when the volume averaged fraction of neutral hydrogen is found to be fH ≥ 10−3 and the photoionization rate Γγ ∼ (0.2 − 0.8) · 10−13s−1 . These results are consistent with those expected at the end of the reionization epoch which probably takes place at z ∼ 6. Extrapolation of the mean separation of absorbers discussed in Sec. 4.4 is consis- tent with these conclusions. These results can be compared with expectations of the Zel'dovich approximation (DD02) for q0 = 10−2. The poten- tial of this approach is limited since it cannot describe the nonlinear stages of structure formation and, so, it cannot substitute the high resolution numerical simulations. How- ever, it describes quite well many observed and simulated statistical characteristics of the structure such as the red- shift distribution of absorbers and evolution of their DM column density. This approach does not depend on the box size, number of points and other limitations of numerical simulations and it successfully augments them. This approach allows to estimate the fractions of DM component accumulated by high density clouds, fcl, fila- ments, ff , and pancakes, fpan, at different redshifts. These functions plotted in Fig. 11 for q0 = 10−2 show that at z ∼ 6 only ∼ 3.5% of the matter is condensed within the high den- sity clouds which can be associated with luminous objects. This value can increase up to ∼ 5 -- 6% with more correct description of the clouds collapse. At the same redshifts, ∼ 27% and ∼ 12% of the matter can be already accumulated by pancakes and filaments, respectively. These expectations are smaller than estimates of hfabsi obtained in Sec. 4.4 . The Zel'dovich approximation allows also to estimate Figure 11. Top panel: expected redshift variations of the DM fraction accumulated by high density clouds, filaments and pan- cakes (red, green and blue lines), respectively, at q0 = 10−2. Bot- tom panel: the expected mass function of the DM clouds at z = 6 and q0 = 10−2. the mass function of all structure elements (DD02) at dif- ferent redshifts. For q0 = 10−2 and at z ∼ 6, this function is plotted in Fig. 11. For these q0 and z, the mean DM mass of structure elements is expected to be ∼ 1012M⊙ and the main mass is concentrated within clouds with Mcl ∼ 0.1hMcli. As is seen from Fig. 11, majority of the clouds have mass be- tween 10−3hMcli and 10 hMcli. The formation of low mass clouds with Mcl ≤ 109M⊙ is suppressed due to strong cor- relation of the initial density and velocity fields at scales ≤ lρ ∼ 0.3h−1 Mpc (56). However, the numerous low mass satellites of large central galaxies can be formed in the course of disruption of massive collapsed clouds at the stage of their compression into thin pancake -- like objects (Doroshkevich 1980; Vishniac 1983). The minimal mass of such satellites was estimated in Barkana, Haiman & Ostriker (2001). This means that the investigation of absorbers observed at high redshifts should be supplemented by the study of properties of dwarf isolated galaxies and discrimination be- tween such galaxies and dwarf satellites of more massive galaxies. It seems to be a perspective way to discriminate between models with the Jeans damping (58) and the damp- ing caused by the mass of WDM particles, (57), and, in par- ticular, between models with one and several types of DM particles. Acknowledgments AGD is grateful to Dr. S. Cristiani and Dr. T.S.Kim for the permission to use the unpublished observational data. This paper was supported in part by Denmark's Grund- forskningsfond through its support for an establishment of Theoretical Astrophysics Center and by the Polish State c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Characteristics of Ly-α forest 19 Kim T.S., Cristiani S., & D'Odorico S., 2002, A&A, in press, astro-ph/0201204 Kim T.S., Carswell R.F., Cristiani S., D'Odorico S. & Gial- longo E., 2002, MNRAS, in press, astro-ph/0205237 Kirkman D., & Tytler D., 1997, ApJ., 484, 672. Kulkarni V.P. et al., 1996, MNRAS, 279, 197 Lanzetta K.M., Bowen D.V., Tytler D., & Webb J.K., 1995, ApJ., 442, 538. Le Brune V., Bergeron J., & Boisse P., 1996, A&A, 306, 691 Lin W.B., Huang D.H., Zhang X., Brandenberger R., 2001, Phys.Rev.Lett. 86, 954 Lu L., Sargent W.L.W., Womble D.S., Takada-Hidai M., 1996, ApJ., 472, 509 Matarrese S. & Mohayaee R., 2002, MNRAS, 329, 37 McDonald P. et al., 2000, ApJ., 543, 1 McDonald P. & Miralda-Escude J., 2001, ApJ., 549, L11 McDonald P. et al., 2001, ApJ., 562, 52 McGill C., 1990, MNRAS, 242, 544 Narayanan V. et al., 2000, ApJ., 543, L103 Oort J.H., 1981, Astr.Astrophys., 94, 359. Oort J.H., 1984, Astr.Astrophys., 139, 211. Pentericci L. et al., 2001, astro-ph/0112075 Penton S.V., Stocke J.T., & Shull J.M, 2002, ApJ., 565, 720 Rauch M., 1998, Ann.Rev.Astr.Astrophys., 36, 267 Rodriguez-Pascual P.M., de la Fuente A., 1995, ApJ., 448, 575. Sargent, W.L.W., Young, P.J., Boksenberg, A. & Tytler, D., 1980, Ap.J.Suppl, 42, 41 Schaye J., Theuns T., Leonard A. & Efstathiou G., 1999, MN- RAS, 310, 57 Schaye J.et al., 2000, MNRAS, 318, 817 Scott J., Bechtold J., Dobrzicki A. & Kulkarni V.P., 2000, ApJ- Supl, 130, 67 Shandarin S., Zel'dovich Ya.B., 1989, Rev.Mod.Phys., 61, 185 Shandarin S. et al., 1995, Phys.Rev.Let., 75, 7 Shectman S.A. et al., 1996, ApJ., 470, 172. Sommer-Larsen J. & Dolgov A., 2001, ApJ., 551, 608 Songaila A., 1998, AJ., 115, 2184 Stoughton C. et al., 2001, AJ., 123, 485 Theuns T., Leonard A., Schaye J., Efstathiou G., 1999, MNRAS, 303, L58 Theuns T., Schaye J., Zaroubi S., Kim T.-S., Tzanavaris P., & Carswell R., 2002a, ApJ, 567, L103 Theuns T., Zaroubi S., Kim T.-S., Tzanavaris P., & Carswell R., 2002b, MNRAS, 332, 367 Tytler D., 1995, in Meylan J., ed., QSO Absorption Lines, p. 289. Viel M., Matarrese S., Mo J.H., Theuns T., Haehnelt M.G., 2002, MNRAS, 329, 848 Vishniac E.T., 1983, ApJ., 274, 152 Weinberg D.H., Burles S., Croft R.A.C., et al., 1998, in "Evolution of Large Scale Structure: From Recombination to Garching", eds. A.J. Banday, R.K. Sheth, L.N. Da Costa, p. 346 Zaldarriaga M., Scoccimarro R., Hui L., 2001, astro-ph/0111230 Zel'dovich Ya.B., 1970, Astrophysica, 5, 20 Zeng W., Davidsen A.F., & Kriss G.A., 1998, AJ, 115, 391 Zhang Yu., Meiksin A., Anninos P., Norman M.L., 1998, ApJ., 495, 63 Committee for Scientific Research grant Nr. 2-P03D-014-17. AGD also wishes to acknowledge support from the Center of Cosmo-Particle Physics, Moscow. Furthermore, we wish to thank the anonymous referee for valuable discussion and many useful comments. REFERENCES Bahcall J.N. et al., 1993, ApJS., 87, 1. Bahcall J.N. et al., 1996, ApJ., 457, 19. Bardeen J.M., Bond J.R., Kaiser N., Szalay A., 1986, ApJ., 304, 15 Barkana R., Haiman Z., Ostriker J.P., 2001, ApJ., 558, 482 Becker R.H. et al. 2001, AJ, 122, 2850 Bergeron J., Cristiani S., & Shaver P.A., 1992, A&A, 257, 417 Bi H., & Davidsen A.F., 1997, ApJ., 479, 523 Black J.H., 1981, MNRAS, 197, 553 Bode P., Ostriker J.P. & Turok N, 2001, ApJ., 556, 93 Bond R., Wadsley J.W., 1997, in "Structure and Evolution of the IGM from QSO Absorption Line Systems", Eds. P.Petitjean, S.Charlot, Editions Frontiers, Paris, p. 143 Choudhury R., Padmanabhan T. & Srianand R., 2001, MNRAS, 322, 561 Choudhury R., Srianand R. & Padmanabhan T., 2001, ApJ, 559, 29 Coles P., Melott A., & Shandarin S., 1993, MNRAS, 289, 37 Colin P., Avila-Reese V., & Valenzuela O., 2001, ApJ., 542, 622 Cristiani S., D'Odorico V., 2000, AJ, 120, 1648 Croft R.A.C., Weinberg D.H., Katz N., & Hernquist L., 1998, ApJ., 495, 44 Croft R.A.C. et al., 2001, ApJ., 557, 67 Croft R.A.C. et al., 2002, astro-ph/00132324 Dav´e R., Hernquist L., Katz N., Weinberg D.H., 1999, ApJ., 511, 521, Demia´nski M. & Doroshkevich A., 1999, MNRAS., 306, 779, (DD99). Demia´nski M., Doroshkevich A.G., Muller V., & Turchani- nov V.I., 2000, MNRAS, 318, 665 Demia´nski M., Doroshkevich A.G., & Turchaninov V.I., 2001, MNRAS, 318, 1177, (Paper I) Demia´nski M., Doroshkevich A.G., & Turchaninov V.I., 2001, MNRAS, 318, 1189, (Paper II) Demia´nski M. & Doroshkevich A., 2002, astro-ph/0206282, (DD02) Djorgovski S.G., Castro S., Stern D., Mahabal A.A., 2001, ApJ., 560, L5 Dodelson S. et al., 2002, ApJ, 572, 140 Dolgov A., 2002, Phys.Rep., in press, hep=ph/0202122 Doroshkevich, A.G., 1980, SvA., 24, 152 Doroshkevich, A.G., Tucker, D.L. & Allam S., 2002, astro- ph/0206301 Efstathiou G. et al., 2001, astro-ph/0109152 Fan X. et al., 2001, astro-ph/0111184, ApJ., in press Gnedin N.Y. & Hui L., 1996, ApJ, 472, L73 Gnedin N.Y. & Hamilton A.J.S., 2002, MNRAS, 334, 107 Giallongo E., Cristiani S., Fontana A., & Trevese D., 1993, ApJ., 416, 137 Hu E.M., Tae-Sun K., Cowie L., & Songaila A., 1995, AJ, 110, 1526 Hui L., Gnedin N.Y., 1997, MNRAS, 292, 27 Hui L., Gnedin N.Y. & Zhang Yu., 1997, ApJ, 486, 599 Hui L., Rutledge R.E., 1999, ApJ, 517, 541 Ikeuchi S., Ostriker J.P, 1986, ApJ., 301, 522. Jannuzi B.T. et al. ApJS. 1998, 118, 1 Jacobsen P., et al. 1994, Nature, 370, 35 Khare P. et al., Bechtold J., 1997, MNRAS, 285, 167. c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
0810.4538
1
0810
2008-10-24T20:00:05
Report of IAU Commission 30 on Radial Velocities (2006-2009)
[ "astro-ph" ]
Brief summaries are given on the following subjects: Radial velocities and exoplanets (Toward Earth-mass planets; Retired A stars and their planets; Current status and prospects); Toward higher radial velocity precision; Radial velocities and asteroseismology; Radial velocities in Galactic and extragalactic clusters; Radial velocities for field giants; Galactic structure -- Large surveys (The Geneva-Copenhagen Survey; Sloan Digital Sky Survey; RAVE); Working groups (WG on radial velocity standards; WG on stellar radial velocity bibliography; WG on the catalogue of orbital elements of spectroscopic binaries [SB9]).
astro-ph
astro-ph
Transactions IAU, Volume XXVIIA Reports on Astronomy 2006-2009 Karel A. van der Hucht, ed. c(cid:13) 2009 International Astronomical Union DOI: 00.0000/X000000000000000X COMMISSION 30 RADIAL VELOCITIES PRESIDENT VICE-PRESIDENT PAST PRESIDENT ORGANIZING COMMITTEE VITESSES RADIALES Stephane Udry Guillermo Torres Birgitta Nordstrom Francis C. Fekel, Kenneth C. Freeman, Elena V. Glushkova, Geoffrey W. Marcy, Birgitta Nordstrom, Robert D. Mathieu, Dimitri Pourbaix, Catherine Turon, Tomaz Zwitter COMMISSION 30 WORKING GROUPS Div. IX / Commission 30 WG Div. IX / Commission 30 WG Div. IX / Commission 30 WG Radial-Velocity Standard Stars Stellar Radial Velocity Bibliography Catalogue of Orbital Elements of Spectroscopic Binaries TRIENNIAL REPORT 2006 -- 2009 1. Introduction This three-year period has seen considerable activity in the Commission, with a wide range of applications of radial velocities as well as a significant push toward higher precision. The latter has been driven in large part by the exciting research on extrasolar planets. This field is now on the verge of detecting Earth-mass bodies around nearby stars, as demonstrated by recent work summarized below, and radial velocities continue to play a central role. This is not to say that classical applications of RVs have lagged behind. On the con- trary, this triennium has seen the release of several very large data sets of stellar radial velocities (Galactic and extragalactic) that are sure to have a significant impact on a number of fields for years to come. The era of mass-producing radial velocities has ar- rived. Examples include the Geneva-Copenhagen Survey, the Sloan Digital Sky Survey, and RAVE, and are described below. Due to circumstances beyond our control, the report of Commission 30 for the previous (2003 -- 2006) triennium did not appear in the printed version of the Transactions of the IAU, although it did appear in the electronic version. For progress during the previous period, the reader is therefore encouraged to consult the latter, which is available from the Commission web site. 1 8 0 0 2 t c O 4 2 ] h p - o r t s a [ 1 v 8 3 5 4 . 0 1 8 0 : v i X r a 2 DIVISION IX / COMMISSION 30 2. Radial velocities and exoplanets By G. Torres and J. Johnson 2.1. Toward Earth-mass planets Detections of Jupiter-mass exoplanets by the radial-velocity method relying on measure- ments with precisions of a few m s−1 are now quite routine. This technique has provided by far the majority of the more than 300 planet discoveries to date. The persistence of astronomers and the increasing precision of their instruments has led to larger and larger numbers of multi-planet systems being found. One example is the interesting case of µ Arae (Pepe et al. 2007), with four planets, one of which is as small as 10.5 MEarth. The host star also presents the signature of p-mode oscillations seen clearly in the radial velocities. The record-holder for the most planets is the star 55 Cnc, which is orbited by no less than five planets (Fischer et al. 2008), of which the smallest has a minimum mass of 10.8 MEarth. Exciting discoveries during this period made possible by the high precision and stability of the HARPS instrument on the ESO 3.6-m telescope at La Silla (Chile) include the system Gls 581 (at only 6.3 pc), attended by at least three planets. In addition to the previously known Neptune-mass body orbiting the star with a period of 5.3 days, two other low mass planets were found by Udry et al. (2007) with minimum masses of only 5 MEarth (period 12.9 days) and 7.7 MEarth, the latter being near the outer edge of the habitable zone of the M3V parent star (period 83.6 days). The three-planet orbital solution for this case has an rms residual of only 1.2 m s−1. Another system with three low-mass planets was announced by Mayor et al. (2008) around the nearby (13 pc) metal-poor K2V star HD 40307. The planets weighed in at 4.2, 6.9, and 9.2 MEarth, and the three-planet Keplerian orbital fit gave impressive residuals of just 0.85 m s−1. HARPS has demonstrated that this sort of velocity precision is achievable for "quiet" stars that present a low level of "jitter" in their radial velocities due to astrophysical phenomena such as p-mode oscillations, granulation, or chromospheric activity. Indications are that Neptune-mass or smaller planets are more common around solar-type (F -- K) stars than previously thought (see, e.g., Mayor & Udry 2008). 2.2. Retired A stars and their planets Most Doppler searches for planets have concentrated on main sequence stars of spec- tral types F or later, because the velocity precision for earlier type stars is seriously compromised by line broadening induced by rapid rotation, as well as the overall fewer number of spectral lines available. This difficulty in studying higher-mass stars intro- duces a bias in our understanding of planets, but it can be overcome by looking at such stars after they have left the main sequence. This is precisely the approach of an on- going project to investigate the relationship between stellar mass and planet formation by using the HIRES instrument on the Keck 10-m telescope to search for planets in a sample of 240 intermediate-mass subgiants (1.3 < M∗/M⊙ < 2.2). Subgiants have lower surface temperatures and rotational velocities than their main-sequence progen- itors, making them ideal proxies for A- and F-type stars in Doppler studies. From a smaller sample of subgiants observed previously by Johnson et al. (2007a) at Lick Ob- servatory for 4 years with a typical velocity precision of 4 m s−1, a strong correlation was detected between stellar mass and planet occurrence, with a detection rate of 9% within 2.5 AU among the high-mass sample, compared to 4.5% for Sun-like stars and less than 2% for M dwarfs. A paucity of planets within 1 AU of stars with masses greater than 1.5 M⊙ was found, indicating that stellar mass also plays a key role in planet migration (Johnson et al. 2007b, Johnson et al. 2008). The goal of the expanded Keck survey (with RADIAL VELOCITIES 3 an increased velocity precision of about 2 m s−1) is to map out the relationships between stellar mass and exoplanet properties in greater detail by examining the distribution of planetary minimum masses, eccentricities, semimajor axes, and the rate of multiplicity around evolved A stars. If the 9% occurrence rate is confirmed, some 20 -- 30 new planets should be found in the sample orbiting some of the most massive stars so far examined by the Doppler technique. 2.3. Current status and prospects In May of 2008 NASA convened the Exoplanet Forum 2008, a meeting of experts from the US and other countries in eight different observational techniques related to exoplanet re- search. The purpose was to discuss paths forward for exploring and characterizing planets around other stars, and to provide specific suggestions for space missions, technology de- velopment, and observing programmes that could fulfill the recommendations of a previ- ously held meeting of the Exoplanet Task Force (http://www.nsf.gov/mps/ast/exoptf.jsp). The reports resulting from these meetings are intended to provide input for considera- tion by various advisory committees in the US, and in particular by the Astronomy and Astrophysics Decadal Survey that is currently underway. Radial velocities was one of the eight techniques considered by the Exoplanet Forum 2008. The corresponding chapter of the report, available at http://exep.jpl.nasa.gov/exep exfCommunityReport.cfm , summarized the progress in the field over the last few years, which is illustrated by a veloc- ity precision of 1 m s−1 or slightly better achieved so far, led by the Swiss team using the HARPS instrument on the ESO 3.6-m telescope, and the California-Carnegie team using the HIRES instrument on the Keck 10-m telescope. The factors currently limiting the pre- cision were discussed briefly and have been described in detail by Pepe & Lovis (2007). They include various sources of astrophysical noise (stellar oscillations, granulation, mag- netic cycles, collectively known as "stellar jitter"), guiding, the illumination of the spec- trograph, and the wavelength reference. Good progress has been made in each of these areas. For example, it appears that jitter can be substantially reduced through longer exposures or binning, to the level of perhaps 10 cm s−1 or less. A new thorium-argon line list was developed by Lovis & Pepe (2007) that significantly improves the velocity pre- cision when using this source as the wavelength reference. Further improvements in the velocity precision perhaps reaching a few cm s−1 appear possible using a dense spectrum of lines generated by a femtosecond-pulsed laser ("laser comb"), described in more detail below. The next few years will tell whether this promise can be realized in practice. The report of the Exoplanet Forum also described recent progress in techniques to mea- sure precise velocities in the near infrared (see, e.g., Ramsey et al. 2008), which are now approaching the 10 m s−1 level in initial tests. Longer wavelengths potentially provide a significant advantage for the Doppler detection of very small (even Earth-mass) planets, since these objects produce a larger signal when orbiting less massive stars, which emit most of their flux in the near infrared. In addition to velocity precision, the report pointed out what is currently considered by the community to be the greatest challenge for making progress in the detection of exoplanets by the Doppler technique: the limited access to telescope time. This has a direct impact not only on the size of the samples of solar-type stars that can be studied, but also severely restricts the number of late-type (faint) stars that can be targeted to search for Earth-mass planets. The need for exposure times longer than dictated by Poisson statistics to reduce stellar jitter, as mentioned above, is a further strain 4 DIVISION IX / COMMISSION 30 on the limited resources currently available on telescopes equipped with high-precision spectrographs. 3. Toward higher radial velocity precision By G. Torres During this period agreement has been reached for the construction of an improved copy of the very successful HARPS spectrograph, currently in operation at the ESO 3.6-m telescope at La Silla, for the northern hemisphere (HARPS-NEF). This is a high- resolution (R ≈ 120,000) fibre-fed optical spectrograph with broad wavelength coverage (3780 -- 6910 A) designed for high radial velocity precision. HARPS-NEF is a collabora- tion between the New Earths Facility (NEF) scientists of the Harvard Origins of Life Initiative and the HARPS team at the Geneva Observatory. It is expected to be the workhorse for follow-up of transiting planet candidates for NASA's Kepler mission, and should be operational perhaps in late 2010. HARPS-NEF is a cross-dispersed echelle spectrograph that will benefit not only from updates and improvements over the original HARPS instrument, but in addition it will be installed on a larger telescope aperture in the northern hemisphere (the 4.2-m William Herschel Telescope on La Palma, Canary Islands). It is designed for ultra-high stability (10 -- 20 cm s−1), and like HARPS it will be placed in a vacuum chamber with careful temperature control. One of the key factors that determine the precision of the RVs is the wavelength ref- erence. Existing technologies in the optical (such as the Th-Ar technique and iodine gas absorption cell) have already reached sub-m s−1 precision in some cases, but further improvements are needed if the Doppler method is to reach cm s−1, as is needed to de- tect terrestrial-mass planets. A new technology that has emerged in the last few years and that holds great promise for providing a very stable reference is that of laser "fre- quency combs". As the name suggests, a frequency comb generated from mode-locked femtosecond-pulsed lasers provides a spectrum of very narrow emission lines with a con- stant frequency separation given by the pulse repetition frequency, typically 1 GHz for this application. This frequency can be synchronized with an extremely precise reference such as an atomic clock. For example, using the generally available Global Position- ing System (GPS), the frequencies of comb lines have long-term fractional stability and accuracy of better than 10−12. This is more than enough to measure velocity varia- tions at a photon-limited precision level of 1 cm s−1 in astronomical objects (see, e.g., Murphy et al. 2007). This direct link with GPS as the reference allows the comparison of measurements not only between different instruments, but potentially also over long periods of time. To provide lines with separations that are well matched to the resolv- ing powers of commonly used echelle spectrographs, a recent improvement incorporates a Fabry-P´erot filtering cavity that increases the comb line spacing to ∼40 GHz over a range greater than 1000 A (Li et al. 2008). Prototypes using a titanium-doped sapphire solid-state laser have been built that provide a reference centreed around 8500 A. In practice, of course, Doppler measurements are also affected by other instrumental prob- lems, so that the value of this new technology for highly precise RV measurements is still to be demonstrated. Tests have been initiated during this triennium. Plans call also for the installation of a laser comb on the HARPS-NEF spectrograph described above. Ap- plications of this technique are not limited to stars. For example, a direct measurement of the expansion of the Universe could be made by observing in real time the evolu- tion of the cosmological redshift of distant objects such as quasars. Such a measurement RADIAL VELOCITIES 5 would require a precision in determining Doppler drifts of ∼1 cm s−1 per year (see, e.g., Steinmetz et al. 2008), which a laser comb can in principle deliver. 4. Radial velocities and asteroseismology By G. Torres The significant increase in the precision of velocity measurements over the past few years, driven by exoplanet searches, has enabled important studies of the internal consti- tution of stars through the technique of asteroseismology. A number of spectrographs now reach the precision needed for this type of investigation. During this period Bedding et al. (2006) observed the metal-poor subgiant star ν Ind with the UCLES instrument on the 3.9-m Anglo-Australian Telescope, and with the CORALIE spectrograph on the 1.2-m Swiss telescope at ESO. The precision of those measurements ranged from 5.9 to 9.5 m s−1, and allowed the authors to place constraints on the stellar parameters confirming that the star has a low mass and an old age. This was the first application of asteroseis- mology to a metal-poor star. α Cen A was observed by Bazot et al. (2007) with the HARPS spectrometer on the ESO 3.6-m telescope, and 34 p modes were identified in the acoustic oscillation spectrum of the star. Individual observations had errors well under 1 m s−1. A similar study by Mosser et al. (2008a) was conducted on Procyon (α CMi) using the SOPHIE spectrograph on the 1.9-m telescope at the Haute-Provence Observa- tory, yielding a precision of about 2 m s−1. The HARPS instrument was used again by Mosser et al. (2008b) to study the old Galactic disk, low-metallicity star HD 203608. A total of 15 oscillation modes were identified, and the age of the star was determined to be 7.25 ± 0.07 Gyr. 5. Radial velocities in Galactic and extragalactic clusters By E. V. Glushkova, H. Levato, and G. Torres Searches for spectroscopic binaries in southern open clusters have continued during this period (e.g., Gonz´alez & Levato 2006). These authors have reported results for the open cluster Blanco 1. Forty four stars previously mentioned in the literature as cluster candidates, plus an additional 25 stars in a wider region around the cluster were observed repeatedly during 5 years. Six new spectroscopic binaries have been detected and their orbits determined. All of them are single-lined spectroscopic systems with periods ranging from 1.9 to 1572 days. When considering also all suspected binaries, the spectroscopic binary frequency in this cluster amounts to 34%. Additional velocities were measured in this cluster by Mermilliod et al. (2008a), who obtained a rather similar binary frequency. Results from long term radial velocity studies based on the CORAVEL spectrome- ters have been presented during this period for the open clusters NGC 6192, 6208, and 6268 (Clari´a et al. 2006), as well as for NGC 2112, 2204, 2243, 2420, 2506, and 2682 (Mermilliod & Mayor 2007). These studies were complemented with photometric obser- vations in a variety of systems, and included membership determination and binary stud- ies. A number of new spectroscopic binaries were discovered, and their orbital elements were determined. Other individual cluster studies in the Milky Way, which we merely reference here without giving the details due to space limitations, include: IC 2361 (Platais et al. 2007), NGC 2489 (Piatti et al. 2007), α Per (Mermilliod et al. 2008b), the five distant open clus- ters Ru 4, Ru 7, Be 25, Be 73 and Be 75 (Carraro et al. 2007), Tombaugh 2 (Frinchaboy et al. 2008), 6 DIVISION IX / COMMISSION 30 the Orion Nebula cluster (Fur´esz et al. 2008), the most massive Milky way open clus- ter Westerlund 1 (Mengel & Tacconi-Garman 2008), the Galactic centre star cluster (Trippe et al. 2008), and the globular clusters M4 (Sommariva et al. 2008) and ω Cen (Da Costa & Matthew 2008). This triennium saw the publication of the final results of the 20-year efforts of J.-C. Mermilliod and colleagues to measure the radial velocities of giant stars in open clusters for a variety of studies related to their kinematics, membership, and photometric and spectroscopic properties. A catalogue of spectroscopic orbits for 156 binaries based on more than 4000 individual velocities was published by Mermilliod et al. (2007), based on measurements from CORAVEL and the CfA Digital Speedometers. Orbital periods range from 41 days to more than 40 years, and eccentricities are as high as e = 0.81. Another 133 spectroscopic binaries were discovered but do not have sufficient observations and/or time coverage to determine orbital elements. This material provides a dramatic increase in the body of homogeneous orbital data available for red-giant spectroscopic binaries in open clusters, and should form the basis for a comprehensive discussion of membership, kinematics, and stellar and tidal evolution in the parent clusters. A companion catalogue (Mermilliod et al. 2008c) reports mean radial velocities for 1309 red giants in clusters based on 10,517 individual measurements, and mean radial velocities for 166 open clus- ters among which 57 are new. This information, combined with recent absolute proper motions, will permit a number of investigations of the galactic distribution and space motions of a large sample of open clusters. Frinchaboy & Majewski (2008) reported on a survey of the chemical and dynamical properties of the Milky Way disk as traced by open star clusters. They used medium- resolution spectroscopy (R ≈ 15,000) with the Hydra multi-object spectrographs on the Cerro Tololo Inter-American Observatory 4-m and WIYN 3.5-m telescopes to derive mod- erately high-precision RVs (σ < 3 km s−1) for 3436 stars in the fields of 71 open clusters within 3 kpc of the Sun. Along with the work described in the preceding paragraph, these represent the largest samples of clusters assembled thus far having uniformly determined, high-precision radial velocities. A good deal of activity focused on kinematic analyses of globular cluster (GC) systems in other galaxies. Lee et al. (2008a) measured radial velocities for 748 GC candidates in M31, and Lee et al. (2008b) obtained radial velocities of 111 objects in the field of M60. Konstantopoulos et al. (2008) obtained new spectroscopic observations of the stellar clus- ter population of region B in the prototype starburst galaxy M82. Schuberth et al. (2006) presented the first dynamical study of the GC system of NGC 4636 based on radial ve- locities for 174 clusters. Bridges et al. (2006) measured radial velocities of 38 GCs in the Virgo elliptical galaxy M60, and Bridges et al. (2007) obtained new velocities for 62 GCs in M104. An interesting problem was discussed by Abt(2008), pointing to a possible bias in the RVs of many B-type stars. The author looked at 10 open clusters younger than about 30 million years with sufficient numbers of measured radial velocities, many of them being measured with CORAVEL, and found that in each case, the main-sequence B0 -- B3 stars have larger velocities than earlier- or later-type stars. 6. Radial velocities for field giants By G. Torres A programme to measure precise radial velocities for 179 giant stars has been on- going at the Lick Observatory, with individual errors of 5 -- 8 m s−1 per measurement RADIAL VELOCITIES 7 (Hekker et al. 2006). This study presented a list of 34 stable K giants (with RV stan- dard deviations under 10 m s−1) suitable to serve as reference stars for NASA's Space Interferometry Mission. A follow-up paper (Hekker et al. 2008) reported that 80% of the stars monitored show velocity variations at a level greater than 20 m s−1, of which 43 exhibit significant periodicities. One of the goals was to investigate possible mechanisms that cause these variations. A complex correlation was found between the amplitude of the changes and the surface gravity of the star, in which part of the variation is periodic and uncorrelated with log g, and another component is random and does correlate with surface gravity. Massarotti et al. (2008) reported radial velocities made with the CfA Digital Speedome- ters for a sample of 761 giant stars, selected from the Hipparcos Catalogue to lie within 100 pc. Rotational velocities and other spectroscopic parameters were determined as well. Orbital elements were presented for 35 single-lined spectroscopic binaries and 12 double- lined binaries. These systems were used to investigate stellar rotation in field giants to look for evidence of excess rotation that could be attributed to planets that were engulfed as the parent stars expanded. 7. Galactic structure -- Large surveys By B. Nordstrom and G. Torres 7.1. The Geneva-Copenhagen Survey During the previous 3-year period one of the mayor surveys completed and published is the Geneva-Copenhagen Survey of the Solar Neighbourhood (Nordstrom et al. 2004). Unfortunately the full description of this project and the important new science results that came out of it did not make it into the printed version of the Transactions of the IAU for 2003 -- 2006, so we summarize and update that information here for its significant impact for the study of Galactic structure. This survey provided accurate, multi-epoch radial velocities for a magnitude-complete, all-sky sample of 14,000 F and G dwarfs down to a brightness limit of V = 8.5, and is volume complete to about 40 pc. The catalogue in- cludes new mean radial velocities for 13,464 stars with typical mean errors of 0.25 km s−1, based on 63,000 individual observations made mostly with the CORAVEL photoelectric cross-correlation spectrometers covering both hemispheres. Studies of this rich data set have found evidence for dynamical substructures that are probably due to dynamical perturbations induced by spiral arms and perhaps the Galactic bar. These "dynamical streams" (Famaey et al. 2005) contain stars of different ages and metallicities which do not seem to have a common origin. These features, which dominate the observed U ,V ,W diagrams, make the conventional two-Gaussian decomposition of nearby stars into thin and thick disk members a highly dubious procedure. An analysis by Helmi et al. (2006) suggests that tidal debris from merged satellite galaxies may be found even in the solar neighbourhood. A new release of this large catalogue with updated calibrations as well as new age and metallicity determinations was published during the present triennium by Holmberg et al. (2007), and is available from the CDS at http://cdsweb.u-strasbg.fr/cgi-bin/qcat?J/A+A/475/51 . A follow-up paper and catalogue are expected to be available shortly, containing new kinematic data (U V W velocities) resulting from a re-analysis using the revised Hipparcos parallaxes (van Leeuwen 2007), and online updates. 8 DIVISION IX / COMMISSION 30 7.2. Sloan Digital Sky Survey This period saw the sixth data release of the Sloan Digital Sky Survey (Adelman-McCarthy et al. 2008), which now covers an area of 9583 square degrees on the sky. This release includes nearly 1.1 million spectra of galaxies, quasars, and stars with sufficient signal to be usable, along with redshift determinations, as well as effective temperature, surface gravity, and metallicity determinations for many stars. The spectra cover the wavelength region 3800 -- 9200 A at a resolving power ranging from 1850 to 2200. Velocity precisions range from about 9 km s−1 for A and F stars to about 5 km s−1 for K stars. The zero point of the velocities is in the process of being calibrated using spectra from the ELODIE spec- trograph. These data are a valuable resource for a variety of investigations related to Galactic structure and the evolution and history of the Milky Way. 7.3. RAVE The second data release of the Radial Velocity Experiment (RAVE) was published dur- ing this triennium (Zwitter et al. 2008). This is an ambitious spectroscopic survey to measure radial velocities as well as stellar atmosphere parameters (effective temperature, metallicity, surface gravity, rotational velocity) of up to one million stars using the 6dF multi-object spectrograph on the 1.2-m UK Schmidt telescope of the Anglo-Australian Observatory. The RAVE programme started in 2003, obtaining medium resolution spec- tra (median R = 7500) in the Ca II triplet region (8410 -- 8795 A) for southern hemisphere stars drawn from the Tycho-2 and SuperCOSMOS catalogues, in the magnitude range 9 < I < 12. Following the first data release, the current release doubles the sample of published radial velocities, now reaching 51,829 measurements for 49,327 individual stars observed between 2003 and 2005. Comparison with external data sets indicates that the new data collected since April 2004 show a standard deviation of 1.3 km s−1, about twice as good as for the first data release. For the first time, this data release contains values of stellar parameters from 22,407 spectra of 21,121 individual stars. The data release includes proper motions from the STARNET 2.0, Tycho-2, and UCAC2 catalogues, and photometric measurements from Tycho-2, USNO-B, DENIS, and 2MASS. The data can be accessed via the RAVE web site at http://www.rave-survey.org . Scientific uses of these data include the identification and study of the current structure of the Galaxy and of remnants of its formation, recent accretion events, as well as the discovery of individual peculiar objects and spectroscopic binary stars. For example, kine- matic information derived from the RAVE data set has been used by Smith et al. (2007) to constrain the Galactic escape velocity at the solar radius to Vesc = 536+58 −44 km s−1 (90% confidence). 8. Working groups Below are the progress reports of the three active working groups of Commission 30. Their efforts are focused on providing a service to the astronomical community at large through the compilation of a variety of information related to radial velocities. By S. Udry 8.1. WG on radial velocity standard stars Large radial-velocity surveys are being conducted to search for extrasolar planets around different types of stars, including A to M dwarfs, and G -- K giants (e.g., Udry & Santos 2007). RADIAL VELOCITIES 9 Although not aiming at establishing a set of radial-velocity standard stars, the non- variable stars in these programmes, followed over a long period of time, provide ideal candidates for our list of standards. They will moreover broaden the domain of stellar properties covered (brightness and spectral type). At this point, the results of most of those programmes are still not publicly available and we must still wait a bit in order to fine-tune and enlarge the list presently available at http://obswww.unige.ch/∼udry/std/std.html . In addition to the by-product aspect of planet search programmes, a targeted observa- tional effort, dedicated to the definition of a large sample of RV standards for GAIA, is being pursued with several instruments (CORALIE, SOPHIE, etc). It will provide in a few years a list of several thousand suitable standards spread over the entire sky (Crifo et al. 2007). For all of the efforts above, work remains to be done to combine the data from the different instruments into a common RV system, for example through the observation of minor planets in the solar system (Zwitter et al. 2007). This has still to be done for most of the planet search programmes, but is already included in the GAIA effort. By H. Levato 8.2. WG on stellar radial velocity bibliography During the 2006 -- 2009 triennium, the WG searched for the papers with measurements of radial velocities of stars in 33 journals. As of December 2007 113,658 entries have been catalogued. We expect to finish 2008 with more than 150,000. It is worth mentioning that at the end of 1996 there were 23,358 entries recorded, so that in 10 years the number of entries in the catalogue has expanded by a factor of five. During the triennium we have improved the search engine to search by different parameters. In the main body of the catalogue we have included information about the technical characteristics of the instrumentation used for radial velocity measurements, and comments about the nature of the objects. The catalogue can be accessed at http://www.casleo.gov.ar/catalogue/catalogue.html . 8.3. WG on the catalogue of orbital elements of spectroscopic binaries (SB9) By D. Pourbaix In Manchester, a WG was set up to work on the implementation of the 9th catalogue of orbits of spectroscopic binaries (SB9), superseding the 8th release of Batten et al. (1989) (SB8). SB9 exists in electronic format only. The web site http://sb9.astro.ulb.ac.be was officially released during the summer of 2001. This site is directly accessible from the Commission 26 web site, from BDB (in Besan¸con), and from the CDS, among others. Since the last report, substantial progress has been accomplished, in particular in the way complex systems can be uploaded together with their radial velocities. That is the case, for instance, for triple stars with the light time effect accounted for and systems with a pulsating primary. At the time of this writing SB9 contains 2802 systems (SB8 had 1469) and 3340 orbits 10 DIVISION IX / COMMISSION 30 (1469 in SB8). A total of 563 papers were added since August 2000, although most of them come from outside the WG. Many papers with orbits still await uploading into the catalogue. According to ADS, the release paper (Pourbaix et al. 2004) has been cited a total of 58 times since 2005. This is twice as many as the old Batten et al. catalogue over the same period. Even though this work has been very well received by the community and a number of tools have been designed and implemented to make the job of entering new orbits easier (input file checker, plot generator, etc.), the WG still suffers from a serious lack of manpower. Few colleagues outside the WG spontaneously send their orbits (but they are usually pleased to send their data when we ask for them). Any help (from authors, journal editors, and others) is therefore very welcome. Uploading an orbit into SB9 also involves checking for typos. In this way we have found several mistakes in published solutions, which we have corrected. Sending orbits to SB9 prior to publication (e.g., at the proof stage) would therefore be a way to prevent some mistakes from making it into the literature. Guillermo Torres Vice-President of the Commission References Abt, H. 2008, PASP, 120, 715 Adelman-McCarthy, J. K., Agueros, M. A., Allam, S. S. et al. 2008, ApJS, 175, 297 Batten, A. H., Fletcher, J. M., & MacCarthy, D. G. 1989, Eighth catalogue of the orbital elements of spectroscopic binary systems, Publ. Dom. Astr. Obs., 17, 1 Bazot, M., Bouchy, F., Kjeldsen, J., Charpinet, S., Laymand, M., & Vauclair, S. 2007, A&A, 470, 295 Bedding, T. R., Butler, R. P., Carrier, F. et al. 2006, ApJ, 647, 558 Bridges, T., Gebhardt, K., Sharples, R. et al. 2006, MNRAS, 373, 157 Bridges, T., Rhode, K. L., Zepf, S. E., & Freeman, K. C. 2007, ApJ, 658, 980 Carraro, G., Geisler, D., Villanova, S. et al. 2007, A&A, 476, 217 Claria, J. J., Mermilliod, J.-C., Piatti, A. E., & Parisi, M. C. 2006, A&A, 453, 91 Crifo, F., Jasniewica, G., Soubiran, C. et al. 2007, in Towards a new set of radial velocity standards for GAIA, eds. J. Bouvier, A. Chalabaev, & C. Charbonnel, Proceedings of the Annual meeting of the French Society of Astronomy and Astrophysics, Grenoble (France), p. 459 Da Costa, G. S., & Matthew, C. G. 2008, AJ, 136, 506 Famaey, B., Jorissen, A., Luri, X. et al. 2005, A&A, 430, 165 Fischer, D. A., Marcy, G. W., Butler, R. P. et al. 2008, ApJ, 675, 790 Frinchaboy, P. M., Marino, A. F., & Villanova, S. et al. 2008, MNRAS (in press), arXiv:0809.2559 Frinchaboy, P. M., & Majewski, S. R. 2008, AJ, 136, 118 Fur´esz, G., Hartmann, L. W., Megeath, S. T. et al. 2008, ApJ, 676, 1109 Gonz´alez, J. F., & Levato, H. 2006, RMxAA, 26, 171 Hekker, S., Reffert, S., Quirrenbach, A., Mitchell, D. S., Fischer, D. A., Marcy, G. W., & Butler, R. P. 2006, A&A, 454, 943 Hekker, S., Snellen, I. A. G., Aerts, C., Quirrenbach, A., Reffert, S., & Mitchell, D. S. 2008, A&A, 480, 215 Helmi, A., Navarro, J. F., Nordstrom, B. et al. 2006, MNRAS 365, 1309 Holmberg, J., Nordstrom, B., & Andersen, J. 2007, A&A, 475, 519 Johnson, J. A. et al. 2007a, ApJ, 670, 833 Johnson, J. A., Marcy, G. W., Fischer, D. A. et al. 2008, ApJ, 675, 784 Johnson, J. A., Butler, R. P., Marcy, G. W., Fischer, D. A., Vogt, S. S., Wright, J. T., & Peek, K. M. G. 2007b, ApJ, 665, 785 RADIAL VELOCITIES 11 Karchenko, N. V., Scholz, R.-D., Piskunov, A. E. et al. 2007, AN, 328, 889 Konstantopoulos, I. S., Bastian, N., Smith, L. J. et al. 2008, ApJ, 674, 846 Lee, M. G., Hwang, Ho S., Kim, S. Ch. et al. 2008a. ApJ, 674, 886 Lee, M. G., Hwang, Ho S., Park, H. S. et al. 2008b, ApJ, 674, 857 Li, Ch.-H., Benedick, A. J., Fendel, P. et al. 2008, Nature, 452, 610 Lovis, C., & Pepe, F. 2007, A&A, 468, 1115 Massarotti, A., Latham, D. W., Stefanik, R. P., & Fogel, J. 2008, AJ, 135, 209 Mayor, M., & Udry, S. 2008, Phys. Scr., 130 (in press) Mayor, M., Udry, S., Lovis, C. et al. 2008, A&A (in press), arXiv:0806.4587 Mengel, S., & Tacconi-Garman, L. E. 2008, in Young massive star clusters - Initial conditions and environments, Granada, Spain (in press), arXiv:0803.4471 Mermilliod, J.-C., Platais, I., James, D. J., Grenon, M., & Cargile, P. A. 2008a, A&A, 485, 95 Mermilliod, J.-C., & Mayor, M. 2007, A&A, 470, 919 Mermilliod, J.-C., Andersen, J., Latham, D. W., & Mayor, M. 2007, A&A, 473, 829 Mermilliod, J.-C., Queloz, D., & Mayor, M. 2008b, A&A, 488, 409 Mermilliod, J.-C., Mayor, M., & Udry, S. 2008c, A&A, 485, 303 Mosser, B., Bouchy, F., Marti´c, M. et al. 2008a, A&A, 478, 197 Mosser, B., Deheuvels, S., Michel, E. et al. 2008b, A&A, 488, 635 Murphy, M. T., Udem, Th., Holzwarth, R. et al. 2007, MNRAS, 380, 839 Nordstrom, B., Mayor, M., Andersen, J. et al. 2004, A&A, 418, 989 Pepe, F., Correia, A. C. M., Mayor, M. et al. 2007, A&A, 462, 769 Pepe, F. A., & Lovis, C. 2007, in Physics of Planetary Systems, Nobel Symposium 135, in press Piatti, A., Clari´a, J. J., Mermilliod, J.-C., Parisi, M. C., & Ahumada, A. V. 2007, MNRAS, 377, 1737 Platais, I., Melo, C., Mermilliod, J.-C. et al. 2007, A&A, 461, 509 Pourbaix, D., Tokovinin, A. A., Batten, A. H. et al. 2004, A&A, 424, 727 Ramsey, L. W., Barnes, J., Redman, S. L., Jones, H. R. A., Wolszczan, A., Bongiorno, S., Engel, L., & Jenkins, J. 2008, PASP, 120, 887 Schuberth, Y., Richtler, T., Dirsch, B. et al. 2006, A&A, 459, 391 Smith, M. C., Ruchti, G. R., Helmi, A. et al. 2007, MNRAS, 379, 755 Sommariva, V., Piotto, G., Rejkuba, M. et al. 2008, A&A (in press), arXiv:0810.1897 Steinmetz, T., Wilken, T., Araujo-Hauck, C. et al. 2008, Science, 321, 1335 Trippe, S., Gillessen, S., Gerhard, O. E. et al. 2008, A&A (in press), arXiv:0810.1040 Udry, S., & Santos, N. C. 2007, ARA&A, 45, 397 Udry, S., Bonfils, X., Delfosse, X. et al. 2007, A&A, 469, L43 van Leeuwen, F. 2007, A&A, 474, 653 Zwitter, T., Mignard, F., & Crifo, F. 2007, A&A, 462, 795 Zwitter, T., Siebert, A., Munari, U. et al. 2008, AJ, 136, 421
astro-ph/9701141
1
9701
1997-01-20T17:19:19
The ages and distances of globular clusters with the luminosity function method: the case of M5 and M55
[ "astro-ph" ]
We present new age and distance determinations for the Galactic Globular Clusters M55 and M5, using the luminosity function method (Jimenez & Padoan 1996, Padoan & Jimenez 1997). We find an age of $11.8 \pm 1.5$ Gyr for M55 and $11.1 \pm 0.7$ Gyr for M5. This confirms previous results (Jimenez et al. 1996, Sandquist et al. 1996) and allows to conclude that the oldest stars in the Universe are not older than 14 Gyr. We also find $m-M=14.13 \pm 0.11$ for M55, and $m-M=14.49 \pm 0.06$ for M5. These values agree with the ones obtained using the tip of the red giant branch (Jimenez et al. 1996) and the sub-dwarf fitting method (Sandquist et al. 1996).
astro-ph
astro-ph
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000) Printed 14 August 2018 (MN LATEX style file v1.4) The ages and distances of globular clusters with the luminosity function method: the case of M5 and M55. Raul Jimenez,1 and Paolo Padoan2 1Royal Observatory, Blackford Hill, Edinburgh EH9-3HJ, UK 2Theoretical Astrophysics Center, Juliane Maries Vej 30, DK-2100 Copenhagen 0, Denmark 14 August 2018 ABSTRACT We present new age and distance determinations for the Galactic Globular Clusters M55 and M5, using the luminosity function method (Jimenez & Padoan 1996, Padoan & Jimenez 1997). We find an age of 11.8 ± 1.5 Gyr for M55 and 11.1 ± 0.7 Gyr for M5. This confirms previous results (Jimenez et al. 1996, Sandquist et al. 1996) and allows to conclude that the oldest stars in the Universe are not older than 14 Gyr. We also find m− M = 14.13± 0.11 for M55, and m− M = 14.49± 0.06 for M5. These values agree with the ones obtained using the tip of the red giant branch (Jimenez et al. 1996) and the sub-dwarf fitting method (Sandquist et al. 1996). Key words: globular clusters: general -- globular clusters: individual (M5, M55) 1 INTRODUCTION An accurate determination of the ages and distances of Globular Clusters (GCs) is an im- portant constraint for the age of the Universe, and for the theory of galaxy formation. In particular it is important to compute very accurate relative ages to understand if there is a spread in ages among the Galactic GCs or not. The use of the stellar luminosity function (LF) to compute ages of GCs was first proposed by Paczynski (1984). Later on, Jimenez & Padoan (1996) and Padoan & Jimenez (1996) developed a method to determine the age and the distance of a GC simultaneously, using c(cid:13) 0000 RAS 2 R. Jimenez & P. Padoan the LF. The method is described in detail in Padoan & Jimenez (1996), where it is concluded, on the basis of artificial data, that an uncertainty of about 0.6 Gyr in the age and 0.06 mag in the distance modulus can be achieved, if the number of stars, in 1 mag-wide luminosity bins, is known with an uncertainty of 3%. In this paper we use recent observations of the Galactic Globular Clusters M5 (Sandquist et al. 1996) and M55 (Desidera & Ortolani, private communication) to apply the LF method and compute accurate ages and distance module. These two clusters are very adequate since M55 is a metal-poor one and M5 has intermediate metallicity, so we can investigate the spread in ages (if any) in the formation of the GC system. In this letter we apply for the first time the LF method to real data and we show that the method is much more superior to traditional methods (isochrone fitting to the main sequence turn off point, ∆V method, or any other methods that involve the fitting of the main sequence turn off). The method is superior because it allows to determine the age and the distance simultaneously and independently and because the errors in computing the age and distance are straightforward to calculate. Furthermore, it gives age determinations with sufficient accuracy to make cosmological predictions. This first application of our LF method to real data shows that our previous theoretical predictions were correct. 2 THE DATA In order to apply our LF method it is necessary to obtain the complete LF of the glob- ular cluster from almost the tip of the red giant branch (RGB), down to the upper main sequence. The number of observed stars should be very large, in order to keep statistical errors sufficiently low. Recently, two LFs that fulfill these requirements have been obtained by Sandquist et al. (1996) for M5, and by Desidera & Ortolani (private communication) for M55. M5 is a massive globular cluster, with an average metallicity of [F e/H] = −1.17 ± 0.01, according to Sneden et al. (1992), and [F e/H] = −1.4, according to Zinn & West (1984). Since it is a high altitude cluster (b = 46.8o), it is not seriously affected by contamination and interstellar reddening. For the metallicity of this cluster we used [F e/H] = −1.3, and adopt Y = 0.24. The completeness of the LF is discussed in detail in Sandquist et al. (1996). M55 is among the poor-metal clusters. The main advantage of M55 is that it is not very concentrated and therefore it is possible to resolve its core into stars. Again due to its c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 The ages and distances of globular clusters with the luminosity function method: the case of M5 and M55. 3 high galactic latitude (b = −23o), interstellar reddening and contamination are negligible. We adopt a metallicity of [F e/H] = −1.9 (Briley et al. 1993) and Y = 0.24. The LF was provided to us by Desidera & Ortolani (private communication.), who have performed extensive tests with artificial stars, in order to compute the completeness of the sample. In this letter we use the data in the filter band V for M55 and in I for M5. This allows us to illustrate the robustness of the method in two very different bands. The other colours were used in both cases to remove the horizontal and asymptotic giant branch (AGB) stars. In the case of the horizontal branch it is quite easy to distinguish them from the RGB. Even if this is not the case for the AGB stars, the number of stars used in the LF method is so large in that bin that a small mistake in distinguishing RGB and AGB stars does not contribute significantly to the errors in the final age and distance determination. In producing LFs to study both GCs we have used [α/F e] = 0.4 and compute this effect in our solar-scaled tracks using the approach by Chieffi, Straniero & Salaris (1991). This value of alpha enhancement is well justified by spectroscopic observations of giants in GCs (Minniti et al 1996), and this enhancement is valid for a metallicity range ([F e/H] = −1.5 to −2.0). 3 THE LUMINOSITY FUNCTION METHOD To determine the age and the distance of a GC, two independent constraints are needed from the LF, that is to say the number of stars in two different bins. One more bin is needed for the normalization, and a fourth bin is useful to estimate the completeness of the data, but it is not used in this work, since the completeness was previously estimated performing experiments with artificial stars (nevertheless, we have checked that the fourth bin gives the same completeness estimation as the artificial star tests). We use a bin positioned at the RGB to normalize the LF, and two bins around the sub-giant region to constrain age and distance modulus. As discussed in Padoan & Jimenez (1996), the precise position of the two bins at the sub-giant region is extremely important. In fact, our LF method is based on the careful optimization of those two bins. The result of the optimization process is a contour plot of the quantity R(t, m − M), on the plane (t, m − M), where: R2(t, m − M) = [n2,th(t) − n2,obs(t, m − M)]2 c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 4 R. Jimenez & P. Padoan +[n3,th(t) − n3,obs(t, m − M)]2 (1) where t is the age, m− M the distance modulus, ni,obs the normalized ratio Ni,obs/N1,obs, Ni,obs the number of stars in the ith observational bin and ni,th the corresponding theoretical ratio. The values of ni,th are only functions of the age of the GC, since the shape of the theoretical LF depends only on the age (for a given chemical composition), while the values of ni,obs are also functions of the distance modulus, because the observational LF depends on the distance modulus, if the bins are defined in absolute magnitudes. If the set of bins is not optimal, the contour plot of R shows only open lines, which define a relation between age and distance modulus. Once the set of bins is optimized, the contour plot shows also closed lines, that define both the age and the distance modulus of the GC at the same time. If the lines start to become closed only for R ≤ 0.1, the degeneracy age-distance modulus is broken only if stellar counts are available with uncertainty smaller than 10%. This is the reason why the method requires LF with a large number of stars and excellent photometry. The optimization process, applied to M55 and M5 gave the following set of bins: MV,optimal,M 55 = (tGyr − 9.0) × 0.05 +[4.01, 3.01, 2.01, 0.01]mag MI,optimal,M 5 = (tGyr − 8.0) × 0.05 +[2.87, 2.07, 1.27, −3.13]mag The optimal bins shift by 0.05 mag/Gyr, as determined in our previous work (Padoan & Jimenez 1996). This is an essential point in the attempt to obtain the contour plot of R. Note that for M5 we could use two bins narrower than 1.0 mag, in order to improve the sensitivity of the method, thanks to the very large number of stars in the bins and to the good quality of the photometry. Our results for M55 is shown in Fig. 1, and for M5 in Fig. 2. In both cases, closed contour lines are obtained for R ≤ 0.1, and the uncertainty in age and m− M, obtained at any given level of R, is comparable for the two clusters, and similar to what we previously predicted with artificial LFs (Padoan & Jimenez, 1996) Given the photometric uncertainty and the statistical uncertainty in the stellar counts (1/√N, where N is the number of stars), we estimate a global uncertainty of 6% for the c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 The ages and distances of globular clusters with the luminosity function method: the case of M5 and M55. 5 M5 M55 age m-M 14.49 ± 0.06 11.1 ± 0.7 11.8 ± 1.5 14.13 ± 0.11 Table 1. The table gives the values for the age and distance modulus for M5 and M55. These values have been determined simultaneously using the luminosity function method described in the text. case of M55, and of 4% for the case of M5. Entering Fig. 1 and Fig. 2 with R = 0.06 and R = 0.04 respectively, one gets the results listed in Table 1. 4 DISCUSSION AND CONCLUSIONS The results listed in Table 1 show that the age and the distance modulus of M55 are in good agreement with previous determinations by Mandushev et al. (1996) and Alcaino et al. (1992). In the case of M5 the results agree with Sanquist et al. conclusions. They estimate in fact an age of 13.5 ± 1 Gyr, for [F e/H] = −1.17, and they state that the age would be 11.5 Gyr, for [F e/H] = −1.4. We use [F e/H] = −1.3, and get an age of 11.1 ± 0.7 Gyr. They also estimate m − M = 14.50 ± 0.07 mag for [F e/H] = −1.17, and m − M = 14.41 ± 0.07 mag, for [F e/H] = −1.4, using the sub-dwarf fitting of the main sequence. We get m− M = 14.49 ± 0.06 mag, for [F e/H] = −1.3. Note that in Padoan & Jimenez (1996) we estimated a variation of 0.02 mag in m − M, for a shift of 0.1 in metallicity; so we would predict m − M = 14.47 ± 0.06 mag, for [F e/H] = −1.4. The distance modulus measured with the LF method is therefore in excellent agreement with the distance modulus determined with the sub-dwarf fitting. The uncertainty of our estimates is very small (±0.06 mag), and no assumption on the age is required. The LF method is superior to the main sequence turn-off (MSTO) method (Chaboyer, Demarque & Sarajedini 1996), to determine the absolute age of globular clusters, because it is not affected by the three largest sources of theoretical uncertainty affecting the MSTO method, that is to say the determination of the value of the mixing length parameter, the morphology of the MSTO and the color-Teff calibration (see Jimenez et al. 1996 for a detailed discussion of the main uncertainties in the MSTO method). Furthermore, the MSTO method needs to know the distance in order to determine the age, and it is unable to break this degeneracy. The absolute ages, determined in this work for M55 and M5, seem to indicate that the oldest GCs are not older than 14 Gyr. c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 6 R. Jimenez & P. Padoan Figure 1. The figure shows the contour plots of R(t, m − M ) (see text) in determining simultaneously the distance modulus and age of M55. Notice that the contours closed around a central value, showing that the method works quite well in breaking the age-distance degeneracy. The LF method is a very powerful tool to investigate relative ages, since most uncertain- ties of stellar evolution theories are in that case avoided. From the comparison of the ages of M5 and M55 we can conclude that the age of the two GCs is not significantly different. We conclude by remarking that most methods to determine age and distance module of GCs share two common problems: some degree of dependence of age on distance modulus (or vice-versa), and a somewhat fuzzy procedure to estimate the uncertainty of the final result. Our LF method, instead, gives constraints for both age and distance modulus independently, and estimates both most probable values and uncertainties in a straightforward way. On the basis of the present work, we think that very high quality data for GCs, together with the LF method, may shed new light on the problems of the age of the oldest stars in the Universe and the formation of the Galaxy. ACKNOWLEDGEMENTS We are grateful to S. Desidera % S. Ortolani for providing us with unpublished data of M55. c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 The ages and distances of globular clusters with the luminosity function method: the case of M5 and M55. 7 Figure 2. The same as before but for M5. The estimated uncertainty of 4% is marked with dashed lines. REFERENCES Alcaino, G., Liller, W., Alvarado, F., Wenderoth, E., 1992, AJ, 104, 190 Briley, M.M., Smith, G.H., Hesser, J.E., Bell, R.A., 1993, ApJ, 306,142 Chaboyer, B., Demarque, P., & Sarajedini, A. 1996, ApJ, 459, 558 Chieffi, A., Straniero, O., Salaris, M., 1991, in ASP Conf. Ser. 13, The Formation and Evolution of Star Clusters, ed. K. Janes (San Francisco: ASP), 219 Jimenez, R., Thejll, P., Jørgensen, U.G., MacDonald, J., Pagel, B., 1996, MNRAS, 282, 926 Jimenez, R., Padoan, P., 1996, ApJ, 463, L17 Mandushev, G.I., Fahlman, G.G., Richer, H.B., Thompson, I.B., 1996, AJ, 112, 1536 Minniti, D., Petterson, R.C., Geisler, D., Claria, J.J., 1996, ApJ, 470, 953 Paczynski, B., 1984, ApJ, 284, 670 Padoan, P. & Jimenez, R. 1997, ApJ, 475, 580 Sandquist, E.L., Bolte, M., Stetson, P.B., Hesser, J.E., 1996, ApJ, 470, 910 Sneden, C., Kraft, R.P., Prosser, C.F., Langer, G.E., 1992, AJ, 104, 2121 Zinn, R. & West, M. J. 1984, ApJS, 55, 45 c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
astro-ph/9408068
1
9408
1994-08-18T21:08:01
The Damping and Excitation of Galactic Warps by Dynamical Friction
[ "astro-ph" ]
We investigate the dynamical interaction of galactic warps with the surrounding dark matter halo, using analytic perturbation theory. A precessing warp induces a density wake in the collisionless dark matter, which acts back on the original warp, transferring energy and angular momentum between the warp and halo (dynamical friction). In most cases dynamical friction damps the warp, but in unusual circumstances (such as a halo that rotates in the same direction as the precession of the warp, or a warp in the equatorial plane of an axisymmetric prolate halo) friction can excite a warp. The damping/excitation time is usually short compared to the Hubble time for realistic systems. Thus most warps cannot be primordial; they must be maintained by some ongoing excitation mechanism.
astro-ph
astro-ph
The Damping and Excitation of Galactic Warps by Dynamical Friction Robert W. Nelson and Scott Tremaine Canadian Institute for Theoretical Astrophysics ABSTRACT We investigate the dynamical interaction of galactic warps with the surrounding dark matter halo, using analytic perturbation theory. A precessing warp induces a density wake in the collisionless dark matter, which acts back on the original warp, transferring energy and angular momentum between the warp and halo (dynamical friction). In most cases dynamical friction damps the warp, but in unusual circumstances (such as a halo that rotates in the same direction as the precession of the warp, or a warp in the equatorial plane of an axisymmetric prolate halo) friction can excite a warp. The damping/excitation time is usually short compared to the Hubble time for realistic systems. Thus most warps cannot be primordial; they must be maintained by some ongoing excitation mechanism. 1 Introduction The origin and maintenance of the `integral sign' warps commonly seen in the outer parts of disk galaxies is a longstanding puzzle in galactic dynamics (for a recent review see Binney 1992). Many if not most disk galaxieseven isolated onesappear to be warped, which implies that warps are either long-lived or repeatedly excited (Sanchez-Saavedra et al. 1990, Bosma 1991). Following an initial suggestion by Lynden-Bell (1965), Hunter & Toomre (1969; hereafter HT) explored the possibility that warps are discrete modes of oscillation of a thin, self-gravitating disk. They found, however, that isolated disks with realistic density pro les actually support a continuum of bending modes; consequently, any packet of bending waves should disperse. Dekel & Shlosman (1983) and Toomre (1983) suggested a solution to this problem, which was analyzed in detail by Sparke & Casertano (1988; hereafter SC): realistic disks can support discrete bending modes if they are embedded in the static potential of a attened dark matter halothese modes are distorted versions of the neutral tilt mode of an isolated disk. Thus the gravitational in uence of galaxy halos may play an important role in the maintenance of galactic warps; conversely, by studying the dynamics of warps in galaxies we may learn about the shape and dynamics of the unseen halos. Unfortunately, galaxy halos are not simply static potentials: the collisionless halo material whatever its naturemust respond to the gravitational eld of an embedded disk. A warped, precessing disk induces a density wake in the dark matter halo and the gravitational eld from the wake will then act back on the disk; in this manner energy and angular momentum are transferred { 2 { between disk and halo in a manner analogous to the Landau damping of waves in a collisionless plasma or dynamical friction on a point mass travelling through a stellar system. In general, then, any bending mode of a galaxy disk is not precisely neutral; the question is whether the dynamical coupling between the halo and the disk is strong enough to dampor excite, as we shall seebending modes of a galaxy on timescales less than the age of the galaxy itself . 1 This issue was rst investigated by Bertin & Mark (1980), who approximated the warp as a tightly-wound (WKB) bending wave, embedded in an in nite, homogeneous, and isotropic `halo' having a Maxwell-Boltzmann velocity distribution. Their analysis predicted that the amplitude of the bending wave would grow inside the corotation radiusand on a short timescale. Unfortunately, this interesting conclusion is suspect, because Bertin and Mark neglected the unperturbed halo's contribution to the vertical restoring force (see Section 3.1). In this paper we examine the role of a responsive halo in damping or exciting bending disturbances in galactic disks. We extend Bertin and Mark's seminal analysis to include warps and halos with more realistic shapes, as well as halos with anisotropic velocity distribution and net rotation. In particular, if the orbits of halo particleswe shall refer to these as `stars' although their nature is unknownare quasiperiodic, energy and angular momentum exchange between disk and halo occurs via stars whose orbits are nearly resonant with the frequency of the time-varying disk potential (Tremaine & Weinberg 1984). In this situation, the direction of energy and angular momentum ow can be from disk to halo or vice versa, so that dynamical friction with the halo may either damp or excite warps, depending on the details of the halo structure. Our analysis of this phenomenon is more elaborate than any in the literature so far, but still has a number of limitations. We do not consider the self-gravity of the halo response. We calculate the unperturbed orbits of stars in the disk plus halo system by assuming that the potential is spherical, and assume that the halo is axisymmetric for the purpose of computing the precession rate and energy of the bending disturbance; whereas cosmological N-body simulations of the collapse of protogalactic density peaks typically result in triaxial halos. We use linear perturbation theory, whose validity is suspect since warps often have substantial amplitudes. Despite these shortcomings, we believe that the principal results of our analysis are robust. In particular, we nd that dynamical friction from the halo can either damp or excite bending disturbances in a disk, although damping is the most common outcome for plausible disk/halo con gurations. The characteristic growth or decay time is generally less than a Hubble time, implying that dynamical friction plays a ma jor role in determining the present properties of galaxy warps. In Section 2 we review the dynamics of bending disturbances in a thin disk embedded in a static halo potential. In Section 3 we relax this assumption and consider the coupling of the 1 `One key question about this kind of proposal is how long the central disk can thereafter remain tilted despite dynamical friction against the doubtless abby halo' (Toomre 1983). { 3 { bending perturbation to the dynamics of the halo stars. In particular, we give general formulae for the damping/excitation rate of bending disturbances from dynamical friction with the halo. In Section 4 we present numerical calculations of damping/excitation times for reasonably realistic halo and disk potentials and warp shapes. Finally, in Section 5 we discuss the implications of our results. 2 Warp Dynamics We review the behavior of linear bending waves of a zero-thickness, axisymmetric, self-gravitating disk embedded in a static halo potential, delaying until Section 3 consideration of the coupling of bending waves to a responsive halo. Many, but not all, of the results we describe are presented in HT, Sparke (1984), and SC. 2.1 The equation of motion We work in cylindrical coordinates (R; ; z ) and take the unperturbed disk to lie in the z = 0 plane, so that position in the disk is described by the vector R  (R; ). We consider a razor-thin axisymmetric disk of surface density (R), rotating at angular speed (R) > 0. The equation of motion in the inertial frame for small vertical displacements, Z (R; ; t), is (HT) 2 2   D Z @ @  + Z = F + F + F ; (1) h sg ext 2 Dt @ t @  where the three terms on the right side are the vertical components of the force per unit mass arising from the gravitational eld of the static halo (F ) and the disk (F ), and from any other h sg sources (F ). We assume that the halo is axisymmetric and that its density is smooth near the ext disk, so that the halo potential can be expanded in a Taylor series U (R; z ) = U (R; 0) +  (R)z ; where  (R) = (R; z = 0) (2) h h 2 h h 2 @ z 1 2 2 2 2 @ U h is the square of the frequency of small vertical oscillations in the eld of the halo. Thus F (R; t) =  (R)Z (R; t): (3) h h 2 It is sometimes useful to decompose the radial force R into its contributions from the halo 2 and the disk, 2 2 2 (R) = (R) + (R); (4) h d where 2 2 h 1 @U G d dR (R ) Z 0 0 (R) = (R; z = 0); (R) = : (5) h d 0 R @R R dR jR R j { 4 { The vertical restoring force from a slightly distorted disk is given by HT, Z 0 0 [Z (R; t) Z (R ; t)] 0 F (R; t) = G dR (R ) : (6) sg jR R j 0 3 We may write the disturbance Z as a superposition of eigenfunctions of the form Z (R; t) = Re h(R)e ; (7) n o i(m!t) where we assume m  0 without loss of generality. If ! is real and m 6= 0, the pattern speed  !=m is the angular speed of the rotating frame in which the disturbance is stationary. p When external forces are absent (F = 0), the dynamical equation (1) for a single ext eigenfunction reduces to n o Z 1 [! m (R)]  (R) h(R) = G dR R (R )[h(R)H (R; R ) h(R )I (R; R )] (8) h m 2 2 0 0 0 0 0 0 0 where 2 Z Z 2 2 0 0 d d cos m H (R; R ) = ; I (R; R ) = ; (9) 2 0 0 3=2 2 0 0 3=2 2 2 m 0 0 (R + R 2RR cos ) (R + R 2RR cos ) these are Laplace coecients and can be evaluated in terms of elliptic integrals. The component of the torque on the disk parallel to the z = 0 plane may be written  =  e +  e , where x x y y Z  + i = dR(R)[F (y ix) z (F iF )]; (10) x y z y x here F = (F ; F ; F ) is the force per unit mass on the disk at R, composed of a radial component x y z 2 2 R and a vertical component  Z + F . Thus sg h Z  + i = i dR(R)Re [(  )Z (R; t) + F ]: (11) x y sg h i 2 2 Since the torque from the disk on itself must be zero we have (Kuijken 1991) Z i 2 0 = dR(R)Re [ Z (R; t) + F ]; (12) d sg (a result which can also be derived by manipulating eqs. 5 and 6) and  + i = i dR(R)Re (  )Z (R; t): (13) x y h h Z i 2 2 2 0 The functions H and I diverge as R ! R but the integral in equation (8) is nevertheless well-de ned. m { 5 { These equations degenerate to 0 = 0 when m 6= 1 but when m = 1 we may use equations (1) and (7) to reduce equation (12) to the form Z 1 0 2 2 2 2 dRR (R)[ 2 + (  )]h(R) = 0: (14) p h h p This constraint provides a quadratic equation for the pattern speed in terms of the eigenfunction p h(R); its unique feature is that the self-gravity of the disk does not have to be evaluated explicitly. In general, one of the roots of equation (14) is spurious and we shall assume that the correct root is the one with the smallest absolute value. The motivation for this choice is that we focus on the distorted tilt mode, which has a small pattern speed. Numerical solutions of the eigenvalue equation (8) for various disk and halo density pro les are given by HT, Sparke (1984) and SC. The following special cases o er useful insight: 1. In the absence of disk self-gravity, equation (8) reduces to ! = m (R)   (R); (15) h as does (14) when m = 1. In general, the combination of terms on the right side will depend on radius, so the eigenfunctions are singular (van Kampen modes), and any spatially extended initial disturbance will damp by phase mixing. 2. If the halo is spherical, the disk is neutrally stable to tilting. To see this from equation (8) we use equations (5) and (9) to write Z 1 G 2 0 0 0 0 0 0 (R) = dR R (R )[RH (R; R ) R I (R; R )]; (16) d 1 R 0 it follows that when m = 1 and the halo is spherical ( = ) a solution of (8) is ! = 0, h h h(R) / R (the `tilt mode'). 3. For a nearly spherical halo, the tilt mode is distorted and acquires a small but non-zero pattern speed. In equation (14) we may therefore drop the term proportional to ; p 2 moreover, in a rst approximation we can replace the eigenfunction h(R) by the tilt mode h(R) / R. Thus we obtain (SC) R 1 2 2 3 (  )(R)R dR 0 h h R ' ; (17) p 1 3 2 (R)(R)R dR 0 note that the angular frequency in the denominator includes the contribution from both disk and halo, while the frequencies in the numerator represent only the halo contribution. Equation (17) simply says that the precession rate of the warped disk is the torque from the halo divided by the o -axis angular momentum (Kuijken 1991); we shall refer to it as the Sparke-Casertano formula for the pattern speed. { 6 { The distorted tilt mode is closely related to the Laplace invariable plane : each annulus of 3 the disk is in the Laplace plane determined by the halo and the other annuli. Thus we shall refer to this mode as the `Laplace mode' (SC use the term `modi ed tilt mode'). 2.2 Energy and angular momentum To determine the energy E associated with a bending disturbance Z (R; t), we assume that the w disturbance has been generated by an external potential U (r; t), where r = (R; ; z ) = (R; z ). If ext the disk has density (r; t) and velocity v(r; t) then the rate at which the external potential does work on the disk is Z Z dE w dt = drv  F = drv  rU : (18) ext ext We may write v = (0; R; DZ=Dt); the radial and azimuthal velocities have been set to their unperturbed values since we are working to rst order in the small quantity Z (R; t) and to this order the horizontal and vertical motions are decoupled. Thus dE w Z = dr [ U + (Z; + Z; )U ] ; ext; t  ext;z dt Z = dr [  U Z; U +  Z; U ] ; (19) ; ext t ext;z ;z  ext where the second line follows from integration by parts. The density of our razor-thin disks may be written (r; t) = (R) [z Z (R; t)]. Thus  =  (z Z )Z; =  Z; : (20) ;  ;z  0 Substituting this result in equation (19) we nd that the rst and third terms cancel, so that Z Z = drZ; U = dRF Z; (21) t ext;z ext t dE w dt where as usual F is the z -component of the external force (note that the expression is not R ext dRF (DZ=Dt), as one might expect from considering only vertical motions). ext We may eliminate the external force F from equation (21) using equations (1), (3) and (6). ext Then using integration by parts with respect to , we can convert the right side of equation (21) 3 Consider a attened planet that is accompanied by a massive satellite on an inclined orbit. A test particle orbiting the planet in the Laplace plane will remain in that plane (except for short-period oscillations) despite the perturbations from the planet's equatorial bulge and the massive satellite. The normal to the Laplace plane lies in the plane determined by the spin and orbital angular momentum vectors of the planet and satellite, while the orientation of the plane is xed so that the angular momentum vectors of the planet, satellite and test particle all precess at the same rate. { 7 { into the total time derivative of a quantity which must equal the energy E to within a constant. w Setting this constant to zero when Z = 0, we obtain Z Z G [Z (R; t) Z (R ; t)] 0 2 1 2 2 2 2 2 0 0 E = dR(R)[Z; + Z Z; ] + dRdR (R)(R ) : (22) w 2 t h  4 jR R j 0 3 In the absence of an external force, the energy (22) is conserved . 4 Note that the energy of a disturbance is not guaranteed to be positive, since the rst integrand can be negative. HT examined the sign of the energy for azimuthal wavenumbers m = 0 (axisymmetric) and m = 1 (tilt or warp). When m = 0 Z; = 0 so equation (22) implies that the  energy is positive (or zero, in the trivial case  = 0, Z =constant, corresponding to the translation h of an isolated disk). Since unstable modes must have zero energy, all m = 0 modes are stable. In the case m = 1 it is useful to decompose the angular speed into contributions from the disk and halo (eqs. 4, 5). HT then show that Z Z G [Z (R; t) Z (R ; t)] 0 2 1 2 2 0 0 dR(R) (R)Z; + dRdR (R)(R ) 2 Z G d  0 3 4 jR R j = dRdR d(R)(R )[RZ (R ; ; t) R Z (R; ; t)] I (R; R ); (24) 1 0 0 0 0 2 0 4 where I is de ned in equation (9). Thus the energy of an m = 1 disturbance may be written m Z Z G E = dR(R)[Z; +( )Z ]+ dRdR d(R)(R )[RZ (R ; ; t)R Z (R; ; t)] I (R; R ): 1 2 2 2 2 0 0 0 0 2 0 w 2 t h h 4 1 (25) If there is no halo ( = = 0) the energy is positive or zero; zero energy corresponds to the h h disturbance Z = AR cos(  ) where A and  are constants (physically, an isolated disk is 0 0 neutrally stable to a uniform tilt). Since unstable modes must have zero energy, all m = 1 modes of an isolated disk are stable, a conclusion due to HT. When the disk is embedded in a halo, m = 1 disturbances can have positive or negative energy. We shall call an axisymmetric halo `vertically sti ' if  > at all radii (which is h h generally the case if the disk lies in the equatorial plane of an oblate mass distribution), and `vertically soft' if  < (which occurs for prolate distributions). Equation (25) shows that the h h total energy of an m = 1 disturbance is positiveso that m = 1 modes are stablewhenever the halo is vertically sti . 4 This result for the energy is perhaps somewhat unexpected: in the absence of self-gravity, the most natural way to derive an energy is to treat the disk particles as independent vertical harmonic oscillators. This procedure yields Z Z 0 1 2 2 2 2 2 2 2 2 1 E = dR(R)[(DZ=Dt) +  Z ] = dR(R)(Z; + Z + Z; +2 Z; Z; ); (23) w h t h  t  2 2 which is not the same as the rst term of (22). Both E and E are conserved if F = F = 0, but equation (22) w sg ext w 0 is the only energy-like quantity conserved when self-gravity is present (except in the special case of uniform rotation, which is described at the end of this section). { 8 { The expression for the energy (22) can be rewritten in a simpler form. We operate on equation R (1) with dR(R); setting F = 0 we obtain ext Z dR(R)(ZZ; +2 ZZ; + ZZ; + Z ) tt t  h 2 2 2 Z Z (R; t)[Z (R; t) Z (R ; t)] 0 = G dRdR (R)(R ) 0 0 jR R j 0 3 Z 0 2 G [Z (R; t) Z (R ; t)] 0 0 = dRdR (R)(R ) ; (26) 2 jR R j 0 3 where the last expression is obtained from the second by interchanging primed and unprimed variables and averaging with the original expression. Using this result to eliminate the last term from equation (22) we nd Z 1 2 2 2 2 E = dR(R)[Z; ZZ; 2 ZZ; Z; ZZ; ]: (27) w tt t  2 t  The terms proportional to cancel upon integrating by parts with respect to . Thus 2 Z 1 2 E = dR(R)[Z; ZZ; 2 ZZ; ]: (28) w tt t 2 t For a disturbance of the form (7) with m 6= 0 and real frequency ! = m , we have p Z 2 2 E = m dRR(R)jh(R)j ( ): (29) w p p Thus disturbances with zero pattern speed have zero energy; if the pattern speed is negative (retrograde precession) the energy is always positive; if the pattern speed is small and positive the energy is negative. We can also work out the z -component of angular momentum associated with a disturbance excited by an external potential: Z dJ w dt = drU ext; Z = dr; U  ext Z = dr; Z; U z  ext Z = drU Z; ; (30) ext;z  where the third line follows from equation (20). We eliminate U = F using equations (1) ext;z ext and (3) and derive an expression for the angular momentum by a similar procedure to the one used to derive equation (22). In this derivation we use the results Z; Z; = (Z; Z; ); (Z; ); (31)  tt  t t  2 t 1 2 { 9 { and Z Z 0 0  Z; (Z Z ) 0 dR(R)Z; F = G dRdR (R)(R )  sg jR R j 0 3 Z 0 2 G [(Z Z ) ]; 0 0  = dRdR (R)(R ) 2 jR R j 0 3 Z G @ 0 0 0 2 0 3 = dRdR (R)(R )(Z Z ) jR R j ; (32) 2 @  which is zero, because the integrand to the left of @ =@  is symmetric under interchange of primed and unprimed variables, while @ jR R j =@  is antisymmetric. 0 3 In this way we nd Z J = dR(R)(Z; Z; + Z; ): (33) w t   2 Thus, in the absence of external forces, there are two independent conserved quantities that are quadratic in the disturbance strength. These quantities correspond to the energy E and w z -component of angular momentum J . For a disturbance of the form (7) with real frequency ! = m , we have p w Z J = m dRR(R)jh(R)j ( ): (34) w p 2 2 Comparison of equation (34) with equation (29) shows that E = J : (35) w p w We close this section by considering the special case of disks with constant angular speed, (R) = =const. Combining equations (22) and (33) we obtain the conserved quantity 0 Z Z G [Z (R; t) Z (R ; t)] 0 2 1 2 2 2 0 0 E J = dR(R)[(Z; + Z; ) +  Z ] + dRdR (R)(R ) : w 0 w t 0  2 h 4 jR R j 0 3 (36) Since this quantity is non-negative, all bending disturbances in a uniformly rotating disk are stable. 2.3 The WKB approximation The equation of motion (1) for bending waves can be solved analytically in the WKB or short-wavelength limit. This solution o ers useful insights, even though the WKB approximation is poor for the bending waves actually seen in galaxies. In the WKB limit the radial wavelength  is much shorter than any relevant radial scale length, i.e.,   R;   ; (37) dR 1 d ln (R) but long compared to the wave amplitude, { 10 {   jZ j: (38) We write the vertical disturbance as Z (R; t) = Re A(R; t)e e ; (39)   R R i K (R)dR i(m!t) where K (R) = 2=(R) is the local wavenumber and A(R; t) is a slowly varying amplitude. The vertical gravitational force from the disturbed disk (eq. 6) can be evaluated using local Cartesian coordinates, R = R + xe + ye , where at R the xaxis points radially outward and x y 0 the yaxis points azimuthally. Because the radial wavelength is short, we neglect the radial variation in the surface density (R ) and the azimuthal variation in Z , and write 0 0 i(K x!t) i!t Z (R ; t) = Re[A e ]; Z (R; t) = Re[A e ]; (40) 0 0 where R R i K (R)dR im(R) A  A(R; t)e e : (41) 0 Then equation (6) yields F (R; t) = G(R) Re A e dx dy sg 0 2 2 3=2 (x + y )  Z  i!t [1 exp(iK x)] = 4G(R)jK jZ (R; t) du 0 0 (1 + v ) 2 3=2 2 u Z Z 1 1 dv (1 cos u) = 2G(R)jK jZ (R; t): (42) The WKB dispersion relation follows from the dynamical equation (1), D(! ; K )  [! m (R)] 2G(R)jK j  (R) = 0: (43) h 2 2 At the corotation radius, the angular speed of the disk equals the pattern speed, (R) = = !=m. p At the vertical resonance radii K (R) = 0, that is, ! = m (R)   (R) (cf. eq. 15). Because 2 2 2 h the dispersion relation can only be satis ed when m ( )   , there is a forbidden region p h between the two vertical resonance radii and including corotation, in which WKB bending waves cannot propagate. The radial group velocity for WKB bending waves is (Toomre 1969, Whitham 1974) @ ! G(R) sgn(K ) c = = : (44) g @K m[ (R) ] p Thus the direction of propagation depends on whether the waves are leading (K < 0) or trailing (K > 0), and on whether they are inside corotation ( > , assuming that the angular speed p decreases outward) or outside. { 11 { The equation of motion for vertical disturbances (1) can be derived from a Lagrangian. The dynamics in the WKB limit are therefore described by an averaged Lagrangian that is obtained by averaging over the short-wavelength oscillations (Whitham 1974), L(! ; K; A) = RjAj D(! ; K ); (45) 2 1 2 where jAj is the wave amplitude and D(! ; K ) is the dispersion relation (43). The energy density of the wave per unit radius is @L 2 2 (R) = ! = m R(R) ( )jAj : (46) p p @ ! The energy density is positive if the pattern speed is prograde ( > 0) and outside corotation p ( < ), or if the pattern speed is retrograde. The total energy of the disturbance is p Z E = (R)dR; (47) w which is the same as the energy derived in the previous section (eq. 29)in this case it happens that the WKB approximation is exact. The density of zangular momentum per unit radius is (Bertin & Mark 1980) which is exactly consistent with equation (34). j (R) = m = ; (48) @L (R) @ ! p 3 Dynamical Friction Between a Warped Disk and a Spherical Halo So far we have treated the halo potential as a static background eld, which in uences the disk only through its contribution to the vertical and radial restoring force. In fact, the halo is composed of stars whose orbits respond to the time-varying gravitational eld of the warped disk. The gravitational attraction between the coherent response of the halo and the warped disk transfers energy and angular momentum between the two components, thereby dampingor possibly excitingthe warp (dynamical friction). To describe this response, we write the halo potential as U (r) = U (r) + U (r; t), where the h h h 0 1 two terms denote the static, unperturbed halo potential and the potential arising from its response to the warp. The corresponding densities are  (r) and  (r; t). For consistency with Section h h 0 1 2.2, we de ne the energy of the bending disturbance in the disk, E , to include the kinetic and w self-gravitational energies of the disk material plus its potential energy in the static eld U , but h 0 not its potential energy in the response eld U ; the forces arising from the latter eld are treated h 1 as an external force F = rU . Thus ext h 1 { 12 { dE w dt Z Z = dr v  F = drr  ( v )U d d ext d d 1 Z Z @  d @   d 1 0 h 1 h = dr U = G drdr h 0 @ t @ t jr r j Z Z @U @U 1 1 d w = dr = dr : (49) h h @ t @ t where we have integrated by parts and used the continuity equation for the disk mass; the subscript `d' denotes disk, and in the last line we have replaced the disk potential U by the warp potential d U since the potential of the unwarped disk is stationary. To rst order in the perturbation w strength we may write  = r  ( r), where r(r; t) is the Lagrangian displacement of the h h 1 0 halo stars caused by the perturbation. The last line of equation (49) may now be written Z dE @U w w 0 = dr r  r : (50) dt @ t h There are two equivalent ways to determine the dynamical friction exerted by the halo on the disk: (i) calculate the density and potential perturbation induced in the halo by the disk and then integrate the resulting gravitational force from the halo over the disk mass distributionthis is analogous to the standard discussion of Landau damping in a collisionless plasma (cf. Ichimaru 1973) and corresponds to the rst line in equation (49); (ii) calculate the rate of change of energy of each halo star due to the warp, and integrate these energy changes over the halo (which corresponds to eq. 50). We use the second approach in this paper. Our calculations are based on the action-angle formalism developed in papers by Lynden-Bell & Kalna js (1972), Tremaine & Weinberg (1984), and Weinberg (1986), which may be consulted for details. Our analysis employs three (not very good) approximations: (i) the warp is small, so that we can use perturbation theory in which the small parameter is the amplitude of the warp; in this approximation both r and @U =@ t are proportional to the warp amplitude so dE =dt d w is second-order in the amplitude (this is the same order as the energy of the warp computed in Section 2.2; thus the damping rate E =E is independent of amplitude); (ii) the unperturbed w w _ potential (halo plus at disk) is spherical; this is certainly incorrectand moreover inconsistent, since we assume that the halo is axisymmetric but non-spherical when estimating the pattern speed and energy of bending modesbut orbits in spherical potentials are easier to analyze than orbits in attened potentials and the dynamical friction force should not depend strongly on the shape of the halo orbits; (iii) we neglect the self-gravity of the halo response; again, this is certainly incorrect but unlikely to cause more than a factor of two or so error in the results. In the unperturbed state (no warp) the motion of the halo stars is described by an autonomous Hamiltonian H . The phase-space coordinates of a star are speci ed using the action-angle 0 variables of this Hamiltonian (I; w); in these coordinates H = H (I) so that Hamilton's equations 0 0 become _ @H @H 0 0 I = = 0; w = _  (I); (51) @w @ I { 13 { with solutions I(t)  I = constant w(t)  w (t) = (I )t + constant: (52) 0 0 0 The density of halo stars in phase space is described by a distribution function (hereafter df) F and Jeans' theorem implies that F = F (I). When a warp is present the Hamiltonian becomes Hamilton's equations become H = H + U : (53) 0 w _ @U @U w w I = ; w = (I) + _ : (54) @w @ I The perturbing potential can be expanded in an action-angle Fourier series (cf. Appendix A), 8 9 < = 1 1 X X i(lw! t) l U (I; w; t) = Re (I)e : (55) w l l l 1 2 3 : ; l =0 l ;l =1 3 1 2 The rst-order corrections to the orbit (52) caused by the perturbation are @  @  I = ; w = (56) @w @ I where 8 9 Z t < i(lw! t) = 1 1 X X e l  = U dt = Re ; (57) w l l l 1 2 3 1 i(l  ! ) l : ; l =0 l ;l =1 3 1 2 and these quantities are evaluated along the unperturbed orbit (I ; w ). We have assumed that 0 0 ! has a small positive imaginary part,  , as if the potential were turned on slowly in the distant l past. The integrand of equation (50) may be written @U @ U @ U w w w 2 2 r  r = (I ; w ; t)I + (I ; w ; t)w: (58) 0 0 0 0 @ t @ I@ t @w@ t We evaluate this expression using equations (55){(57). Since the df depends only on the actions we can average the result over the unperturbed angles w . As described in Tremaine & Weinberg 0 (1984), the terms in this expression involve combinations of periodic exponentials, most of which vanish upon averaging over initial phases . Combining the surviving terms we nd 5   1 1 X X 2t @U @ 2(! l  )e w 2 l ( ) r  r = l  j (I)j ; (59) @ t @ I 2(jl  ! j +  ) w l l =0 l ;l =1 3 1 2 l 2 2 5 Or are oscillatory in time, in the case l = 0. 3 { 14 { where ! is now taken to be real as the contribution from the small imaginary part  is written l out explicitly. We now use the identity  lim =  (x): (60) !0 x +  + 2 2 Integrating over the df of halo stars nally gives the rate of change of the disk energy (cf. eq. 50), 6 Z Z   _ E = dIdwF (I)r  r = (2 ) dIF (I) r  r d @U w 3 @U w @ t @ t w 1 1 Z X X 4 2 @ F (I) = 4 dI(l  )l  j (I)j  (l  ! ) @ I l l l =0 l ;l =1 3 1 2 1 1 Z X X 4 2 4 dS  l(l  )F (I)j (I)j  (l  ! ); (61) l l l =0 l ;l =1 3 1 2 where dS is the area element on the boundary of action space and we have set t = 0. For most dfs the last (surface) term vanishes, as the following argument shows: Let dS be an area element on the boundary of action space. The ux of stars through this element with angles in the interval dw is dwdS  IF (I); using equations (54) and (55) this ux equals _ P dwdS  lF (I)Refi (I) exp[i(l  w ! t)]g. The terms in this sum are linearly independent l l l functions of w. Since stars cannot leave action space, each term in the sum must be zero. Thus dS  lF (I) (I) must vanish, which implies that the surface term in (61) must vanish. (This l argument can fail if the df diverges on the boundary of action space.) If the surface term vanishes, equation (61) simpli es to _ 4 2 @ F (I) 1 1 Z X X E = 4 dI(l  )l  j (I)j  (l  ! ): (62) d l l l =0 l ;l =1 3 1 2 @ I Only resonant orbits with l  (I) = ! contribute to the energy transfer; a particular resonant l triplet l = (l ; l ; l ) contributes to a net damping or growth of the perturbation depending on the 1 2 3 components of (I) and the gradient of F (I) along the resonant triplet. Finally, we note that the formalism we have developed is related to Goodman's (1988) stability criterion for galaxies, which is based on the sign of a quadratic functional of a trial perturbation. Consider a trial perturbation U with time dependence e with  real (not e as we have w t im t+t p assumed so far). Then Goodman shows that a sucient criterion for instability is Z 0 2 G [Z (R; t) Z (R ; t)] _ 0 0 E > dRdR (R)(R ) ; (63) w 2 jR R j 0 3 6 Our result is consistent with equation (65) of Tremaine & Weinberg (1984), who assume l 6= 0 and a single pattern 3 _ speed, so that ! = l and the torque is related to the rate of energy change by E =  (Jacobi's integral). 3 p p z l We have assumed that the evolution of the warp is suciently rapid that we are in the `fast' limit of Tremaine and Weinberg, that is, that the damping rate is large compared to the perturbation strength; this assumption is always justi ed if the warp amplitude is suciently small, because the damping time is independent of amplitude. { 15 { where E is de ned by equation (49) and evaluated using equation (59)compare Goodman's w _ equation (24). 3.1 Halos with isotropic distribution functions An important special case is an unperturbed halo df that depends on energy alone, F = F (E ) = F [H (I)]. Then equations (51) and (62) imply 0 _ 4 2 2 dF 1 1 Z X X E = 4 dI(l  ) j (I)j  (l  ! ): (64) d l l l =0 l ;l =1 3 1 2 dE In most model galaxies with F = F (E ) the halo df is a decreasing function of energy, dF =dE < 0; following Goodman (1988), we call these `iddf halos' (for `isotropic, decreasing distribution function'). In iddf halos, equation (64) implies that E < 0. In other words, in iddf halos, a d _ warped disk always loses energy to the halo. Thus disk disturbances with positive energy damp, while disturbances with negative energy grow. Equation (22) shows that bending disturbances with a single azimuthal wavenumber m  0 have positive energy if  (R) > m (R) at all radii. Thus in iddf halos, dynamical friction damps h all bending disturbances so long as  (R) > m (R) at all R: (65) h For m = 1 disturbances, equation (25) implies that the damping criterion (65) can be replaced by the stronger damping criterion  (R) > (R) at all R; (66) h h which is the condition that the halo is vertically sti . In other words, in vertical ly sti iddf halos, al l m = 1 bending disturbances damp. Bertin & Mark (1980) examined the evolution of short-wavelength bending waves of a thin disk embedded in a halo. The halo was assumed to have an isotropic Maxwellian velocity distribution, and the disk dynamics were analyzed in the WKB approximation. They concluded that dynamical friction from the halo would excite all bending waves propagating inside the corotation radius. Toomre (1983) has pointed out that the validity of this conclusion is suspect, because Bertin and Mark ignored the halo's contribution to the vertical frequency, e ectively setting  = 0 and thereby ensuring that the halo was vertically soft. The more general approach h developed in this paper allows a rapid derivation of the correct answer: if, like Bertin and Mark, we assume an iddf halo, then bending waves are excited by dynamical friction from the halo if and only if they have negative energy. From equation (46) the energy density of a WKB bending wave is negative if and only if 0 < < (R); (67) p i.e. for prograde waves inside corotation. { 16 { Equation (67) is the condition for excitation of WKB bending waves for any iddf halo. The WKB dispersion relation (43) for m = 1 waves implies that j j   , so a necessary condition p h 7 for excitation is  < , which is consistent with the damping criterion (65) . These results imply h that the excitation of WKB bending waves by the halo is considerably more dicult than Bertin & Mark (1980) assumed. We shall not pursue the analysis of WKB bending waves further, since the WKB approximation is invalid for most observed galactic warps. Instead we shall focus on numerical evaluation of damping and excitation rates for more realistic warps. 3.2 A model for the halo In this section we describe the analytic model for the halo df that we shall use in our numerical estimates of damping rates. In a universe dominated by collisionless dark matter, galaxy halos form by the gravitational collapse of random peaks in the cosmological density eld (cf. Warren et al. 1992 and references therein). Both analytic arguments and numerical simulations show that halos have power-law density pro les that are roughly consistent with the pro les  / r implied by observations 2 of at rotation curves of spiral galaxies. The simulated halos tend to be triaxial with minor to ma jor axis ratios c=a  0:5, although the uncertain e ects of dissipation are likely to reduce the triaxiality (Katz & Gunn 1991; Dubinski 1994). The velocity ellipsoid in the halo is radially elongated: the ratio of the mean-square radial velocity to the mean-square tangential velocity is given by 2hv i=hv i ' 1:4 (Dubinski 1992), whereas an isotropic dispersion tensor would have the r t 2 2 value unity. Tidal torquing from neighboring structures gives the halo a net angular momentum; Dubinski nds that the ratio of the maximum rotation velocity to the central line-of-sight velocity dispersion is roughly correlated with the dimensionless spin parameter, v =  4, where max c  = J jE j G M has a median value of about 0:05 (White 1984; Barnes & Efstathiou 1987; 1=2 1 5=2 Zurek et al. 1988). To simplify our calculations we ignore the intrinsic triaxiality of the halo and the contribution of the disk to the total potential: thus we assume a halo df, halo orbits, and a rotation curve for the disk that are consistent with a spherical potential provided entirely by the halo. This approximation is not strictly consistent, since we assume that the halo is axisymmetric but non-spherical when estimating the pattern speed of bending modes or calculating the mode energy 7 In the WKB approximation the damping criteria (65) and (66) are generally the same when m = 1, because there is no distinction between and . The reason is that the dispersion relation (43) can usually be satis ed h in the WKB limit (jK Rj  1) only if the surface density  is small, so the disk's contribution to the total rotation speed is negligible. { 17 { (see eq. 75 at the end of this subsection). The bias caused by the spherical approximation should lead us to underestimate the damping and excitation rates in the likely case that the halo is attened towards the disk plane, since the halo density will be higher and the velocity of the halo stars will be lower than we have assumed. Our model for the halo df: (i) assumes that the halo potential generates a at rotation curve with circular speed v , c   2 r which implies a halo density U (r) = v ln ; (68) h c r s 2 2 r U (r) v c (r) = = (69) 4G 4Gr 2 (r is a ducial radius, which eventually scales out of the problem); (ii) allows for a range of s velocity ansiotropy, expressed by the ratio 2hv i=hv i; (iii) allows a non-zero mean rotation speed r t 2 2 hv i, which is taken to be independent of radius.  The df that we use is F (E ; J; J ) = F J e (1 + J =J ); (70) z 0 z 2( 1) E = 2 where E is the energy, J is the angular momentum, and J is the component of angular momentum z normal to the unperturbed disk. The parameter controls the halo rotation, while controls the velocity anisotropy; we must have j j  1 so that the df is non-negative, and > 0 so that the spatial density is nite . In order that this df generates the density (69) we must have 8 2 +1=2 2 c v 1  = ; F = ; (71) 2 4 r Gv s c J ( ) s 0 5=2 2 2( 1) where J = r v . In the special case = 0, = 1, the df and potential reduce to those of the s s c singular isothermal sphere (the normalization is the same as Weinberg 1986, eq. 24). In turn, F is a special case of the larger classes of scale-free dfs discussed by Gerhard (1991) and Evans (1994). The mean-square velocities in the radial and tangential directions are 2 2 2 2 hv i =  ; hv i = 2  ; (72) r t so the ratio of the principal axes of the velocity ellipsoid is given by 2 2hv i 1 r 2 = ; (73) hv i t 8 Galaxies with predominantly radial orbits are likely to be unstable; in fact Palmer & Papaloizou (1987) have shown that all galaxies with distribution functions that are unbounded as J ! 0 are unstable. However their proof assumes that the resonant combination of frequencies 2 / J as J ! 0; while this is generally true for galaxies 1 2 with a at central core, in the singular isothermal sphere 2 / 1=j ln J j as J ! 0, and the Palmer-Papaloizou 1 2 proof does not apply. In any case, it is not crucial for our purposes to work with stable halo models, because the self-gravity of the halo responsewhich is central to the instabilityis neglected in our calculations of dynamical friction. { 18 { a halo with nearly radial orbits corresponds to ! 0 while one with nearly circular orbits has ! 1. The mean rotation speed is hv i ( + ) 1  v c = sin : (74) 2 1=2 2 ( ) The cosmological simulations described above suggest that typical parameter values are ' 0:7 (from the velocity anisotropy) and ' 0:4 (we have identi ed v with hv i( =  ) and 3 max  2 c 1 2 with hv i + hv i). r t 2 2 Finally, when estimating the pattern speed or energy of a bending mode, we generalize the expression for the halo potential (68) to an axisymmetric halo:   1 2 2 2 2 U (R; z ) = v ln R + z =q ; (75) h 2 c  where q is the axial ratio of the equipotential surfaces and   1 q is their ellipticity ( > 0     for oblate halos and < 0 for prolate halos). In this case the vertical and azimuthal frequencies in the eld of the halo are (eqs. 2 and 5) v c v c  = ; = : (76) h h q R R  3.3 A model for the warp To calculate the damping rate in a realistic warped galaxy we need to know the shape of the warp h(R) (eq. 7) and its associated gravitational potential U (r). Rather than solving equation (8) to w obtain an exact eigenfunction, we have chosen a simple analytic parametrization of the warp. We assume that the surface density of the disk is exponential, (R) =  e ; (77) 0 R=R d so that the total disk mass M = 2 R . We set the parameter d 0 d 2 GM d 2 R v d c = 1:6; (78) in all of our simulations, which is roughly correct for our Galaxy. We write h(R) = R f (R=R ) d d where R is the scale length of the disk. If the mode has real eigenfrequency then we can assume d that h(R) is real (by the anti-spiral theorem of Lynden-Bell & Ostriker 1967); in this case the line of nodes is radial, which is approximately consistent with observations. Our parametrization has the form f (x) = C x e ; (79) n x where C is an arbitrary normalization (recall that the damping rate is independent of the amplitude). This parametrization permits us to model two distinct types of warps. (i) For n = 1 { 19 { the inner part of the disk is tilted relative to the halo (as in the Laplace mode) and the outer disk is at; in this case equation (79) gives a warp whose curvature in its outer parts is more gradual than observed warps (see Figure 1a), but this is not a ma jor defect since most of the interaction with the halo comes from the tilted inner disk, which is not curved either in our model or in real galaxies. (ii) For n  1 the inner disk is at and the outer disk is tilted. In this case the shape of the warp can be made to closely resemble observed warps (except that at very large radii the disk becomes at againits maximum height is at x = n=but if x  1 the surface density in max max the region where the disk becomes at again will be negligible). An advantage of the functional form (79) is that the spherical harmonic expansion of the warp potential U (r)Y (; ) can be evaluated in terms of incomplete gamma functions (see Appendix ln ln B). Fig. 1a. The solid line shows the shape of a typical warp used in our estimates of the warp damping/excitation rate, derived from eq. (79) with C = 1, n = 1,  = 0:2. The dotted and dashed lines indicate the rst two non-zero components, U (r); U (r), of the spherical harmonic 21 41 expansion of the potential of the warped disk with m = 1; the scale for the potentials is shown on the right hand axis in units of GM =R . d d { 20 { Fig. 1b. The pattern speed in units of v =R (the angular speed of the disk at one scale length), c d as a function of the ellipticity of the halo potential. The pattern speed is computed using the warp shape in Figure 1a and eq. (14)the solution of the quadratic equation with the smaller value of j j is shown. The dotted line corresponds to the Sparke-Casertano formula (eq. 17). p Fig. 1c. Energy of the warp shown in (a) as a function of the ellipticity of the halo potential, computed from eqs. (29), (77) and (78). { 21 { Figure 1a shows an example of a warp shape parametrized by equation (79) with n = 1 and  = 0:2, along with the radial potential functions associated with the rst two nonvanishing m = 1 spherical harmonics, U (r), U (r). Figure 1b shows the pattern speed for this warp, calculated 21 41 using equation (14), as a function of the halo ellipticity  (eq. 75). The dotted line shows the  Sparke-Casertano formula for the pattern speed (eq. 17), which is a good approximation to the more accurate formula. Figure 1c shows the energy of the warp (eq. 29) with C = 1. The gure illustrates that warps embedded in vertically sti halos ( > 0) generally precess in a retrograde  direction ( < 0) and have positive energy, while warps embedded in vertically soft halos ( < 0) p  precess in a prograde direction ( > 0) and have negative energy. p 4 Calculation of Damping and Excitation Rates We now calculate the rate of change of the warp energy, and the resulting damping or growth rate caused by dynamical friction from the halo. Weinberg (1986) has described similar calculations, based on a formula analogous to (62), for the rate of decay of a satellite on a circular orbit in a spherical potential. Thus we have simply summarized the principal formulae in Appendix A and refer the reader to Weinberg (1986) for details. The formula (62) for the rate of change of the warp energy is combined with equation (29) for the energy to obtain the characteristic evolution rate 1 dE w = ; (80) E dt w > 0 implies excitation and < 0 implies damping. In our model the evolution rate is a function of the velocity anisotropy of the halo (parametrized by ), the rotation of the halo (parametrized by ), the warp pattern speed (written in dimensionless form as R =v ), and the p d c shape of the bending disturbance; it is proportional to the mass of the disk but independent of the amplitude of the disturbance (for small perturbations) since both the energy and the rate of energy change scale with the square of the amplitude. We evaluate the action-angle transform of the perturbing potential using equations (A1){(A3) and (A23), along with equations (79) and (B9){(B10) for the warp shape and potential. We treat the integrals in equations (A10), (A11) and (A23) as the solutions to a coupled set of ordinary di erential equations, which we integrate using an adaptive Runge-Kutta routine. We use Romberg integration to evaluate the integral in equation (A26), summing over the possible resonant triplets in the integrand itself, rather than evaluating an integral for each term separately. Since the number of resonant triplets grows / l , we must cut o the summation at a relatively 3 small maximum harmonic l . Experiments on a few points show that E (l ) = E + C l , with max max max _ _ ' 1. In practice we evaluated E (l ) at l = 2; 4, and 6, tting these values to a quadratic max max _ 1 in l , and extrapolate to l ! 1. We believe that the damping/excitation rates we have max max obtained are accurate to within 3%. { 22 { We have veri ed our results in the epicycle limit (1   1) where the functional form for the potential transforms becomes analytic, and also compared our formulas with similar transforms calculated by Weinberg (1985). Fig. 2. The damping rate = E =E as a function of the ellipticity of the halo potential. The w w _ warp shape is shown in Figure 1a. The halo df is described by eq. (70) with = 1 (isotropic velocity distribution). Warps with > 0 grow, while those with < 0 damp. Damping or excitation occurs within a Hubble time if jj > 0:0015 (assuming R =v = 1:5  10 y). The small errorbar centered d c 7 at  = 0 corresponds to the range of damping rates where damping or excitation does not occur  within a Hubble time. The dotted, solid, and dashed lines correspond respectively to halos with rotation parameters = 1; 0; 1 described in the text; these values span the range of possible rotation rates. Figure 2 shows the dimensionless damping/excitation rate R =v as a function of the d c halo ellipticity  . The warp has the shape shown in Figure 1a and is embedded in a halo  with an isotropic velocity distribution, 2hv i=hv i = 1. For parameters corresponding to those r t 2 2 of our Galaxy, R =v = 1:5  10 y, the damping/excitation time is less than 10 y when d c 7 10 jjR =v > 0:0015. d c { 23 { We plot results for three values of which span the range of possible rotation rates consistent with our model (eq. 74). First consider the non-rotating halo ( = 0) with isotropic velocity distribution (an iddf halo in the notation of Section 3.1). As shown in that section, a warp always loses energy to an iddf halo. Thus warps with positive energy damp. However, if the halo is vertically soft ( < 0, which obtains if the disk lies in the equatorial plane of a prolate  halo potential), the energy of the warp is negative (Figure 1c). Thus, a warp embedded in a non-rotating, vertically soft iddf halo can grow in amplitude, while one embedded in a vertically sti halo must damp. Halo rotation tends to increase the rate of change of warp energy E if the halo rotates in w _ the same sense as the pattern speed, and decreases E if the halo rotates in the opposite sense. w _ Thus, adding rotation tends to increase (or decrease) the warp amplitude if the sign of =E p w is positive (negative). If the pattern speed is small, E = is always negative (eq. 29); thus halo w p rotation contributes to warp growth only if the halo rotates in the opposite sense to the disk. Fig. 3. Same as Figure 2, but with for a halo with predominantly radial orbits, 2hv i=hv i = 1:5. r t 2 2 Note that the vertical range of the graph is a factor of ten larger than Figure 2. Figure 3 shows the damping/excitation rates for warps embedded in halos with predominantly { 24 { radial orbits, 2hv i=hv i = 1:5. The damping rate is much larger than in the isotropic case; in fact 2 2 r t > so large (jj j j) that our treatment of the warp as a slowly evolving normal mode is suspect.  p Fig. 4. Same as Figure 2, but with for a halo with predominantly circular orbits, 2hv i=hv i = 0:5. r t 2 2 Note that the vertical range is only half as large as Figure 2. Figure 4 shows the damping/excitation rates for warps embedded in halos with predominantly circular orbits, 2hv i=hv i = 0:5. In this case energy ows from the halo to the warp, a consequence r t 2 2 of the strong anisotropy in the phase space distribution function. Thus positive energy warps ( > 0) grow, while negative energy warps damp.  Figure 5 shows the dimensionless damping rate as a function of the halo velocity anisotropy, for a warped disk with a retrograde pattern speed, R =v = 0:5; this pattern speed corresponds d p c to a halo whose equipotentials have ellipticity  = 0:5 (Figure 1b). For halos with predominantly  circular orbits the warp is excited, while in halos dominated by radial orbits it is damped. An interesting feature is that the damping rate diverges as hv i=hv i ! 2, corresponding to ! . r t 2 2 2 1 We show in Appendi<x C that the apparent divergence is an artifact of the perturbation theory that we have used: according to equation (61) the rate of energy loss E arising from the response w _ of the halo to a perturbation potential is O( ), but in fact the response is O( ) where 2 p { 25 { Fig. 5. The damping rate as a function of the velocity anisotropy of the halo. The halo equipotentials are oblate,  = 0:5. The dotted, solid, and dashed lines correspond to halos with  dimensionless rotation = 1:0; 0:0; 1:0. Halos with predominantly circular orbits excite the warp, while halos with predominantly radial orbits damp the warp. { 26 { p = min(2; 2 + 1); thus for < the energy loss rate is of lower order than O( ). Our analysis 1 2 2 1 is not powerful enough to determine the damping rate for halos with < ; nevertheless it is clear 2 that the damping is very strong when the orbits of the halo stars are predominantly radial. We have also examined the damping of disks that are only warped relative to the halo in their outer parts (eq. 79 with n = 5,  = 1). Although the damping/excitation times are substantially longer, signi cant damping still generally occurs in less than a Hubble time. 5 Discussion and Conclusions The most striking feature shown by Figures 2{4 is the rapid evolution of the warps from dynamical friction. All of the models we examined had damping/excitation times less than a Hubble time when the halo ellipticity was larger than 0.2, and in general the damping/excitation times were much shorter than a Hubble time. For example, in a non-rotating halo with a plausible value of the velocity anisotropy, 2hv i=hv i = 1:5 (recall from Section 3.2 that cosmological simulations r t 2 2 yield a typical value of 1.4 for this parameter), the damping time = 10 y even when the 1 8 ellipticity of the halo potential is as small as  = 0:2.  Thus our numerical results show that dynamical friction between a warped galaxy disk and its surrounding dark matter halo usually leads to signi cant evolution of the warp over a Hubble time. The rate of evolution depends strongly on the velocity anisotropy and net rotation of the halo stars and the ellipticity of the halo potential. Dynamical friction can either damp or excite a warp: warps can have either positive or negative energy, and friction can either add energy to or remove energy from the warp. If the disk lies in the equatorial plane of an axisymmetric halo, then the energy of an m = 1 warp is generally positive if the halo is oblate or vertically sti , and negative if the halo is prolate or vertically soft. A non-rotating iddf halo always removes energy from the warp. Halos with predominantly radial orbits remove energy more rapidly than halos with an isotropic velocity distribution, while in the unlikely case that the halo orbits are predominantly circular, dynamical friction can add energy to the warp. The e ect of halo rotation is to add energy to the warp if the mean rotation rate is in the same direction as the precession of the warp, and to remove energy otherwise. Our assumption that the halo is axisymmetric is a ma jor oversimpli cation, since cosmological simulations of galaxy formation typically produce halos that have highly triaxial shapes. The formation of gaseous disks in triaxial potentials has been investigated by analytic arguments and numerical simulations (Steiman-Cameron & Durisen 1984, Habe & Ikeuchi 1985, Pearce & Thomas 1991, Thomas et al. 1994). Disks can form in the principal plane normal to the ma jor or the minor axis (orbits in the principal plane normal to the intermediate axis are unstable). As a rst approximation, we may treat the halo as vertically sti if the disk is normal to the minor axis and vertically soft if the disk is normal to the ma jor axis. The principal plane in which accreted gas will settle depends on its initial angular momentum; simple analytic arguments { 27 { based on perturbation theory imply that the probability that gas with randomly directed angular momentum will settle into the principal plane normal to the ma jor axis is 2I = , where crit 2 2 2 b a sin I = : (81) crit 2 2 c a Taking typical halo axis ratios of c=a = 0:45, b=a = 0:7 (Dubinski & Carlberg 1991, Warren et al. 1992), we nd that the probability that disks lie in the principal plane normal to the ma jor axis is roughly 0.3; this is an upper limit since in practice the angular momentum vector of the infalling gas is strongly correlated with that of the halo material (e.g. Katz & Gunn 1991) For the minority of galaxies in which the disk forms in the principal plane normal to the halo ma jor axis, our results show that the disk is likely to be strongly unstable to warping; the outcome of this instability is presumably the settling of the disk into the principal plane normal to the minor 9 axis . It is possible that some warped galaxies are a manifestation of this instability. Rotating halos can also excite warps. Warps in vertically sti halos generally have retrograde pattern speeds (eq. 17) and positive energy, so that excitation requires that the halo and the disk rotate in opposite directions; as we have mentioned, cosmological simulations show that the angular momentum vectors of halos and their embedded disks are generally similar (although polar rings [Casertano et al. 1991] and counter-rotating disks [Rubin et al. 1992] o er counter-examples). Warps can also be excited if the halo orbits are predominantly circular, but simulations of galaxy formation suggest that halo orbits are radial rather than circular. Thus excitation by either of these mechanisms is unlikely to be the ma jor cause of warps either. A general diculty with explaining warps by excitation from dynamical friction is that the excitation timescales are short compared to a Hubble time, and there is no obvious nonlinear process to limit the amplitude of the warps at their observed values. In the ma jority of galaxies, for which dynamical friction damps rather than excites warps, the damping time is short compared to the Hubble time. In the plausible case that halos have predominantly radial orbits, the damping time is often shorter than even the precession time 2=j j (Figure 2). p These results bring into sharp focus the problem of exciting warps. Ostriker & Binney (1989) have argued that normal galaxy formation processes will not excite the Laplace mode; our calculations show that even if the mode were excited during galaxy formation, it would by now have been damped away. Thus we require an excitation mechanism that operates throughout the life of the galaxy. One plausible mechanism is twisting of the halo by cosmic infall (Ostriker & Binney 1989). Both analytic arguments and N-body simulations show that the direction of the angular momentum vector of a haloand hence the orientation of its principal planesis reoriented by 9 A corollary is that massive polar rings should be unstable. { 28 { infalling material with each doubling of the halo age (see Binney 1992 for a review). This steadily changing orientation is communicated to the inner halo and disk by gravitational torques; the warp arises because it is necessary to transport o -axis angular momentum between the inner and outer disk. In this model the warp amplitude is determined by the present rate of twisting rather than the past history of excitation and damping. Our results can be applied to warps with other azimuthal wavenumbers; although we have done no calculations we can make some comments. Sparke (1994) has argued that axisymmetric (m = 0) `bowl-shaped' modes may be present in spiral galaxies and polar rings. The analysis of Section 2.2 shows that all m = 0 warps have positive energy; hence they will be damped in an iddf halo. Symmetry arguments show that halo rotation will not tend to excite axisymmetric modes either. We expect the damping from dynamical friction to be rapid, since the frequency of the m = 0 mode is usually much larger than the pattern speed of the Lagrange mode. Thus it is unlikely that m = 0 modes are excited by dynamical friction; and if they are excited by other mechanisms, they will be rapidly damped. For higher azimuthal wavenumbers, m  2, there is a continuum of bending modes in realistic disks, so such modes will disperse even without damping from the halo. We close this section with a summary of the approximations we have made: 1. We have used linear perturbation theory (in the warp amplitude), which is inadequate: observed warps may have amplitudes 10 , which represents only the di erence in  >  inclination between the inner disk and the outer edge. In perturbation theory the tilted disk tends to align with the equatorial plane of the halo; in reality a tilted disk with mass comparable to the inner halo has much more angular momentum than the halo, so that the inner halo tends to align with the disk rather than the disk with the halo (then over a longer timescale the outer halo continues to exert signi cant frictional force and eventually brings all three components to a common alignment). 2. In calculating the orbits of halo stars, we have assumed that the unperturbed potential of the halo plus disk is spherical, whereas in fact the halo potential is probably triaxial and the disk provides an additional attened component. With a more realistic potential the friction would probably be increased, since the halo would generally be attened towards the disk plane so its interaction with the warp would be stronger. An inconsistency in our calculations is that we assume the halo is axisymmetric but non-spherical when computing the warp energy and pattern speed. 3. We have approximated the halo potential as that of a singular isothermal sphere. This approximation leads us to underestimate the damping/excitation for warps with small pattern speeds, since the absence of a core implies that fewer halo stars are close to the (l ; l ) = (1; 2) resonance that dominates the friction at small pattern speeds. 1 2 4. We have neglected the self-gravity of the halo response in order to simplify the friction { 29 { calculations. Including the self-gravity of the halo response should generally increase the dynamical friction (for satellites, the inclusion of the response self-gravity decreases the friction [Hernquist & Weinberg 1989], but this re ects mainly the diculty of treating the motion of the center of mass consistently without self-gravity), and therefore strengthen our conclusion that the damping times are short. 5. We have treated the disk as razor-thin, with no internal structure. The dispersion relation for bending or corrugation waves in stellar systems with slab geometry (uniform in x and y , self-gravitating, and symmetric about the z = 0 plane) has been investigated by Araki (1985), Weinberg (1991), and Toomre (1966, 1994). The bending wave is damped by near-resonant stars; if the horizontal variation of the bending wave is / exp[i(k  x ! t)], the resonant stars are those with k  v = n (E ) + ! , where n is an integer and  (E ) is the z z vertical oscillation frequency as a function of vertical energy (this approximation neglects the curvature of the epicyclic orbits in a real galaxy disk, but this should not invalidate our analysis since the epicycle frequency is small compared to  ). If the typical horizontal velocity dispersion is  , the criterion for signi cant dampingplenty of resonant starsis k= (0) > 1 + j! j= (0). The term j! j= (0) is usually negligible for Laplace modes, which have small pattern speeds. Taking  = 2=k = 10 kpc,  = 40km s ,  (0) = 3  10 s 1 15 1 (corresponding to a local density of 0:15M pc ), we nd k= (0) = 0:3; which is probably 3 too small for signi cant damping in a typical warp. Many of the concerns raised by these assumptions can be addressed by N-body simulations, which complement the analytic approach described here. Dubinski & Kuijken (1994) have recently examined the evolution of tilted disks (both rigid and N-body) embedded in N-body halos. They nd that the disk rapidly aligns with the inner halo, and the damping times that they observe are even shorter than the ones we calculate (probably because the disk contains most of the angular momentum, so that the halo aligns with the disk rather than vice versa). Our results suggest other worthwhile N-body experiments. It would be interesting to con rm whether warps can be excited in rotating or vertically soft halos, to test our prediction that the damping is strongest in halos with predominantly radial orbits, and to investigate the fate of massive polar rings. Our conclusions are summarized brie y in the abstract. We acknowledge helpful discussions with James Binney, Omer Blaes, Ray Carlberg, John Dubinski, Peter Goldreich, Linda Sparke, and Alar Toomre. This research was supported by NSERC. A The Dimensionless Energy Loss Rate We wish to evaluate the rate of energy transfer to the disk given by equation (62). { 30 { An orbit in a spherical potential can be described by the canonical actions (I ; I ; I ) = (I ; J; J ), and their conjugate angles (w ; w ; w ); I is the radial action, and 1 2 3 r z 1 2 3 r J and J are the magnitude of the total angular momentum and its component along e which z z we take to be normal to the unperturbed disk. The conjugate angles vary from 0 to 2 , with frequencies given by Hamilton's equations (51). Following Tremaine & Weinberg (1984) and Weinberg (1986), we expand the perturbing potential from the warped disk in both spherical harmonics and an action-angle Fourier series (cf. eq. 55) , 10 U (r; t) = U (r)Y (; 0)e w ln ln in( t) p 1 l X X l=1 n=l 8 9 < = 1 1 X X = Re (I)e (A1) i(lw! t) l : ; l l l 1 2 3 l =0 l ;l =1 3 1 2 where l = jnj, ! = l , U = (1) U , and the radial functions U are derived in Appendix 3 3 p l;n ln l ln  n B. The amplitudes in the two lines of equation (A1) are related by   1 X 2 l 1 (I ; I ; I ) = V ( )W (I ; I ); (A2) l l l 1 2 3 ll l 1 2 1 2 3 2 3 ll l 2 3 l=1 1 +  l 0 3 where V ( ) = r ( )Y (  ; 0)i ; (A3) ll l 2 3 l l 2 3 ll 2 2 l 1 l l 3 2 l 1 Z  1 W (I ; I ) = dw cos[l w l ( w )]U (r); (A4) ll l 2 3 1 2 1 1 1 2 2 ll 3  0 here is the angle in the orbital plane measured from the ascending node, and explicit formulae for the angles w and w are given in Tremaine & Weinberg (1984) and below (eqs. A10 and A11). 1 2 Here cos = J =J is the cosine of the orbital inclination, and r ( ) are rotation matrices which z l l 2 3 l satisfy the orthogonality condition (cf. Gottfried 1979, section 34.5) Z  l l 0 2 d sin r ( )r ( ) = mn mn  0 : ll (A5) 0 2l + 1 It can be shown that = . l l  For the logarithmic potential given in equation (68) it is convenient to de ne a dimensionless radius and angular momentum (Weinberg 1986) r 2 J E =v c ~r = e ;  = ; (A6) r s J (E ) c 10 i!t in t p When n = 0 the time-dependence must be written as e rather than e . For simplicity, we shall not explicitly include this special case in our formulae. { 31 { where J (E ) = J e is the maximum angular momentum possible for a given energy (a c s 2 (E =v 1=2) c circular orbit) and J = r v . In these variables the orbital frequencies can be written in the form s s c (E ; J )  e  (); (A7) i i E =v 2 c where the new frequencies are given by  = (v =r )  , e i c s i  I  1 1 d~r e  () = 1 ; (A8) 2 (2 ln ~r  =e~r ) 2 2 1=2 I e  () 1 d~r e  () = 2 ; (A9) 1=2 2 2 2 1=2 2e ~r (2 ln ~r  =e~r ) and  = 0. The integrations are taken over a complete orbit with turning points determined by 3 the roots of the denominators. 11 The angles are de ned as a function of the dimensionless radius by Z ~r d~r w (~r) =  () ; (A10) e 1 1 ~r p (2 ln ~r  =e~r ) 2 2 1=2  ()  d~r 2 Z ~r w = w : (A11) 2 1  () 1 e ~r (2 ln ~r  =e~r ) ~r p 1=2 2 2 2 1=2 We divide the directional derivative of the df in equation (70) into two parts, even and odd in cos , @ F @ F @ F @ F       l  = l  + l + l = F (E ; J ) + F (E ; J ) cos ; (A12) 2 3 1 2 @ I @E @ J @ J J;J z E ;J E ;J z z where   F (E ; J ) = F J e + [2( 1)l + l ] ; (A13) 1 0 2 3 2  J 2( 1) E = 2 l  1 and   F (E ; J ) = F J e l : (A14) 2 0 2 J  2 2( 1) E = 2 (2 3) l  Here we have used = @E =@ I. Using dI dI = J dJ d(cos ), the integral over in equation (62) 2 3 due to the rst term, F (E ; J ), is proportional to 1 Z  0 d sin j (I)j = jY ( ; 0)j jW (I ; I )j ; (A15) l 2 ll 2 2 ll l 2 3 1 2 2 1 2 l 1 2 (2l + 1)(1 +  ) l 0 3 l=1 1 X 8 11 1=2 Weinberg sometimes uses dimensionless units r =  = 1; in these units v = 2 . s c { 32 { which follows from equations (A2) and (A5). The integral arising from the second term, F (E ; J ) cos , is proportional to 2 Z  0 d sin cos j (I)j = W (I ; I )W (I ; I )Y (  ; 0)Y (  ; 0) l 2 ll l 2 3 1 2 1 2 ll 0 l l l 2 3 2 2 0 l l 2 2 2 l 1  l 1 1  1 4 1 X (1 +  ) l 0 3 0 l;l =1 Z  l l 0  d sin cos r ( )r ( ): (A16) l l l l 2 3 2 3 0 0 The last integral vanishes unless l = l  1 (cf. Gottfried 1979); we show in Appendix B that the potential contains only terms odd in l + l so this integral will vanish. Consequently, using (A15) 3 in (62) we nd X X 1 4 32 ! Z _ l 1 2 l 1 2 E = jY (  ; 0)j dI J dJ jW j F (E ; J ) (l  ! ): (A17) w 2 ll 2 2 1 1 l ll l 2 3 l l=1 (2l + 1)(1 +  ) l 0 3 The resonance constraint can be written as a delta function in energy,  [E E ()] v l   l c 2    (l  ! ) = =  [E E ()] (A18) l l j@ l  =@Ej j! j !  l l where   2 l   () E () = v ln (A19) l c ! l is the resonant energy associated with a particular  and  is the unit step function. We replace dI by dE = (E ; J ); then using equations (71), (A6), (A7) and (A13) we nd 1 1 J dJ F (E ; J ) 2F J  1 0 s 2 2( 1) (E ; J ) v e  () e 1 1 c 2 = f ()d (A20) where f () = l   () e [( 1)l + l ]: (A21) e 2 3 2 1=2 1 and we have replaced the frequencies  by their dimensionless counterparts  . After integration e i i over energy, the integral in equation (A17) becomes Z l 1 2 ! dI J dJ jW j F (E ; J ) (l  ! ) = l l 1 1 ll l 2 3 2 2 2 2( 1) 1 Z   2F J G M  l   e 0 s d f l 1 2 sgn(! ) d f ()j W ()j  ; (A22) 2 e R d l ll l 2 3 0 1 e  () e ! l where the dimensionless transform is (cf. eq. A4) f l 1 e Z ~r a 1 dw 1   ~rl   e e W (; ! ) = d~r cos[l w l ( w )] U ; (A23) ll l 2 3 l 1 1 2 2 ll 3  d r e ~r p e ! l e ! = R ! =v , and U (r) = (GM =R ) U (r=R ). l l 3 3 d c ll d d ll d e { 33 { Finally, inserting this result in equation (A17) and using equation (71) we obtain the form we use for numerical computations, 2 GM v d c _ _ e E = E ( ; ; ); (A24) w p R 2 d where and _ e 1 2 l _ 1 e E ( ; ; ) = jY (  ; 0)j E ( ; ; ); (A25) p ll 2 2 ll l 2 3 p 1 1 1 X X X 4 l =0 l ;l =1 l=1 3 1 2 (2l + 1) l 1 _ e p f e l 1 2 sgn( ) 4  l   e E ( ; ; ) = d f ()j W (; l )j  : (A26) ll l 2 3 2 ll l 2 3 p 3 p (1 +  ) e ( ) e  () l 0 3 0 1 e l 3 p e 1 1=2 +1=2 2( 1) Z ! When = 1 and = 0, the df is that of a singular isothermal sphere. In this case, we recover the result of Weinberg (1986; eq. 40), after correcting two minor typographical errors (the factor 2l + 2 in Weinberg's paper should be 2l + 1; and the sum over l should run from 0 to l, not l to 3 l) and converting from torque to energy loss. Note the symmetry which implies that l 1 _ e l 1 _ e E ( ; ; ) = E ( ; ; ); (A27) ll l l;l ;l 2 3 2 3 p p _ e _ e E ( ; ; ) = E ( ; ; ): (A28) p p Thus if the halo is non-rotating ( = 0) and l 6= 0, E is independent of the sign of the pattern 3 w _ speed as one would expect from symmetry. The term in equation (A21) proportional to the halo spin contributes energy to the bending wave if the pattern speed is in the same direction as the halo rotation, and removes energy if the pattern speed and halo rotation are opposite. B Gravitational Potential of a Disk Bending Mode The potential of a thin, warped disk of surface density (R) is Z 0 (R ) 0 U (r; t) = G dR (B1) d jr r (R ; t)j 0 0 where r = R + Z (R ; t)e . Subtracting o the potential of the unperturbed ( at) disk, and z 0 0 0 expanding to rst order in the small displacement Z , we obtain the potential due to the warp Z 0 0 0 @ 1 U (r; t) = G dR (R )Z (R ; t) : (B2) w 0 0 @ z jr r j 0 z =0 In spherical coordinates { 34 { 1 4 r X l <  0 0 = Y (; )Y ( ;  ); (B3) 0 l+1 ln ln jr r j 2l + 1 ln r > so that @ 1 X l  4 r 1 @ Y 0 < in ln 1 = Y (; ) (  ; 0)e : (B4) ln 0 0 l+1 0 2 @ z jr r j 0 2l + 1 R @ cos  z =0 ln r > The values of the spherical harmonics are @ Y 2l + 1 (l n)! (l + n)!! ln 1 (l+n1)=2 (  ; 0) = (1) n + l odd (B5) @ cos  4 (l + n)! (l n 1)!! 2   1=2 s where the double factorial terms are (2n 1)!! = ; (2n)!! = 2 n!: (B6) (2n)! n n 2 n! Terms with l + n even vanish. We now substitute equation (7) for Z (R; t). Then we may write U (r; t) = U (r)Y (; 0)e ; (B7) ln ln X i(n! t) n l;n where 2  l 1 Z 4 G @ Y r lm 1 0 0 < 0  0 U (r) = (  ; 0) dR (R ) [h(R ) + h (R ) ]; (B8) ln mn m;n 2l + 1 @ cos  0 r > 2 l+1 and ! = ! , ! = ! . m m  In this paper we mostly consider disturbances with m = 1. Since n = m and n + l must be odd, only terms with even l will be non-zero. We assume that the disk surface density and warp shape are given by equations (77) and (79). Then equation (B8) can be evaluated as GM 2 @ Y d  Z 1 l x l1 1 < x U (r) = (  ; 0) dxf (x) e (B9) l1 2 l+1 R 2l + 1 @ cos  d 0 x > where Z    1 l l+1 x R (n + l + 1; (1 + )r=R ) < x d d dxf (x) e = C 0 x > l+1 n+l+1 r (1 + )    l r (n l; (1 + )r=R ) d + : (B10) R (1 + ) d nl In this equation the gamma functions are de ned by Z Z z 1 p1 x p1 x (p; z ) = x e dx; (p; z ) = x e dx; (B11) 0 z { 35 { note that If p is a positive integer then (p; z ) + (p; z ) = (p) = (p 1)! (B12) (p; z ) = (p 1)!e : (B13) p1 X k z z For negative integers, n  0, k! k=0 1 where E (z ) is an exponential integral. n (n; z ) = E (z ); (B14) n+1 n z C The Contribution of Near-Radial Orbits to Dynamical Friction In this Appendix we elucidate the reason for the apparent divergence in the damping/excitation rate in halos with predominantly radial orbits. Suppose that a torque  (t) is exerted on the star in a nearly radial orbit. The resulting change in the angular momentum vector may be written where  =  (t)dt. The change in J = jJj is then R J = ; (C1) J    (J  ) 2 3 " # 2 2 J =  +  + O( ): (C2) J 2J 2J 3 As J ! 0, in general  approaches a xed, non-zero value. Thus as J ! 0 J !  cos + ; (C3) 2 2 2   sin 2J 3 where is the angle between  and J (plus terms that are O( ), which we neglect). The O() term can be dropped, since it averages to zero in a spherically symmetric distribution of stars, and in any case does not diverge as J ! 0. In a spherical distribution, the average of sin is , so 2 2 1 the average change is hJ i =   =J . The corresponding energy change is 4 1 2 2 @H 0 hE i = hJ i = hJ i; (C4) @ J 2 which also diverges as J , since is generally constant and non-zero as J ! 0. 2 1 The number of stars in a small interval of energy and angular momentum is dN = 16 F (E ; J )dE J dJ= . Now assume that the df has the form (70), that is 1 3 F (E ; J ) = J g (E ). Since is constant as J ! 0, the total energy change of the 1 2( 1) { 36 { stars of a given energy on near-radial orbits is proportional to J hJ idJ =   J dJ , 4 R R 2 1 2 2 2 2 1 which diverges if  . 2 1 Thus we have explained the apparent divergence in dynamical friction for dfs with  . 2 1 Our arguments also show that the divergence is an artifact of the expansion procedure: the exact change in angular momentum is J = (J + 2J   +   ) J; (C5) exact 2 2 2 1=2 and the total energy or angular momentum change from near-radial orbits (J < J , say) is max proportional to Z J max 2 1 J hJ idJ: (C6) exact 1 2 0 When > this integral is O( ) (after averaging over orientations assuming hcos i = 0); this 2 is the usual case in which the frictional energy or angular momentum change is second-order in the perturbation strength . When < the integral is O( ); thus the frictional force is of a 1 2 +1 lower order in the perturbation strength than  , but is not divergent. 2 2 REFERENCES Araki, S. 1985, unpublished Ph.D. thesis, MIT Barnes, J., & Efstathiou, G. 1987, ApJ 319, 575 Bertin, G., & Mark, J. W.-K. 1980, A&A 88, 289 Binney, J. 1992, ARA&A 30, 51 Bosma, A. 1991, in Casertano et al. (1991), 181 Casertano, S., Sackett, P. D., & Briggs, F. H., eds. 1991, Warped Disks and Inclined Rings around Galaxies, (Cambridge: Cambridge University Press) Dekel, A., & Shlosman, I. 1983, in Internal Kinematics and Dynamics of Galaxies, IAU Symposium 100, ed. E. Athanassoula (Dordrecht: Reidel), 187 Dubinski, J. 1992, ApJ 401, 441 Dubinski, J. 1994, preprint Dubinski, J., & Carlberg, R. G. 1991, ApJ 378, 496 Dubinski, J., and Kuijken, K. 1994, preprint Evans, N. W. 1994, MNRAS 267, 333 { 37 { Gerhard, O. E. 1991, MNRAS 250, 812 Goodman, J. 1988, ApJ 329, 612 Gottfried, K. 1979, Quantum Mechanics, London: W.A. Benjamin, Inc. Habe, A., & Ikeuchi, S. 1985, ApJ 289, 540 Hernquist, L., & Weinberg, M. D. 1989, MNRAS 238, 407 Hunter, C., & Toomre, A. 1969, ApJ 155, 747 (HT) Ichimaru, S. 1973, Basic Principles of Plasma Physics, Reading: W.A. Benjamin, Inc. Katz, N., & Gunn, J. E. 1991, ApJ 377, 365 Kuijken, K. 1991, ApJ 376, 467 Lynden-Bell, D. 1965, MNRAS 129, 299 Lynden-Bell, D., & Kalna js, A. J. 1972, MNRAS 157, 1 Lynden-Bell, D., & Ostriker, J. P. 1967, MNRAS 136, 293 Ostriker, E. C., & Binney, J. J. 1989, MNRAS 237, 785 Palmer, P. L., & Papaloizou, J. 1987, MNRAS 224, 1043 Pearce, F. R., & Thomas, P. A. 1991, MNRAS 248, 699 Rubin, V. C., Graham, J. A., & Kenney, J.D.P. 1992, ApJ 394, L9 Sanchez-Saavedra, M. L., Battaner, E., & Florido, E. 1990, MNRAS 246, 458 Sparke, L. S. 1984, ApJ 280, 117 Sparke, L. S. 1994, preprint Sparke, L. S., & Casertano, S. 1988, MNRAS 234, 873 (SC) Steiman-Cameron, T. Y., & Durisen, R. H. 1984, ApJ 276, 101 Thomas, P. A., Vine, S., & Pearce, F. R. 1991, MNRAS 268, 253 Toomre, A. 1966, Notes on the 1966 Summer Study Program in Geophysical Fluid Dynamics, Woods Hole Oceanographic Institution, 111 Toomre, A. 1969, ApJ 158, 899 Toomre, A. 1983, in Internal Kinematics and Dynamics of Galaxies, IAU Symposium 100, ed. E. Athanassoula (Dordrecht: Reidel), 177 { 38 { Toomre, A. 1994, private communication Tremaine, S., & Weinberg, M. D. 1984, MNRAS 209, 729 Warren, M. S., Quinn, P. J., Salmon, J. K., & Zurek, W. H. 1992, ApJ 399, 405 Weinberg, M. D. 1985, MNRAS 213, 451 Weinberg, M. D. 1986, ApJ 300, 93 Weinberg, M. D. 1991, ApJ 373, 391 White, S.D.M. 1984, ApJ 286, 38 Whitham, G. B. 1974, Linear and Nonlinear Waves (New York: Wiley) Zurek, W. H., Quinn, P. J., & Salmon, J. K. 1988, ApJ 330, 519 This preprint was prepared with the AAS L T X macros v3.0. A E { 39 { Authors' Addresses Robert W. Nelson and Scott Tremaine CITA, University of Toronto 60 St. George Street Toronto, ON, M5S 1A7 CANADA Email: [email protected], [email protected]
0812.1756
2
0812
2009-05-21T09:22:33
Analysis of non-Gaussian CMB maps based on the N-pdf. Application to WMAP data
[ "astro-ph" ]
We present a new method based on the N-point probability distribution (pdf) to study non-Gaussianity in cosmic microwave background (CMB) maps. Likelihood and Bayesian estimation are applied to a local non-linear perturbed model up to third order, characterized by a linear term which is described by a Gaussian N-pdf, and a second and third order terms which are proportional to the square and the cube of the linear one. We also explore a set of model selection techniques (the Akaike and the Bayesian Information Criteria, the minimum description length, the Bayesian Evidence and the Generalized Likelihood Ratio Test) and their application to decide whether a given data set is better described by the proposed local non-Gaussian model, rather than by the standard Gaussian temperature distribution. As an application, we consider the analysis of the WMAP 5-year data at a resolution of around 2 degrees. At this angular scale (the Sachs-Wolfe regime), the non-Gaussian description proposed in this work defaults (under certain conditions) to an approximative local form of the weak non-linear coupling inflationary model (e.g. Komatsu & Spergel 2001) previously addressed in the literature. For this particular case, we obtain an estimation for the non-linear coupling parameter of -94 < F_nl < 154 at 95% CL. Equally, model selection criteria also indicate that the Gaussian hypothesis is favored against the particular local non-Gaussian model proposed in this work. This result is in agreement with previous findings obtained for equivalent non-Gaussian models and with different non-Gaussian estimators. However, our estimator based on the N-pdf is more efficient than previous estimators and, therefore, provides tighter constraints on the coupling parameter at degree angular resolution.
astro-ph
astro-ph
Mon. Not. R. Astron. Soc. 000, 1 -- 13 (1998) Printed 25 October 2018 (MN LATEX style file v2.2) Analysis of non-Gaussian CMB maps based on the N-pdf. Application to WMAP data P. Vielva1, J. L. Sanz1,2 1Instituto de F´ısica de Cantabria (CSIC - Univ. de Cantabria), Avda. Los Castros s/n, 39005 - Santander, Spain 2CNR Istituto de Scienza e Tecnologie dell'Informazione, 56124, Pisa, Italy E-mails : [email protected], [email protected] Accepted ???. Received ???; in original form 25 October 2018 ABSTRACT We present a new method based on the N-point probability distribution (pdf) to study non-Gaussianity in cosmic microwave background (CMB) maps. Likelihood and Bayesian estimation are applied to a local non-linear perturbed model up to third order, characterized by a linear term which is described by a Gaussian N-pdf, and a second and third order terms which are proportional to the square and the cube of the linear one. We also explore a set of model selection techniques (the Akaike and the Bayesian Information Criteria, the minimum description length, the Bayesian Evidence and the Generalized Likelihood Ratio Test) and their application to decide whether a given data set is better described by the proposed local non-Gaussian model, rather than by the standard Gaussian temperature distribution. As an application, we consider the analysis of the WMAP 5-year data at a resolution of ≈ 2◦. At this angular scale (the Sachs-Wolfe regime), the non-Gaussian description proposed in this work defaults (under certain conditions) to an approximative local form of the weak non- linear coupling inflationary model (e.g. Komatsu et al. 2001) previously addressed in the literature. For this particular case, we obtain an estimation for the non-linear coupling parameter of −94 < fNL < 154 at 95% CL. Equally, model selection criteria also indicate that the Gaussian hypothesis is favored against the particular local non- Gaussian model proposed in this work. This result is in agreement with previous findings obtained for equivalent non-Gaussian models and with different non-Gaussian estimators. However, our estimator based on the N-pdf is more efficient than previous estimators and, therefore, provides tighter constraints on the coupling parameter at degree angular resolution. Key words: cosmology: observations -- cosmology: cosmic microwave background -- methods: data analysis -- methods: statistical 9 0 0 2 y a M 1 2 ] h p - o r t s a [ 2 v 6 5 7 1 . 2 1 8 0 : v i X r a 1 INTRODUCTION The Cosmic Microwave Background (CMB) fluctuations, a relic radiation originated around 400,000 years after the Big- Bang, is one of the most outstanding sources for understand- ing the evolution and energy/matter content of the Uni- verse. The large amount of high quality data provided by recent CMB experiments and other complementary astro- nomical observations, have provided with a consistent pic- ture for a flat Universe filled with cold dark matter (CDM) and dark energy in the form of a cosmological constant (Λ), plus the standard baryonic and electromagnetic com- ponents: the concordance model (e.g. Komatsu et al. 2008). However, beyond the strength of the CMB measurements to put constraints on the cosmological parameters, like the ones already provided by the NASA Wilkinson Microwave Anisotropy Probe (WMAP, Hinshaw et al. 2008) and the ones expected from the incoming ESA Planck satellite, the CMB is a unique tool to probe fundamental principles and assumptions of the so-called standard model. In particular, the application of sophisticated statistical analysis to CMB data might help us to understand whether the temperature fluctuations of the primordial radiation are compatible with the fundamental isotropic and Gaussian predictions from the inflationary phase. The basic inflationary scenario relates the CMB fluctuations, as well as the large-scale structure of the Universe, to the Gaussian quantum energy density perturbations present during the early Universe (see for in- stance Liddle & Lyth 2000). The present homogeneity and isotropy of the Universe is compatible with an inflation- ary era in the early universe and this idea is the only way, 2 P. Vielva and J. L. Sanz nowadays, to explain efficiently current observations rang- ing from galaxies to the CMB. In particular, this fundamen- tal hypothesis predicts that the CMB fluctuations follow an isotropic and Gaussian random. In fact, the estimation of the cosmological parameters defining the concordance model is done by assuming these statistical properties. The quality of current CMB data (in particular the ones provided by the WMAP satellite) has allowed for a systematic probe of the statistical properties of the relic radiation. Indeed, the interest of the scientific community in this field has experimented a significant growth, since the analysis of the WMAP data has reported several hints for departure from isotropy and Gaussianity of the CMB temperature distribution. Some of these works are the following. Park (2004) detected a Gaussianity deviation with a genus-based statistic; Vielva et al. (2004) found a significant non-Gaussian signature on the 1-year WMAP data on the kurtosis of the Spherical Mexican Hat (SMHW) wavelet coefficients at scales of around 10 degrees, pointing out a very large cold spot (CS) on the southern hemisphere as a possible source for this non-Gaussianity. A posterior analysis (Cruz et al. 2005) confirmed the anomalous nature of the CS by performing an area-based statistical analysis of the wavelet coefficients. This detection, as well as new ones, were confirmed by analyzing WMAP with various wavelet bases and several statistical estimators (Mukherjee & Wang 2004; Cay´on et al. 2005; McEwen et al. 2005; Cruz et al. 2006; Mart´ınez-Gonz´alez et al. 2006; Cruz et al. 2007a; McEwen et al. 2006; Pietrobon et al. 2008; Wiaux et al. 2008). Isotropy deviations were also reported in different manners: an anomalous alignment of the low multipoles of the CMB (Copi et al. 2004; de Oliveira-Costa et al. 2004; Katz & Weeks 2004; Schwarz et al. 2004; Bielewicz et al. 2005; Land & Magueijo 2006; Freeman et al. 2006; Land & Magueijo 2007); north-south asymmetries of (Eriksen et al. 2004a,b; Hansen et al. 2004a,b; Donoghue & Donoghue 2005; Eriksen et al. 2005a; Bernui et al. 2006, 2007; Eriksen et al. 2007; Rath et al. 2007; Gordon 2007); an anomalous variance value (Monteser´ın et al. 2008); unexpected correlation among the CMB phases (Chiang et al. 2003; Coles et al. 2004; Chiang & Naselsky 2006); unbalanced distribution of the temperature extrema (Tojeiro et al. 2006); and anoma- lous alignment of CMB structures (Wiaux et al. 2006; Vielva et al. 2006, 2007). the CMB fluctuations 2005b; Abramo et al. 2005; Land & Magueijo All the previous analyses can be considered as blind approaches, since null tests were performed to probe the CMB compatibility with the isotropic and Gaussian hy- potheses. Complementary to these works, the reader can also find in the literature several studies where targeted departures from Gaussianity are explored, based on non- standard physical models. In particular, several analy- ses have studied the WMAP data compatibility with anisotropic universes: the Bianchi VIIh model (Jaffe et al. 2006a,b,c; Bridges et al. 2007a, 2008). Recently, some works have proposed non-rotational invariant models like Bohmer & Mota (2008) and Ackerman et al. (2007) (ex- prored by Groeneboom & Eriksen 2008, in the context of WMAP data). Although these non-rotational invariant models are promising and provide us with anisotropic tem- plates that could help to fix some anomalies in WMAP data, more work is still needed to connect these anisotropic patterns of the CMB fluctuations to a satisfactory physi- cal model describing the evolution of the anisotropic field (Himmetoglu et al. 2008a,b). In addition to the previous analyses, several works have studied different hypotheses to explore the anomalous na- ture of the CS; for instance Cruz et al. (2008) explored the CS compatibility with different non-standard models, point- ing out that an explanation in terms of a cosmic texture (as proposed by Cruz et al. 2007b) is much more favored than other alternatives already discussed in the literature, like a very large void in the large scale structure of the Universe or contamination, in the form of the Sunyaev- Zeldovich emission, due to a large and nearby galaxy clus- ter. However, the study of non-standard inflationary mod- els is the problem that has attracted a larger attention. For instance, and for the WMAP case, the non-linear cou- pling parameter fNL that describes the non-linear evolution of the inflationary potential (see e.g. Bartolo et al. 2004, and references therein) has been constrained by several groups: using the angular bispectrum (Komatsu et al. 2003; Creminelli et al. 2006; Spergel et al. 2007; Komatsu et al. 2008); applying the Minkowski functionals (Komatsu et al. 2003; Spergel et al. 2007; Gott et al. 2007; Hikage et al. 2008; Komatsu et al. 2008); and using different statistics based on wavelets (Mukherjee & Wang 2004; Cabella et al. 2005; Curto et al. 2008). Besides a claim for fNL > 0 with a probability greater than 95% (Yadav & Wandelt 2008), there is a general consensus on the WMAP compatibility with the predictions made by the standard inflationary sce- nario at least at 95% confidence level. The current best limits (Curto et al. 2008) are: −8 < fNL < 111 at 95% CL. The present paper is related to non-blind or targeted probes for Gaussianity deviations, and, more specifically, it addresses two common topics arose in these kind of studies: the definition of an optimal estimator for the non-standard model (and, therefore, the optimal estimation of the param- eters that define such a model) and the complementary is- sue of model selection. The former aspect is, usually, quite complex, since, for many physical models and realistic obser- vational limitations (e.g. incomplete sky coverage) there is not a trivial solution and, therefore, non-optimal estimators are usually proposed (which are posteriorly characterized by simulations). The latter issue on model selection is equally complex, since, generally, relies on heuristic principles. Some authors adopt a fully Bayesian view, whereas others adopt asymptotic measurements for the distance between two hy- pothesis, and others prefer to rely on the information that can be obtained from the likelihood itself. In this work we propose to study a non-Gaussian model for the CMB that is a local non-linear expansion of the temperature fluctua- tions. For this model -- that, at large scales, is considered by some authors as an approximation to the non-linear cou- pling parameter fNL -- we are able to build the exact like- lihood on pixel space. To work in pixel space allows one to include easily non-ideal observational conditions, like incom- plete sky coverage and anisotropic noise. We show how, from this likelihood, it is straightforward to obtain an analytical expression for the optimal estimator for the non-Gaussian term. In addition, we also explore different model selection criteria, like the Akaike information criterion (Akaike 1973), the Bayesian information criterion (Schwarz 1978), the min- Analysis of NG CMB maps based on the N-pdf 3 imum description length (Rissanen 2001), the Bayesian evi- dence, and the generalized likelihood ratio test. The paper is organized as follows. In Section 2 we de- scribe the physical model based on the local expansion of the CMB fluctuations and derive the full posterior proba- bility. In Section 3 we address the issue of the parameter estimation (3.1) and model selection (3.2). The methodol- ogy is explored on WMAP-like simulations in Section 4 and it is applied to WMAP 5-year data in Section 5. Finally, conclusions are given in Section 6. model. Indeed, the reason to select the specific model given by equation 1 is twofold. One the one hand, it is an useful parametrization for describing a small departure from Gaus- sianity, that allows us to present a new methodology. On the other hand, it is a model previously addressed by other au- thors (using other estimators), and, therefore, it is easier to make a straightforward comparison among the results. For simplicity, let us transform the Gaussian part (∆T i)G into a zero mean and unity variance random vari- able φi, hence, equation 1 can be rewritten as: 2 THE NON-GAUSSIAN MODEL where: xi = φi + ǫ`φ2 i − 1´ + αǫ2φ3 i , Current observations indicate that CMB temperature fluc- tuations can be well described by random Gaussian fluctu- ations, as predicted by the standard inflationary scenario. However, as it was discussed in the Introduction, these ob- servations still allow for a small departure from the Gaussian distribution, which could reflect the role played by physical processes described by non-standard models for the struc- ture formation. In this paper, we will focus on a parametric non- Gaussian model, that accounts for a small and local (i.e. point-to-point) perturbation of the CMB temperature fluc- tuations, around its intrinsic Gaussian distribution: G −(∆T i)2 G¸ + b (∆T i)3 ∆T i = (∆T i)G + a(∆T i)2 G . (1) (∆T i)G (the linear term) is the Gaussian part, whose N- point probability density function (N-pdf) can be easily de- scribed in terms of the standard inflationary model. The second and third terms on the right-hand side are the quadratic and cubic perturbation terms, respectively. Their corresponding contribution to the observed CMB fluctua- tions (∆T i) is governed by the a and b parameters. Subindex i refers to the direction corresponding to a certain pixeliza- tion on the sphere, and the operator h·i averages over all the pixels defining the sky coverage. Notice that the previ- ous expression does not include instrumental noise. We have adopted this simplification, since we aim to focus the work on large-scale CMB data sets (> 1◦), where (as it is the case for WMAP), the contribution from noise is typically negligible. Let us point out that the particular situation for b ≡ 0 defaults into a well known case that has been al- ready addressed in the literature (e.g. Komatsu et al. 2001; Cay´on et al. 2003; Curto et al. 2007). It describes a local approximation to the weak non-linear coupling inflation- ary model (e..g Komatsu et al. 2001; Liguori et al. 2003) at scales larger than the horizon scale at the recombination time (i.e. above the degree scale). In this context, the a fac- tor is usally related to the non-linear coupling parameter fNL by: a ≡ 3fNL T0 , (2) where T0 = 2, 725 mK is the CMB temperature, and we follow the sign convention in Liddle & Lyth (2000) for the relation between the temperature fluctuations and the gravi- tational potential at the Sachs-Wolfe regime. Let us remark that, of course, this model do not pretend to incorporate all the gravitational (like lensing) and non-gravitational ef- fects, due to the evolution of the initial quadratic potential 1 σ (∆T )G , ǫ ≡ aσ, α ≡ bσ2 ǫ2 , (3) (4) 1 σ x ≡ ∆T , φ ≡ and σ2 ≡(∆T i)2 G¸ is the rms of the CMB fluctuations. Let us remark that, since the proposed non-Gaussian model is a perturbation of the standard Gaussian one, the non-linear parameters have to satisfy: ǫ ≪ 1, α . 1. (5) It is straightforward to show that the normalized Gaussian field φ satisfies: hφii = hφ3 ii = 0, hφ2 ii = 1, hφiφji = ξij , (6) where ξij represents the normalized correlation between pix- els i and j. The N-pdf of the φ = {φ1, φ2, ..., φN} random field (where N refers to the number of pixels on the sphere that are observed) is given by a multivariate Gaussian: p(φ) = 1 (2π)N/2(det ξ)1/2 e− 1 2 φξ−1φ t , (7) where ξ denotes the correlation matrix and operator ·t de- notes standard matrix/vector transpose. Our goal is to compute the N-pdf associated to the non- Gaussian x = {x1, x2, ..., xN} field, as a function of the non- linear coupling parameters based on the full N-pdf for the underlaying Gaussian signal φ. Hence, let us make first the inversion of equation 3 to find the expression of φi as a function of xi: i − 1) + ǫ2[(2 − α)x3 φi = xi − ǫ(x2 Obviously, since previous equation is a local transfor- mation, the Jacobian matrix is diagonal and, therefore, the Jacobian (Z) is given by: i − 2xi] + O(3). (8) Z = det» ∂φi ∂xj -- =Yi „ ∂φi ∂xi« . (9) It is more convenient to work with the log-Jacobian (log Z), which, taking into account equation 8 and expanding the log function up to second order, is given by: log Z = Xi = −2ǫXi log„ ∂φi ∂xi« xi + ǫ2"−2N + (4 − 3α)Xi (10) x2 i# + 0(3). Now, taking into account equation 3 and recalling that φ is a Gaussian field, it is straightforward to prove that the data 4 P. Vielva and J. L. Sanz x satisfy: Pi xi = 0 and 1 the log-Jacobian is given by: N Pi x2 i = 1 + O(2). Therefore, log Z = N ǫ2 (2 − 3α) + O(3), (11) Finally, it is easy to calculate p(xǫ), the N-point pdf of x given the ǫ parameter: p(xǫ) = p(φ = φ(x))Z = p(x0)el(xǫ) (12) where the probability p(x0) is given by: (2π)N/2(det (ξ))1/2 e− 1 p(x0) = 1 2 xξ−1x t , (13) i.e., it is the N-point Gaussian pdf (i.e. p(x0) ≡ p(φ ≡ x)), and the log-likelihood l(xǫ) reads as: l(xǫ) = log Z − 1 2 φξ−1φt − xξ−1xt . It is easy to show that l(xǫ) is given by: l(xǫ) ≡ N [ǫR − ǫ2Q + O(3)], where the functions R and Q are given by: and , R = 1 N xξ−1x2 − It Q = −2 + J + αS, non-Gaussian model in equation 1 can be seen as an ap- proximation at large scales) is really negligible or not with respect to the quadratic contribution. For the former sce- nario (negligibility of the cubic term), the particular value of α becomes an irrelevant issue since, naturally, one will have that S ≪ −2 + J (of course, α should always satisfy the condition given in equation 5). However, for the latter case different results could be obtained, depending on the specific value for α. In the following, we will discuss pa- rameter estimation and model selection assuming α ≡ 0 in equation 17. However, we will explore (analyzing WMAP data and simulations) whether the condition S ≪ −2 + J is naturally satisfied or not. 3 STATISTICAL ANALYSIS In this Section we aim to address two aspects very much linked one to the other: the estimation of the parameter defining the local non-Gaussian model (ǫ, Section 3.1) and the computation of some heuristic rules to decide whether a given data set is better described by the local non-Gaussian model rather than by the standard Gaussian one (Sec- tion 3.2). (14) (15) (16) (17) being I the unity vector of dimension N and 3.1 Parameter estimation J = and + 1 N 2xξ−1x`x2 − I´t N Iξ−1It − S = 3 1 2 x2 − I ξ−1x2 − Itff xξ−1x3tff . 1 3 (18) (19) Let us remark that R and Q could be seen as a kind of generalized third-order and fourth-order moments, respec- tively (or, equivalently, in terms of the spherical harmonic coefficients, to the bispectrum and the trispectrum). As an example, for the particular case of uncorrelated data (i.e., ξ ≡ δ), it is straightforward to show that R = k3 and N Pi xn Q = −5/2 + (3α − 2)k2 + (5/2 − α)k4, with kn = 1 i . The N-pdf p(xǫ) given by equation 12 contains all the required information, on the one hand, to estimate the non- linear coupling parameter ǫ and, on the other hand, to per- form a model selection (Gaussian vs. non-Gaussian). These two aspects will be studied in next Section. Notice that the parameter α (or, equivalently, b) can- not be estimated in this framework: it just appears as an arbitrary constant in the definition of Q (equation 17): it just controls the relevance of S in Q. If one were inter- ested in obtaining a posterior probability of the data x given both non-linear parameters (ǫ and α), then it would be necessary to expand the local non-Gaussian model be- yond 0(3). However, we would always find an expression in which the non-linear parameter controlling the highest order in the expansion, acts as an arbitrary constant. For that rea- son, we keep terms up to 0(3). There is a discussion within the field (e.g., Okamoto & Hu 2002; Kogo & Komatsu 2006; Babich 2005; Creminelli et al. 2007) about to what extend the cubic term in the description of the weak non-linear coupling inflationary model (for which, we recall, the local The description of the full N-pdf for the non-Gaussian model proposed in Section 2 allows one to obtain an optimal esti- mation of the non-linear coupling parameter ǫ. Let us recall that an optimal estimation of the ǫ pa- rameter is possible, since it would be derived from the full pdf for the non-Gaussian model, p(xǫ). In other words, we could obtain an unbiased and minimum variance estimator. This is possible, precisely, for the specific selection of the local non-Gaussian model in equation 3: for other physical non-Gaussian models it is not always trivial to obtain a full description of the posterior probability of the data given the parameters (at least, under realistic observational conditions like incomplete sky coverage) and shortcuts have to be taken by defining pseudo-optimal estimators that are, afterwards, validated with simulations. 3.1.1 Parameter estimation from the log-likelihood We shall define the optimal estimator for the non-linear pa- rameter as the value, ǫ, that maximizes the probability of x given ǫ. From equation 12, it is obvious that maximizing this probability is equivalent to maximize l(xǫ) in equation 15. By derivation one obtains: ǫ = R 2Q . (20) We can also estimate the error associated to ǫ from the Fisher matrix Fǫ ≡ − d2l dǫ2 = 2N Q: σǫ = F −1/2 ǫ = (2N Q)−1/2. (21) Notice that the error on the estimation of the parameter ǫ is constant, up to the order considered in equation 3. 3.1.2 Bayesian parameter estimation Within the Bayesian framework we can include any a priori information that we might have in relation to the ǫ parame- ter. In particular, following Bayes' theorem, the probability of ǫ given the data x read as: p(ǫx) ∝ p(xǫ)p(ǫ), (22) where p(ǫ) is the prior probability function for the parame- ter ǫ. Of course, for the case of the local non-Gaussian model proposed in this work, there is not a clear physical motiva- tion to choose a particular prior. Let us however explore, as an exercise, two simple sce- narios, which could be useful for more general purposes. We consider first a uniform prior given by: 1 ǫM−ǫm 0 if ǫ ∈ [ǫm, ǫM] otherwise , (23) p(ǫ)8< : where, obviously, the range allowed to ǫ is such that ǫ ≪ 1 for any ǫ ∈ [ǫm, ǫM]. For this particular case, it is trivial to show that the Bayesian estimation for the non-linear cou- pling parameter (¯ǫ) is equivalent to the one obtained via the maximum-likelihood estimation (i.e, ¯ǫ ≡ ǫ) if ǫ ∈ [ǫm, ǫM]. The second case we want to address corresponds to a Gaussian prior p(ǫ), described by a most probable value ǫ∗ and a dispersion σ∗: p(ǫ) = (ǫ−ǫ∗ )2 2σ∗ 2 . − e 1 √2πσ∗ (24) By deriving the posterior probability, it is trivial to obtain the Bayesian estimation for the non-linear coupling param- eter (¯ǫ): ¯ǫ = N σ∗ 2N σ∗ 2R + ǫ∗ 2Q + 1 . (25) For the particular case of σ∗ → 0, i.e. a very strong prior for ǫ, peaked around ǫ∗, one trivially obtain ¯ǫ ≡ ǫ∗, that is, the prior dominates Bayesian estimation, leading to a most probable value for ǫ equal to the maximum value for the prior. Also trivially one finds that, for a non-informative scenario (i.e., σ∗ → ∞), ¯ǫ ≡ ǫ = R/2Q. 3.2 Model selection In this subsection we aim to calculate under which condi- tions (according to different model selection criteria) a given observation x is better described by a local non-Gaussian model as the one described by equation 3 with ǫ > 0 (hereinafter H1) rather than by a Gaussian random field described just in term of the N-point correlation function ξ (hereinafter H0). Some of the model selection approaches inves- tigated in this paper have been previously applied to different astronomical/cosmological problems For in- stance, (Szydlowski & Godlowski 2006; Szydlowski et al 2006; Borowiec et al. 2006) applied the Akaike and the Bayesian information criteria (AIC and BIC, recpectively) to study whether astronomical data sets favored simplest models for the accelerating universe against more complex ones. This issue was also addressed by Davis et al. (2007) by Analysis of NG CMB maps based on the N-pdf 5 analyzing the ESSENCE supernova survey data, and appliy- ing Bayesian evidence (BE) in addition to the AIC and the BIC approaches, to study. These three model selection crite- ria (AIC, BIC and BE) were also applied to study the impact of non-standard physical models on the Friedmann equa- tions (Szydlowski et al. 2008). Liddle (2004) used the AIC and the BIC techniques to study, on the one hand, whether WMAP 1-year data preferred a spatially flat cosmology ver- sus a closed one, and, on the other hand the significance of the running spectral index detected on this WMAP data release. In a posterior work (Liddle 2007), BE was added to the AIC and the BIC approaches to study the suitability of different cosmological models to the WMAP 3-year data. Bayesian evidence (BE) is, probably, the model selec- tion criterion that has attracted a greater interest from cosmologists during the past years. In addition to the works mentioned above (where it was compared with other model selection criteria), it has been also applied to several problems where competing cosmological mod- els were explored. Some of these applications are the fol- lowing: Mukherjee et al. (2006) followed a BE approach to study cosmological models with different matter power spectra and dark energy evolution models; Liddle et al. (2006) studied different dark energy evolving scenar- ios; Bridges et al. (2006, 2007b) performed a model selec- tion on the matter power spectrum from the BE analy- sis of the WMAP; Bridges et al. (2007a, 2008) followed a similar approach for analyzing the WMAP compatibility with anisotropic Bianchi VIIh models; Cruz et al. (2007b, 2008) used it to decide whether the WMAP cold spot was compatible or not with predictions from non-standard models like the cosmological defects; Mukherjee & Liddle (2008) studied the Planck ability to discriminate between several re-ionization models; the BE criterion was also applied (Carvalho et al. 2008; Feroz et al. 2008b) to the problem of compact source detection on microwave data; and more recently, Feroz et al. (2008a) investigated differ- ent properties of the constrained minimal supersymmetric model (mSUGRA), using WMAP data. 3.2.1 The Akaike information criterion (AIC) The Akaike information criterion (AIC, Akaike 1973) pro- vides with a selection index to decide among competing hy- pothesis, being the model associated to the lowest index the most favored one. The Akaike index corresponding to a given model or hypotheses Hi defined by p parameters and with a maximum value for the log-likelihood of l is given by: AIC(Hi) = 2"p − l" . (26) From equation 15, one can find that AIC(H1) = 2"1 − N R2 4Q", whereas, trivially, AIC(H0) = 0. Therefore, according to the AIC, the decision rule reads as: H0 H1 AIC :8>< >: if R2 Q 6 4 N if R2 Q > 4 N (27) where ν is an arbitrary value indicating the strength in choosing H1 instead of H0. Therefore, according to the GLRT, the decision rule reads as: H0 H1 if R2 Q 6 4 N ν (33) if R2 Q > 4 N ν GLRT :8>< >: (28) Notice that the case ν ≡ 1 provides the same decision rule as the AIC (equation 27), and that ν ≡ ln√N corresponds to the BIC case (equation 29). 6 P. Vielva and J. L. Sanz 3.2.2 Bayesian information criterium (BIC) This asymptotic bayesian criterion introduced by Schwarz (1978) is prior independent. It is based on the BIC function, that provides a measurement of the goodness-of-fit of the model to the data, taking into account the number of pa- rameters defining the model as well as the amount of data (N ): BIC(Hi) ="−2l + p ln N" , where p is the number of parameters defining the data and l is the maximum value for the log-likelihood. As for AIC, BIC provides a ranging index for competing hypothesis, where the one with the lower BIC value is the most favored one. From equation 15, one can find that (for our specific prob- 2Q + ln (N ) and BIC(H0) = 0. There- lem) BIC(H1) = −N R2 fore, according to the BIC, the decision rule reads as: H0 H1 BIC :8>< >: if R2 Q 6 2 N ln N (29) if R2 Q > 2 N ln N 3.2.3 Minimum description length (MDL) MDL (e.g. Rissanen 2001) is an inference approach mostly developed during the 80s and 90s, based on the key idea that the more regular a given data set is, the higher is the compression degree to which we can code the data, and, therefore, the more we can learn on the properties of the data. Among many statistical applications, MDL is used to select between competing models describing the data, select- ing the one allowing for a higher compression degree (which can be seen as an alternative formulation of the Occam's Razor). For our particular case described by equation 3, the MDL measurement of compression is given by: MDL(ǫ) = −l + 1 2 ln„ N 2π« + lnZǫ∈Ω dǫ[det Fǫ]1/2, (30) where Fǫ is the Fisher matrix of ǫ. Therefore, taking into account equations 12 and 15, one can easily compute MDL(H1) and MDL(H0) providing a decision rule that reads as: MDL : , (31) H0 H1 8>>>>< >>>>: if R2 Q 6 4 if R2 Q > 4 N ln„N Ωq Q π« N ln„N Ωq Q π« where Ω is the interval where ǫ is defined. 3.2.5 Bayesian evidence (BE) BE is defined as the average likelihood of the model Hi in the prior p(ǫ): EHi (x) =Z dǫ p(ǫ, Hi)p(xǫ, Hi), (34) where p(xǫ, Hi) is given by equation 12. Model selection in terms of the BE grounds on the Bayes' factor, B10: B10(x) = EH1 (x) EH0 (x) . (35) BE framework provides a rule to quantify how strong the decision is. In the literature it is commonly accepted the Jef- freys' scale (Jeffreys 1961), that provides a recipe in terms of the logarithmic Bayes' factor. Roughly speaking, it is com- monly said that the evidence for H1 against H0 is not sig- nificant if 0 6 ln B10(x) < 1, mild if 1 6 ln B10(x) 6 3 and strong if ln B10(x) > 3. As it was already discussed, the alternative hypothesis H1 representing the local non-Gaussian model (equation 3) does not offer any physical motivation for a particular prior. Even thus, we study here the two particular cases already mentioned in subsection 3.1.2: the Gaussian (equation 24) and the uniform (equation 23) priors. For the former, it is straightforward to prove that: B10 = σǫ ǫ + σ2 ∗ pσ2 whereas for the latter, one can obtain: ǫ2 2σ2 ∗ ∗ + − e ǫ +ǫσ2 (ǫ∗ σ2 ǫ (σ2 σ2 2σ2 ∗)2 ǫ +σ2 ∗ ∗) , (36) B10 =r π 2 σǫ ǫM − ǫm ǫ2 2σ2 e ǫ »erf„ ǫM − ǫ √2σǫ « + erf„ ǫ − ǫm√2σǫ « -- , (37) where ǫ is the maximum-likelihood estimation for the non- linear parameter ǫ (equation 20) and σǫ is the error on this estimation (equation 21). Finally, notice that for the par- ticular cases of σ∗ ≫ σǫ or ǫM − ǫm ≫ σǫ in equations 36 and 37, respectively (i.e., a broad prior), one obtains: 3.2.4 Generalized likelihood ratio test (GLRT) Generalized likelihood ratio test (GLRT) is one of the most common approaches in model selection and its particular application to solve astronomical/cosmological problems has been very extensive. The criterion established by the GLRT to accept the al- ternative hypothesis H1 against H0 is given by p(xǫ, H1) > eνp(x0, H0) or, equivalently by: l ≡ l(ǫ) = N R2 4Q > ν, (32) B10 ≃ ǫ2 2σ2 ǫ , e σǫ γ (38) where γ = σ∗ (ǫM − ǫm) /√2π (for the uniform prior). (for the Gaussian prior) or γ = 4 APPLICATION TO WMAP SIMULATIONS In this Section we aim to explore the performance of the parameter estimators and the model selection crite- ria described in the previous Section. We apply them to Analysis of NG CMB maps based on the N-pdf 7 computed this large number of simulations to assure an ac- curate description of the CMB Gaussian temperature fluc- tuations. Additional 1,000 simulations were also generated to carry out a statistical analysis on the performance of the different parameters estimators and model selection criteria. Each of these 1,000 (∆T )G simulations are transformed into x (following equations 1 and 4) to study the response of the statistical tools as a function of the non-linear ǫ pa- rameter defining the local non-Gaussian model proposed in equation 3. 4.1 Parameter estimation Let us first consider the estimation made via maximum- likelihood estimation (Subsection 3.1.1). As it was men- tioned above, we have generated 1,000 non-Gaussian WMAP-like observations according to equation 3 for a range of values of ǫ, in particular, we have considered ǫ ∈ [0, 0.035], or equivalently, in terms of the most common coupling fNL parameter (equation 2), we explore fNL ∈ [0, 500]. We only explore positive values of the non-linear parameter, since the response of the proposed methodology does not depend on the sign of ǫ. In figure 2 we present the ǫ distributions obtained from the analysis of local non-Gaussian simulations. From left to right and from top to bottom, the panels show the cases: ǫ = 0, 0.001, 0.002, 0.003, 0.004, 0.005, 0.010, 0.015, 0.020, 0.025, 0.030 and 0.035. Notice that, as expected, the pa- rameter estimation is unbiased for reasonable values of the ǫ parameter: all the distributions are peaked around the value of ǫ used to generate the simulations. Only for ǫ > 0.025 (or, equivalently, fNL ' 350) a bias starts to appear. This effect comes from the fact that these values of ǫ are not small enough to assure a local non-Gaussian model as the one de- scribed by equation 3, i.e., a non-Gaussian model that is a local perturbation of the underlaying CMB Gaussian signal. Also as expected (see equation 21), the width of these dis- tributions does not depend on the particular value of ǫ, if, once more, ǫ is small enough to assure a proper expansion for the local non-Gaussian model. In this regime, we obtain, on average, σǫ ≈ 0.004, or, equivalently, σ fNL ≈ 60. Again, for ǫ > 0.025 the width of the distributions starts to be slightly smaller, indicating an inadequate value of the non-linear pa- rameter. These two effects (bias of the maximum-likelihood parameter estimation and dependence on the error on the parameter estimation) provide a natural range (at least for a pixel resolution of ≈ 2◦) where the non-linear parameter is allowed to take values: ǫ ∈ [−0.025, 0.025]. This allowed range for ǫ can be seen as a natural prior p (ǫ), that could be used for performing a parameter estima- tion within a Bayesian framework. Obviously (as discussed in Subsection 3.1.2), the Bayesian estimation made with this uniform prior produces the same estimations for ǫ already reported from the maximum-likelihood. Finally, we have also investigated whether the value of S in equation 17 is negligible as compared to −2 + J, which, as discussed in Section 2, would lead to a situation where the choice of a particular value for the non-linear parameter α becomes an irrelevant problem. We found that, actually, this is not the case: on average, S/(−2 + J) ≈ 0.7. This implies that, first, different values of α could provide different re- sults, not only to the ones presented in this work, but also for Figure 1. Mask at NSIDE=32 HEALPix resolution used in this work. It corresponds to the WMAP KQ75 mask, although the point source masking has not been considered, since the point like-emission due to extragalactic sources is negligible at the con- sidered resolution. At this pixel resolution, the mask keeps around 69% of the sky. CMB simulations of the WMAP 5-year data at NSIDE=32 HEALPix (G´orski et al. 2005) resolution (≈ 2◦). The procedure to generate a CMB Gaussian simulation -- (∆T )G in equation 1 -- is as follows. First, using the Cℓ obtained with the cosmological parameters provided by the best-fit to WMAP data alone (Table 6 in Hinshaw et al. 2008), we simulate WMAP observations (taking into ac- count the corresponding beam window functions) for the Q1, Q2, V1, V2, W1, W2, W3, W4 difference assemblies at NSIDE=512 HEALPix resolution. We obtain a single co- added CMB map through a noise-weighted linear combi- nation of the eight maps (from Q1 to W4). Weights are proportional to the inverse mean noise variance. They are independent on the position (i.e., they are uniform across the sky for a given difference assembly) and they are nor- malized to unity. Notice that we do not add a random noise realization to each map, since we have checked that noise plays a negligible role at the angular resolution in which we are interested in (≈ 2◦). However, we perform the linear combination of the difference assembly maps following the procedure described above, since it will be the same pro- cess that we will follow with the WMAP data.1 Afterwards, the co-added map at NSIDE=512 is degraded down to the final resolution of NSIDE=32. Finally, a mask representing a sky coverage like the one allowed by the WMAP KQ75 mask (Gold et al. 2008) is adopted. At NSIDE=32 the mask keeps around 69% of the sky (notice that we do not consider the masking due to point sources, since at this resolution the contribution from individual extragalactic point sources is negligible, see figure 1). Let us remark that observational constraints like incomplete sky coverage can be easily taken into account by the local non-Gaussian model proposed in this work, since it is naturally defined in pixel space. We have used 500,000 simulations of (∆T )G, generated as described above, to estimate the correlation matrix ξ ac- counting for the Gaussian CMB cross-correlations. We have 1 Co-added WMAP 5-year data is made in this way to produce a final map with a noise level smaller than, for instance, the one that could be achieved just by averaging the 8 difference assembly maps, assuring better a negligible noise contribution to the final map at resolution of ≈ 2◦. 8 P. Vielva and J. L. Sanz 180 160 140 120 100 80 60 40 20 0 180 160 140 120 100 80 60 40 20 0 180 160 140 120 100 80 60 40 20 0 180 160 140 120 100 80 60 40 20 0 ∈ = 0.000 f NL = 0 −0.02 0 0.02 0.04 ǫ ∈ = 0.003 f NL = 44 −0.02 0 0.02 0.04 ǫ ∈ = 0.010 = 146 f NL −0.02 0 0.02 0.04 ǫ ∈ = 0.025 = 364 f NL −0.02 0 0.02 0.04 ǫ 180 160 140 120 100 80 60 40 20 0 180 160 140 120 100 80 60 40 20 0 180 160 140 120 100 80 60 40 20 0 180 160 140 120 100 80 60 40 20 0 ∈ = 0.001 f NL = 15 −0.02 0 0.02 0.04 ǫ ∈ = 0.004 f NL = 58 −0.02 0 0.02 0.04 ǫ ∈ = 0.015 = 218 f NL −0.02 0 0.02 0.04 ǫ ∈ = 0.030 = 436 f NL −0.02 0 0.02 0.04 ǫ 180 160 140 120 100 80 60 40 20 0 180 160 140 120 100 80 60 40 20 0 180 160 140 120 100 80 60 40 20 0 180 160 140 120 100 80 60 40 20 0 ∈ = 0.002 f NL = 29 −0.02 0 0.02 0.04 ǫ ∈ = 0.005 f NL = 73 −0.02 0 0.02 0.04 ǫ ∈ = 0.020 = 291 f NL −0.02 0 0.02 0.04 ǫ ∈ = 0.035 = 509 f NL −0.02 0 0.02 0.04 ǫ Figure 2. Distributions for the maximum-likelihood estimations of the non-linear parameter ǫ obtained by analyzing 1,000 simulations, according to the local non-Gaussian model given in equation 3. Several values of ǫ, or, equivalently of the coupling fNL parameters (both numbers are written in each panel) are explored. Vertical dashed lines indicate the value of the ǫ used to generate each set of local non-Gaussian simulations. other works in the literature (where, we recall, it is assumed α ≡ 0). In particular, it is trivial to show that, values of α . 0.15 would affect the determination of the coupling pa- rameter ǫ in ≈ 10%. Second, as it has been discussed by some authors (Okamoto & Hu 2002; Kogo & Komatsu 2006), it would indicate that, for the weak non-linear coupling infla- tionary model, the role played by cubic terms could be non negligible as compared to quadratic contributions, since, to some extent, they will contribute to the full trispectrum (as one can notice from equation 17, where α governs the role played by S). This results could be important when describ- ing more complete non-local non-Guassian model, since it would indicate the need of including physical effects up to third order. 4.2 Model selection We discuss which is the performance of the different model selection criteria presented in Subsection 3.2. For the par- ticular case of BE, we assume the natural prior p (ǫ) found in the previous Subsection: ǫ ∈ [−0.025, 0.025]. As for the case of the parameter estimation previously discussed, we have considered a set of local non-Gaussian models given by different values of the non-linear parameter. We present the results, graphically, in figure 3. This plot consist in 6 rows and 5 columns. Each row corresponds to the results obtained for a given value of the ǫ parameter (namely, from top to bottom: ǫ = 0, 0.005, 0.010, 0.015, 0.020 and 0.025). Each column refers to a model selection crite- rion (from left ro right: AIC, BIC, MDL, BE and GLRT). For the first column (i.e., the AIC case) we plot the dis- tributions (obtained after analysing 1,000 simulations) of a statistical variable defined as: WAIC ≡ R2/Q − 4/N , no- tice (from equation 27) that a positve value of WAIC implies to accept H1 against H0. The second column accounts for the distributions of the variable WBIC ≡ R2/Q − 2 ln N/N , which (from equation 29) also satisfies to be positive when favoring H1. Equivalently, third column shows the distribu- tions of WMDL ≡ R2/Q − 4/N ln N ΩpQ/π, that according to equation 31 is also positive when H1 is more likely than H0. Fourth column provides ln B10, obtained from equa- tion 37. Finally, in the fifth column we present the distri- butions obtained for ν = `R2/N´ / (4/N ) which, according to the GLRT decision rule (equation 33) provides a mea- surement of the strength in accepting H1 against H0. In addition to the distributions, we also plot, as an in- dication, a vertical line in each panel. For WAIC, WBIC and WMDL, this vertical line separate the region where H0 is fa- vored (left side) from the one where H1 is more likely (right side). The percentage of the 1,000 simulations that fall on the H1 region is also reported at each panel. The vertical line for the WBE statistic represents the limit for which the logarithmic Bayes' factor (ln B10) is greater than 1, which, in terms of the Jeffreys' rule, indicates that H1 is, at least, mildly favored against H0. We also provide the percentage of the simulations satisfying this condition. Finally, the ver- tical line for the WGLRT statistic indicates the value of ν for which 95% of the 1,000 simulations favor H1 instead of H0. This value of ν is also written in the panels. Notice that, among the asymptotic model selection criteria (AIC, BIC and MDL), AIC offers a less restrictive criterion than BIC, whereas BIC behaves similarly with respect to MDL, Analysis of NG CMB maps based on the N-pdf 9 The figure also shows that BE provides a more conservative criterion than the asymptotic methods. 5 APPLICATION TO WMAP 5-YEAR DATA We have applied the statistical approaches described in Sec- tion 3 to WMAP 5-year data. In particular, we have ana- lyzed a co-added CMB map generated from the global noise- weighted linear combination of the reduced foreground maps for the Q1, Q2, V1, V2, W1, W2, W3 and W4 difference as- semblies (see Gold et al. 2008, for details). Weights are nor- malized to unity and, for each map, they are proportional to the inverse average noise variance across the sky. This op- eration is made at NSIDE=512 HEALPix resolution, being degraded afterwards down to NSIDE=32. Hence, we are in the same conditions as for the analysis on simulations described in the previous Section and, there- fore, the CMB cross-correlation in WMAP data is given by the ξ correlation matrix already defined in Section 4. The es- timated full N-pdf of the WMAP 5-year data given the non- linear parameter ǫ is showed in figure 4 (indeed, it is given in terms of the most common fNL parameter for allowing a bet- ter comparison with previous works). Maximum-likelihood estimation (equation 20) provides fNL = 30 with an error for the parameter (equation 21) of σfNL = 62 (compatible with the values obtained from simulations). Hence our WMAP 5-year data analysis reports: fNL = 30 ± 124 at 95% CL (or, equivalently, ǫ = 0.019 ± 0.078 at 95%). This result is compatible with similar works in the literature reporting WMAP compatibility with Gaussian hypothesis. However, let us remark that this estimation is more efficient than pre- vious ones at similar angular resolution, since it provides a smaller error bar. For instance, Curto et al. (2007) per- formed a Gaussianity test on the WMAP data, at the same HEALPix resolution, although in a smaller region of the sky (16% instead of the 69% considered in this work). Their fNL estimator was based on the three Minkowski functionals, providing an error bar of ≈ 200. The expected σfNL provided by our maximum-likelihood estimator for a similar observed region would be σfNL ≈ 124, i.e., ≈ 40% smaller than the one obtained by the Minkowski functionals. This result was expected since, as it was already discussed, the maximum- likelihood estimation of the non-linear parameter is optimal, given the local non-Gaussian model in equation 1. As it has been mentioned above, this result shows WMAP data compatibility with the Gaussian hypothesis, since fNL ≡ 0 can not be rejected at any significant confi- dence level. Of course, same conclusions are obtained from the model selection criteria described in Subsection 3.2. Nei- ther AIC, BIC, MDL nor BE criteria select H1 against H0, whereas GLRT would favor H1 under the very weak condi- tion for ν in equation 33 of ν ≈ 0.1 (which implies a likeli- hood ratio of ≈ 1.1) Finally, let us remark that, as it also happened for the simulations analyzed in Section 4, the contribution of the S term to the value of Q (see equation 17) is not negligible, in particular, S/(−2 + J) = 0.74. This indicates that, as it was already mentioned, the cubic term in the model given by equation 3 is not naturally negligible as compared to the quadratic term and, therefore, α should be chosen small enough (like the case α ≡ 0 considered in this work). 10 P. Vielva and J. L. Sanz #(>0) = 17.0% ∈ = 0.000 f NL = 0 600 500 400 300 200 100 0 −0.5 0 0.5 1 1.5 W AIC 2 2.5 3 3.5 x 10−3 #(>0) = 43.6% ∈ = 0.005 f NL = 73 0 1 2 W AIC 3 4 x 10−3 #(>0) = 85.4% ∈ = 0.010 = 146 f NL #(>0) = 0.4% ∈ = 0.000 f NL = 0 −2.5 −2 −1.5 −1 −0.5 W BIC 0 0.5 1 1.5 x 10−3 #(>0) = 4.4% ∈ = 0.005 f NL = 73 −2 −1 0 W BIC 1 2 3 x 10−3 #(>0) = 31.8% ∈ = 0.010 = 146 f NL 600 500 400 300 200 100 0 400 350 300 250 200 150 100 50 0 140 120 100 80 60 40 20 0 120 100 80 60 40 20 400 350 300 250 200 150 100 50 0 140 120 100 80 60 40 20 0 120 100 80 60 40 20 0 140 120 100 80 60 40 20 0 140 120 100 80 60 40 20 0 0 2 0 2 4 −2 −1 0 1 2 3 4 5 W BIC 4 6 W AIC 8 x 10−3 #(>0) = 99.1% ∈ = 0.015 = 218 f NL 6 x 10−3 10 x 10−3 #(>0) = 77.8% ∈ = 0.015 = 218 f NL 4 6 8 W BIC #(>0) = 98.0% ∈ = 0.020 = 291 f NL 6 8 10 0 −2 0 2 W AIC x 10−3 #(>0) = 99.9% ∈ = 0.020 = 291 f NL 140 120 100 80 60 40 20 0 5 10 W AIC 15 x 10−3 0 −2 0 2 4 6 W BIC 8 10 12 14 x 10−3 #(>0) = 100.0% ∈ = 0.025 = 364 f NL 10 15 W AIC 140 120 100 80 60 40 20 0 20 x 10−3 0 5 0 5 #(>0) = 99.9% ∈ = 0.025 = 364 f NL 10 15 W BIC x 10−3 500 400 300 200 100 0 350 300 250 200 150 100 50 0 150 100 50 0 140 120 100 80 60 40 20 0 120 100 80 60 40 20 0 140 120 100 80 60 40 20 0 #(>0) = 0.1% ∈ = 0.000 f NL = 0 −3.5 −3 −2.5 −2 −1 −0.5 0 −1.5 W MDL 0.5 x 10−3 #(>0) = 0.9% ∈ = 0.005 f NL = 73 −3 −2 −1 W MDL 0 1 2 x 10−3 #(>0) = 16.0% ∈ = 0.010 = 146 f NL −3 −2 −1 0 1 2 3 4 5 W MDL x 10−3 #(>0) = 59.3% ∈ = 0.015 = 218 f NL −2 0 4 6 2 W MDL 8 x 10−3 #(>0) = 93.5% ∈ = 0.020 = 291 f NL −2 0 2 4 0 5 6 8 10 12 W MDL x 10−3 #(>0) = 99.7% ∈ = 0.025 = 364 f NL 10 15 W MDL x 10−3 500 450 400 350 300 250 200 150 100 50 0 350 300 250 200 150 100 50 0 150 100 50 0 120 100 80 60 40 20 0 120 100 80 60 40 20 0 140 120 100 80 60 40 20 0 #(>1) = 0.0% ∈ = 0.000 f NL = 0 −20 −15 −5 0 −10 ln(B ) 10 #(>1) = 0.0% ∈ = 0.005 f NL = 73 −20 −15 −5 0 −10 ln(B ) 10 #(>1) = 0.0% ∈ = 0.010 = 146 f NL −20 −15 −5 0 −10 ln(B ) 10 #(>1) = 1.1% ∈ = 0.015 = 218 f NL −20 −15 −20 −15 −10 −10 ln(B −5 ) 10 −5 ln(B 0 ) 10 0 5 #(>1) = 16.0% ∈ = 0.020 = 291 f NL 5 10 15 #(>1) = 60.1% ∈ = 0.025 = 364 f NL −15 −10 −5 0 10 15 20 25 5 ln(B ) 10 600 500 400 300 200 100 0 400 350 300 250 200 150 100 50 0 140 120 100 80 60 40 20 0 120 100 80 60 40 20 0 140 120 100 80 60 40 20 0 140 120 100 80 60 40 20 0 ν(>95%) = 0.001 ∈ = 0.000 f NL = 0 0 2 4 ν 6 8 ν(>95%) = 0.006 ∈ = 0.005 f NL = 73 0 2 4 ν 6 8 10 ν(>95%) = 0.372 ∈ = 0.010 = 146 f NL 0 5 10 ν 15 ν(>95%) = 2.263 ∈ = 0.015 = 218 f NL 0 5 10 ν 15 20 25 ν(>95%) = 5.754 ∈ = 0.020 = 291 f NL 0 5 10 15 ν 20 25 30 35 ν(>95%) = 10.737 ∈ = 0.025 = 364 f NL 0 10 20 ν 30 40 Figure 3. From left ro right, columns show the distribution for several statistics referred to different model selection criteria: AIC, BIC, MDL, BE and GLRT. From top to bottom, results for local non-Gaussian models for different non-linear parameters are give: ǫ = 0, 0.005, 0.010, 0.015, 0.020 and 0.025. Panels in columns 1, 2, 3 and 4 present a vertical line, separating the region where H0 and H1 are preferred (to the left and to the right of the vertical line, respectively). The vertical line on the last column represents the value for the ν parameter for which 95% of the non-Gaussian simulations are more likely described by H1 rather than H0. Analysis of NG CMB maps based on the N-pdf 11 x 10−3 7 6 5 4 3 2 1 ) L N f x ( p 0 −400 −200 200 400 0 f NL Figure 4. This curve represent the probability given in equa- tion 12, i.e. the full pdf of the WMAP 5-year data x given the non-linear parameter ǫ (or, equivalently, fNL). Vertical dotted line marks the maximum-likelihood estimation fNL = 30. 6 CONCLUSIONS We have presented a parametric non-Gaussian model for the CMB temperature fluctuations. The non-Gaussian model is a local perturbation (up to third order) of the stan- dard CMB Gaussian field which recovers (for the case of b ≡ 0 in equation 1) an approximative form of the weak non-linear coupling inflationary model (e.g. Komatsu et al. 2001; Liguori et al. 2003) at scales larger than the horizon scale at the recombination time (i.e. above the degree scale). For this model, we are able to build the posterior prob- ability of the data given the non-linear parameter ǫ (see equation 3), from which, in principle, an optimal estimator (i.e., unbiased and with minimum variance) can be derived. Analytical expressions for the maximum-likelihood estima- tion of the non-linear parameter (ǫ) and its associated error (σǫ) are derived. In addition, we also discuss an alterna- tive Bayesian estimation (in terms of the posterior prob- ability of the non-linear parameter), for the hypothetical case in which we might have some prior information for ǫ. As an example, two cases are addressed: a non-informative (i.e, uniform) and a Gaussian priors. We also investigate an issue very much linked to the parameter estimation: the model selection. Indeed, we discuss several well known tech- niques to perform hypotheses test, like the Akaike infor- mation criterion (AIC, Akaike 1973), the Bayesian informa- tion criterion (BIC, Schwarz 1978), the minimum description length (MDL, Rissanen 2001), the generalized likelihood ra- tio test (GLRT) and the Bayesian evidence (BE). We derive analytical expressions, for the particular local non-Gaussian model proposed in this work, for all these model selection techniques. The performance of both, parameter estimators and model selection criteria, are investigated by analyzing non- Gaussian simulations, as they could be observed by WMAP. We check that the maximum-likelihood estimation provides an unbiased and efficient estimation of the non-linear pa- rameter defining the deviations from Gaussianity. We find that, for the HEALPix resolution considered in this work (NSIDE=32), results are consistent up to a value of ǫ = 0.025, which approximately corresponds (at the Sachs-Wolfe regime) to a value of the the non-linear coupling parameter fNL ≈ 350. This parameter is the one commonly used to described the weak non-linear coupling inflationary model (e..g Komatsu et al. 2001). We also find that, among the model selection criteria, AIC is the asymptotic method that provides the less restrictive decision rule, whereas, on the other hand, MDL is the most strict one. We also find that BE, for a uniform prior given by ǫ ∈ [−0.025, 0.025], is even more restrictive than MDL. The proposed methodology is applied to WMAP 5-year data. We obtain a value for the non-linear coupling parame- ter of fNL = 30±124 at 95% CL. This result provides a more efficient estimation than previous works in the literature, at the same angular scales. For instance, comparing with the work by Curto et al. (2007) using Minkowski functionals, we can infer that the maximum-likelihood error bar is ≈ 40% smaller than the one obtained with those geometrical esti- mators. Application of model selection criteria to WMAP data confirms that standard hypothesis of Gaussianity is fa- vored against the alternative hypothesis of non-Gaussianity, for the specific local model proposed in this work, and for the adopted resolution of ≈ 2◦. Finally, we would like to comment that, currently, we are extending the technique based on the N-pdf presented in this work, to deal with a more realistic non-local non- Gaussian model, where higher resolution CMB data are con- sidered, including as well the effect of anisotropic noise. ACKNOWLEDGEMENTS The authors thank Andr´es Curto, R. Bel´en Barreiro and Enrique Mart´ınez-Gonz´alez for useful comments and dis- cussion. We acknowledge partial financial support from the Spanish Ministerio de Ciencia e Innovaci´on project AYA2007-68058-C03-02. PV also acknowledges financial support from the Ram´on y Cajal programme. PV thanks to the CNR Istituto de Scienza e Tecnologie dell'Informazione (ISTI, Pisa) for their warm hospitality during his research stays in March and June 2008. JLS acknowledge partial financial support by the Spanish MEC and thanks the CNR ISTI in Pisa for their hospitality during his sabbati- cal leave. The authors acknowledge the computer resources, technical expertise and assistance provided by the Span- ish Supercomputing Network (RES) node at Universidad de Cantabria. We acknowledge the use of Legacy Archive for Microwave Background Data Analysis (LAMBDA). Support for it is provided by the NASA Office of Space Science. The HEALPix package was used throughout the data analysis G´orski et al. (2005). 12 P. Vielva and J. L. Sanz REFERENCES Abramo L.R., Bernui A., Ferreira I.S., Villela T., Wuensche C.A., 2006, Phys. Rev. D, 74, 063506 Ackerman L., Carroll S.M., Wise M.B. 2007, Phys. Rev. D, 75, 083502 Akaike H., 1973, Proceed. of the 2nd International Sympo- sium on Information Theory (eds. Pertov B.N. & Czaki F), Akad. Kiado, Budapest, 267 Babich D., 2005, Phys. Rev. D., 72, 043003 Bartolo N., Komatsu E., Matarrese S., Riotto A., 2004, Phys. Rep., 402, 103 Bernui A., Villela T., Wuensche C.A., Leonardi R., Ferreira I., 2006, A&A, 454, 409 Bernui A., Mota B., Rebou¸cas M.J., Tavakol R., 2007, A&A, 464, 479 Bielewicz P., Eriksen H.K., Banday A.J., G´orski K.M., Lilje P.B., 2005, ApJ, 635, 750 Bohmer C.G., Mota D.F., 2008, Phys. Lett. B, 663, 168 Borowiec A., Godlowski W., Szydlowski M., 2006, Phys. Rev. D, 74, 043502 Bridges M., Lasenby A.N., Hobson M.P., 2006, MNRAS, 369, 1123 Bridges M., McEwen J.D., Lasenby A.N., Hobson M.P., 2007a, MNRAS, 377, 1473 Bridges M., Lasenby A.N., Hobson M.P., 2007b, MNRAS, 381, 68 Bridges M., McEwen J.D., Cruz M., Lasenby A.N., Hob- son M.P., Vielva P., Mart´ınez-Gonz´alez E., 2008, MNRAS, 390, 1372 Cabella P., Liguori M., Hansen F.K., Marinucci D., Matar- rese S., Moscardini L., Vittorio N., 2005, MNRAS, 358, 684 Carvalho P., Rocha G., Hobson M.P., 2008, MNRAS, sub- mited (preprint arXiv0802.3916) Cay´on L., Mart´ınez-Gonz´alez E., Argueso F., Banday A.J., G´orski K.M., 2003, MNRAS, 339, 1189 Cay´on L., Jin J., Treaster A., 2005, MNRAS, 362, 826 Chiang L.-Y., Naselsky P.D., Verkhodanov O. V., 2003, ApJ, 590, 65 Chiang L.-Y., Naselsky P. D., 2006, Int. J. Mod. Phys. D, 15, 1283 Barreiro R.B., Santos D., D´esert F.-X., Tristram M., 2007, A&A, 474, 23 Curto A., Mart´ınez-Gonz´alez E., Mukherjee P., Barreiro R.B., Hansen F.K., Liguori M., Matarrese S., 2008, MN- RAS, in press Davis T.M. et al., 2007, ApJ, 666, 716 de Oliveira-Costa A., Tegmark M., Zaldarriaga M., Hamil- ton A., 2004, Phys. Rev. D, 69, 063516 Donoghue E.P., Donoghue J.F., 2005, Phys. Rev. D, 71, 043002 Eriksen H.K., Hansen F.K., Banday A.J., G´orski K.M., Lilje P.B., 2004a, ApJ, 605, 14 Eriksen H.K., Novikov D.I., Lilje P.B., Banday A.J., G´orski K.M., 2004b, ApJ, 612, 64 Eriksen H.K., Banday A.J., G´orski K.M., Lilje P.B., 2005, ApJ, 622, 58 Eriksen H.K., Banday A.J., G´orski K.M., Hansen F.K., Lilje P.B., 2007, ApJ, 660, L81 Feroz F., Allanach B.C., Hobson M., Abdus Salam S.S., Trotta R., Weber A.M., 2008a, Journ. of High Energy Phys., 10, 64 Feroz F., Hobson M.P., Zwart J.T.L., Sounders R.D.E., Grainge K.J.B., 2008b, MNRAS, submited (preprint arXiv0811.1199) Freeman P.E., Genovese C.R., Miller C.J., Nichol R.C., Wasserman L., 2006, ApJ, 638, 1 Gordon C., 2007, ApJ, 656, 636 Gold B., 2008, ApJS, in press G´orski K.M., Hivon E., Banday A.J., Wandelt B.D., Hansen F.K., Reinecke M., Bartelmann M., 2005, ApJ, 622, 759 Gott J.R., Colley W.N., Park C.-G., Park C., Mugnolo C., 2007, MNRAS, 377, 1668 Groeneboom N.E., Eriksen H.K., 2008, ApJ, in press Hansen F.K., Cabella P., Marinucci D., Vittorio N., 2004a, ApJ, 607, L67 Hansen F.K., Banday A. J., G´orski K.M., 2004b, MNRAS, 354, 641 0, 063004 Hikage C., Matsubara T., Coles P., Liguori M., Hansen F. K., Matarrese S., 2008, MNRAS, 389, 1439 Himmetoglu B., Contaldi C.R., Peloso M., 2008a, (preprint arXiv0809.2779) Coles P., Dineen P., Earl J., Wright D., 2004, MNRAS, 350, Himmetoglu B., Contaldi C.R., Peloso M., 2008a, (preprint 989 Copi C.J., Huterer D., Starkman G.D., 2004, Phys. Rev. D, 70, 043515 Creminelli P., Nicolis A., Senatore L., Tegmark M., Zaldar- riaga M., 2006, Journal of Cosmology and Astro-Particle Physics, 5, 4 Creminelli P., Senatore L., Zaldarriaga M., 2007, Journal of Cosmology and Astro-Particle Physics, 3, 19 Cruz M., Mart´ınez-Gonz´alez E., Vielva P., Cay´on L., 2005, MNRAS, 356, 29 Cruz M., Tucci M., Mart´ınez-Gonz´alez E., Vielva P., 2006, MNRAS, 369, 57 Cruz M., Cay´on L., Mart´ınez-Gonz´alez E., Vielva P., Jin J., 2007a, ApJ, 655, 11 Cruz M., Turok N., Vielva P., Mart´ınez-Gonz´alez E., Hob- son M.P., 2007b, Science, 318, 1612 Cruz M., Mart´ınez-Gonz´alez E., Vielva P. Diego J.M., Hob- son M.P., Turok N., 2008, MNRAS, 390, 913 Curto A., Aumont J., Mac´ıas-P´erez, Mart´ınez-Gonz´alez E., arXiv0812.1231) Hinshaw G. et al., 2008, ApJS, in press Jaffe T.R., Banday A.J., Eriksen H.K., G´orski K.M., Hansen F.K., 2006a, ApJ, 629, L1 Jaffe T.R., Hervik S., Banday A.J., G´orski K.M., 2006b, ApJ, 644, 701 Jaffe T.R., Banday A.J., Eriksen H.K., G´orski K.M., Hansen F.K., 2006c, preprint (arXiv:astro-ph/0606046v1) Jeffreys H., 1961, Theory of Probability (3rd edition) Ox- ford university Press Katz G., Weeks J., 2004, Phys. Rev. D, 70, 063527 Kogo N., Komatsu E., 2006, Phys. Rev. D., 73, 083007 Komatsu E., Spergel D.N., 2001, Phys. Rev. D, 63, 063002 Komatsu E., et al., 2003, ApJS, 148, 119 Komatsu E., et al., 2008, ApJS, in press Land K., Magueijo J., 2005a, MNRAS, 357, 994 Land K., Magueijo J., 2005b, Phys. Rev. Lett., 95, 071301 Land K., Magueijo J., 2007, MNRAS, 378, 153 Liguori M., Matarrese S., Moscardini L., 2003, ApJ, 597, Analysis of NG CMB maps based on the N-pdf 13 57 Liddle A., Lyth D. H., 2000, Cosmological inflation and large-scale structure, Cambridge University Press Liddle A., 2004, MNRAS, 351, 49 Liddle A., 2007, MNRAS, 377, 74 Liddle A., Mukherjee P., Parkinson D., Wang Y., 2006, Phys. Rev. D, 74, 123506 Mart´ınez-Gonz´alez E., Cruz M., Cay´on L., Vielva P., 2006, New Astron. Rev., 50, 875 McEwen J.D., Hobson M.P., Lasenby A.N., Mortlock D.J., 2005, MNRAS, 259, 1583 McEwen J.D., Hobson M.P., Lasenby A.N., Mortlock D.J., 2006, MNRAS, 371, 50 Monteser´ın C., Barreiro R.B., Vielva P., Mart´ınez- Gonz´alez, Hobson, M.P., Lasenby A.N., 2008, MNRAS, 387, 209 Mukherjee P., Wang Y., 2004, ApJ, 613, 51 Mukherjee P., Parkinson D., Liddle A., 2006, ApJL, 638, 51 Mukherjee P., Liddle A., 2008, MNRAS, 389, 231 Okamoto T., Hu W., 2002, Phys. Rev. D., 66, 063008 Park C.-G., 2004, MNRAS, 349, 313 Pietrobon D., Amblard A., Balbi A., Cabella P., Cooray A., Mirinucci D., 2008, Phys. Rev. D, 10, 3504 Rath C., Schuecker P., Banday A.J., 2007, MNRAS, 380, 466 Reissanen J., 2001, IEEE Transactions on Information The- ory, 47, 5 Schwarz G., 1978, Ann. Stat., 6, 461 Schwarz D.J., Starkman G.D., Huterer D., Copi C.J., 2004, Phys. Rev. Lett., 93, 221301 Spergel D.N. et al., 2007, ApJS, 170, 377 Szydlowski M., Godlowski W., 2006, Phys. Lett. B, 633, 427 Szydlowski M., Kurek A., Krawiec A., 2006, Phys. Lett. B, 642, 171 Szydlowski M., Godlowski W., Stachowiak T., 2008, Phys. Rev. D, 77, 043530 Tojeiro R., Castro P.G., Heavens A.F., Gupta S., 2006, MNRAS, 365, 265 Vielva P., Mart´ınez-Gonz´alez E., Barreiro R.B., Sanz J.L., Cay´on L., 2004, ApJ, 609, 22 Vielva P., Wiaux Y., Mart´ınez-Gonz´alez E., Vandergheynst P., 2006, New Astron. Rev., 50, 880 Vielva P., Wiaux Y., Mart´ınez-Gonz´alez E., Vandergheynst P., 2007, MNRAS, 381, 932 Wiaux Y., Vielva P., Mart´ınez-Gonz´alez E., Vandergheynst P., 2006, Phys. Rev. Lett., 96, 151303 Wiaux Y., Vielva P., Barreiro R.B., Mart´ınez-Gonz´alez, Vandergheynst P., 2008, MNRAS, 385, 939 Yadav A.P.S., Wandelt B.D., 2008, Phys. Rev. Lett., 100, 181301
astro-ph/9806346
1
9806
1998-06-25T20:32:38
Mass Transfer and Accretion in the Eccentric Neutron-Star Binary Circinus X-1
[ "astro-ph" ]
I have carried out a project to study the eccentric neutron-star binary Circinus X-1 through an extensive series of observational studies with the Rossi X-ray Timing Explorer satellite and through theoretical computer models I developed to explore mass transfer and evolution in an eccentric binary. I modeled the evolution of the energy spectrum during intensity dips with a variably absorbed bright component plus a fainter unabsorbed component. I show that variability not attributable to absorption dips is related to the spectral/intensity states of the ``Z source'' class of low-mass X-ray binaries (LMXBs), namely motion along (or shifts of) the horizontal, normal, and flaring branches of the ``Z'' track in color-color and hardness-intensity diagrams. I found rapid X-ray variability properties associated with each spectral/intensity state: On the horizontal branch, quasi-periodic oscillations (QPOs) in the X-ray intensity shift in frequency from 1.3 to 35 Hz. On the normal branch, a different QPO occurs at about 4 Hz. On the flaring branch only strong aperiodic variability occurs. I also modeled the evolution of the energy spectra associated with each of these branches. To study mass transfer in an eccentric binary, I developed computer codes for transfer via Roche-lobe overflow and from a stellar wind. I derive theoretical mass accretion profiles and compare them to the observed profile of the X-ray intensity. In order to explore the possible evolutionary history of Circinus X-1, I developed a binary-evolution computer code for a neutron-star and low-mass companion in an eccentric orbit. I use this code in a population-synthesis study to show that the number of systems in the Galaxy expected to resemble Cir X-1 is of order unity, consistent with its unique status as an LMXB with high eccentricity.
astro-ph
astro-ph
Mass Transfer and Accretion in the Eccentric Neutron-Star Binary Circinus X-1 by Robert E. Shirey Submitted to the Department of Physics in partial fulfillment of the requirements for the degree of Doctor of Philosophy at the MASSACHUSETTS INSTITUTE OF TECHNOLOGY September 1998 c(cid:13) Robert E. Shirey, MCMXCVIII. All rights reserved. The author hereby grants to MIT permission to reproduce and distribute publicly paper and electronic copies of this thesis document in whole or in part, and to grant others the right to do so. Author . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Department of Physics June 19, 1998 Certified by . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Hale V. D. Bradt Professor of Physics Thesis Supervisor Accepted by . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Thomas J. Greytak Associate Department Head for Education Mass Transfer and Accretion in the Eccentric Neutron-Star Binary Circinus X-1 by Robert E. Shirey Submitted to the Department of Physics on June 19, 1998, in partial fulfillment of the requirements for the degree of Doctor of Philosophy Abstract I have carried out a project to study the eccentric neutron-star binary Circinus X-1 through an extensive series of observational studies with the Rossi X-ray Timing Explorer satellite and through theoretical computer models I developed to explore mass transfer and evolution in an eccentric binary. We also organized two multi-frequency campaigns to study correlated variability in different frequency bands. The X-ray observations showed that the intensity of Cir X-1 currently maintains a bright baseline level, with strong flares occurring after phase zero of each 16.55-day cycle of the source. This behavior is thought to be due to enhanced mass transfer occurring near periastron of a highly eccentric binary orbit. Dips below the baseline intensity level also occur near phase zero. I modeled the evolution of the energy spectrum during dips with a variably absorbed bright com- ponent plus a fainter unabsorbed component. I show that variability not attributable to absorption dips is related to the spectral/intensity states of the "Z source" class of low-mass X-ray binaries (LMXBs), namely motion along (or shifts of) the horizontal, normal, and flaring branches of the "Z" track in color-color and hardness-intensity diagrams. I found rapid X-ray variability properties associated with each spectral/intensity state: On the horizontal branch, quasi-periodic oscillations (QPOs) in the X-ray intensity shift in frequency from 1.3 to 35 Hz. On the normal branch, a different QPO occurs at about 4 Hz. On the flaring branch only strong aperiodic variability occurs. I modeled the evolution of the energy spectra associated with each of these branches. To study mass transfer in an eccentric binary, I developed computer codes for transfer via Roche- lobe overflow and from a stellar wind. I derive theoretical mass accretion profiles and compare them to the observed profile of the X-ray intensity. In order to explore the possible evolutionary history of Circinus X-1, I developed a binary- evolution computer code for a neutron-star and low-mass companion in an eccentric orbit. I use this code in a population-synthesis study to show that the number of systems in the Galaxy expected to resemble Cir X-1 is of order unity, consistent with its unique status as an LMXB with high eccentricity. Thesis Supervisor: Hale V. D. Bradt Title: Professor of Physics Acknowledgments First, I would like to thank Prof. Hale Bradt for giving me the opportunity to work with him and participate in the Rossi X-ray Timing Explorer project. He has demonstrated how to think like a scientist and produce clear scientific writings. He has always supported me and encouraged my efforts. I also want to thank Prof. Saul Rappaport for helping with the several theoretical modeling projects in my thesis. Although I was initially reluctant to delve into theoretical calculations, I eventually found that they were not to be feared and could be quite interesting. Ed Morgan got me started on Circinus X-1 several years ago by suggesting I work on a proposal to observe Cir X-1 with RXTE once it launched. I also learned many analysis techniques from Ed. Al Levine has provided much-appreciated scientific guidance, especially in writing papers. His participation has certainly improved the quality of my work. In fact, all the members of the RXTE team at M.I.T. have contributed to my education and to this project. Ron Remillard, Wei Cui, and Deepto Chakrabarty have all given helpful advice and shared experiences with analyzing and interpreting RXTE data. The people I've interacted with most at M.I.T. are the other graduate students, and my office mates in particular. Don Smith came to M.I.T. at the same time I did, so we've shared all the ups and downs of the experience. He, Linqing Wen, and Mike Muno have all been great to work with and to discuss ideas with around the office. Our former office mates Charlie Collins and Chris Becker provided a wealth of knowledge and experience about M.I.T., Boston, and astrophysics. In particular my binary evolution code was based on one Chris developed as part of his thesis. I would like to thank all the members of our multi-frequency campaigns for all their efforts to coordinate observations and willingness to share data: George Nicolson, Ian Glass, Allyn Tennant, Rob Fender, Kinwah Wu, and Helen Johnston. My parents have supported my interest in astronomy from the beginning. They gave me a 6-inch diameter telescope for Christmas when I was in elementary school (and then let me sell it several years later when an Atari system seemed more important). Finally, I want to thank my wife, Anne, for supporting me in so many ways throughout the whole process. When we were married, exactly five years ago, I'm sure she didn't know what she was getting into. She has made many sacrifices to help make this happen. Now five years and two children (Ben and Alex) later, we've finally reached this milestone. I thank her for her patience, support, and love. This thesis is for Anne. Robert Shirey Cambridge, MA June 19, 1998 Contents 1 Introduction 1.1 X-ray Astronomy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2 X-ray Binaries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2.2 HMXBs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2.3 LMXBs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.3 Circinus X-1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.3.1 Intensity Cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.3.2 X-ray Bursts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.3.3 Rapid X-ray Variability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.3.4 Distance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.3.5 Young Runaway Binary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.3.6 Infrared and Optical Counterpart . . . . . . . . . . . . . . . . . . . . . . . . . 1.3.7 Radio Ephemeris . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.4 Objectives and Overview of this Project . . . . . . . . . . . . . . . . . . . . . . . . . 2 The Rossi X-ray Timing Explorer 2.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2 Proportional Counter Array . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3 High-Energy X-ray Timing Experiment . . . . . . . . . . . . . . . . . . . . . . . . . 2.4 All-Sky Monitor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.5 Experiment Data System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3 Overview of RXTE Observations 3.1 ASM Observations of Circinus X-1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1.1 General Features of the ASM Light Curves . . . . . . . . . . . . . . . . . . . 3.1.2 Cycling Hardness Ratios . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1.3 16.55-day Period . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5 23 23 23 23 24 25 28 28 28 29 29 29 30 31 31 33 33 33 35 36 37 39 39 39 41 41 6 CONTENTS 3.1.4 Folded Cycle Profile . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2 PCA Observations of Circinus X-1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2.2 Observation Summaries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2.2.1 Study A: 1996 March 5–21 . . . . . . . . . . . . . . . . . . . . . . . 3.2.2.2 Study B: 1996 April 7 . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2.2.3 Study C: 1996 May 7–14 . . . . . . . . . . . . . . . . . . . . . . . . 3.2.2.4 Study D: 1996 Sept. 20 – Oct. 8 . . . . . . . . . . . . . . . . . . . . 3.2.2.5 Study E: 1997 Feb. 18 – Mar. 4 . . . . . . . . . . . . . . . . . . . . 3.2.2.6 Study F: 1997 May 16 – May 21 . . . . . . . . . . . . . . . . . . . . 3.2.2.7 Study G: 1997 June 4–22 . . . . . . . . . . . . . . . . . . . . . . . . 3.2.2.8 Study H: 1997 Sept. 18 – Oct. 4 . . . . . . . . . . . . . . . . . . . . 4 X-ray Timing and Spectral Evolution vs. Orbital Phase 4.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2 Observations and Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5 QPOs Associated with Spectral Branches 5.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.2 Observations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3 Analysis and Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3.1 Color-color and Hardness-intensity Diagrams . . . . . . . . . . . . . . . . . . 5.3.2 Power Density Spectra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3.3 Temporal Behavior versus Position on Spectral Branches . . . . . . . . . . . . 5.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4.1 Horizontal Branch . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4.2 Normal Branch . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4.3 Flaring Branch . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4.4 Relation to Other Sources . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6 Correlated Timing and Spectral Behavior 6.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.2 1997 June PCA Observations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41 44 44 46 46 48 49 49 50 50 50 50 53 53 54 59 61 61 62 64 64 67 73 73 73 75 76 76 79 81 81 82 CONTENTS 6.3 Complete Spectral Track . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.4 Evolution of the Power Density Spectrum . . . . . . . . . . . . . . . . . . . . . . . . 6.5 Evolution of the Energy Spectrum . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.6 Selection of Spectral Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.7 Fits to Spectra from 20 HID Regions . . . . . . . . . . . . . . . . . . . . . . . . . . . 7 82 89 91 93 96 6.8 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101 7 Absorption Dips 103 7.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103 7.2 Dips in RXTE Light Curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103 7.3 Evolution of Hardness Ratios During Dips . . . . . . . . . . . . . . . . . . . . . . . . 104 7.4 Evolution of Energy Spectra During Dips . . . . . . . . . . . . . . . . . . . . . . . . 109 7.4.1 Spectral Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109 7.4.2 Spectral Fitting Strategy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112 7.4.3 Energy Spectra Inside and Outside Dips . . . . . . . . . . . . . . . . . . . . . 113 7.4.4 Iron Emission Line . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117 7.4.5 Spectral Evolution During Dip Transitions . . . . . . . . . . . . . . . . . . . 118 7.5 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122 8 Multi-frequency Observations 125 8.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125 8.2 May 1996 Campaign . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125 8.2.1 Radio (HartRAO) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127 8.2.2 Infrared Photometry (SAAO) . . . . . . . . . . . . . . . . . . . . . . . . . . . 128 8.2.3 Correlated Multi-frequency Variability . . . . . . . . . . . . . . . . . . . . . . 128 8.3 June 1997 Campaign . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128 8.3.1 Radio (HartRAO) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130 8.3.2 IR Photometry (SAAO) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130 8.3.3 IR Spectroscopy (ANU) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132 8.3.4 Optical Spectroscopy (AAT) . . . . . . . . . . . . . . . . . . . . . . . . . . . 132 8.3.5 Comparison of the AAT Optical Spectrum to HST Results . . . . . . . . . . 132 8.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135 9 Mass Transfer in an Eccentric Binary 137 9.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137 8 CONTENTS 9.2 Mass Transfer via Roche-Lobe Overflow . . . . . . . . . . . . . . . . . . . . . . . . . 137 9.2.1 Equipotential Surfaces in an Eccentric Binary Orbit . . . . . . . . . . . . . . 138 9.2.2 Keplerian Orbits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138 9.2.3 Trajectory of a Particle in the Binary Potential . . . . . . . . . . . . . . . . . 139 9.2.4 Final Outcome of Particle Trajectories . . . . . . . . . . . . . . . . . . . . . . 139 9.2.5 Comparison to the X-ray Intensity Profile . . . . . . . . . . . . . . . . . . . . 142 9.3 Mass Transfer via a Stellar Wind . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142 9.3.1 Bondi-Hoyle Accretion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145 9.3.2 Wind Accretion Profile in an Eccentric Binary . . . . . . . . . . . . . . . . . 146 9.3.3 Comparison to the X-ray Intensity Profile . . . . . . . . . . . . . . . . . . . . 147 10 Eccentric Binary Evolution 149 10.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149 10.2 Binary Evolution Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149 10.2.1 Evolution of Binary Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . 149 10.2.2 Evolution of the Donor Star Radius . . . . . . . . . . . . . . . . . . . . . . . 151 10.2.2.1 Nuclear Evolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151 10.2.2.2 Thermal Readjustment Toward Equilibrium . . . . . . . . . . . . . 152 10.2.2.3 Adiabatic Response to Mass Loss . . . . . . . . . . . . . . . . . . . 152 10.2.3 Evolution of Orbital Parameters . . . . . . . . . . . . . . . . . . . . . . . . . 152 10.2.3.1 Tidal Evolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153 10.2.3.2 Effect of Mass Transfer on the Orbit . . . . . . . . . . . . . . . . . . 154 10.2.3.3 Magnetic Braking . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155 10.2.3.4 Gravitational Radiation . . . . . . . . . . . . . . . . . . . . . . . . . 155 10.2.4 Stable Mass Transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156 10.3 Binary Evolution Code . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158 10.3.1 Code Algorithm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158 10.3.2 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159 10.4 Monte-Carlo Population Synthesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163 10.4.1 Systems Resembling Cir X-1 . . . . . . . . . . . . . . . . . . . . . . . . . . . 163 10.4.2 Predicted System Parameters for Cir X-1 . . . . . . . . . . . . . . . . . . . . 165 10.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170 11 Conclusions 171 CONTENTS A Fourier Timing Techniques 9 175 A.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175 A.2 Power Density Spectra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176 A.2.1 Leahy Normalization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176 A.2.2 Estimate of the Poisson Noise Level . . . . . . . . . . . . . . . . . . . . . . . 176 A.2.3 Fractional RMS Variability . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177 A.2.4 Estimates of Power Variance . . . . . . . . . . . . . . . . . . . . . . . . . . . 177 A.2.5 Data Gaps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178 A.2.6 Variable Number of Detectors . . . . . . . . . . . . . . . . . . . . . . . . . . . 178 A.3 Complex Cross Spectra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180 A.3.1 Time Lags . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180 A.3.2 Coherence Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180 B Standard PCA Light Curves 183 10 CONTENTS List of Figures 1-1 X-ray color-color diagrams and power density spectra typical of Z sources and atoll sources. The soft color is approximately I [3–5 keV] / I [1–3 keV] and the hard color I [6.5–18 keV] / I [5–6.5 keV]. Power spectra are shown for the horizontal, normal, and flaring branches (HB/NB/FB) of Z sources, and the island state, lower banana, and upper banana (IS/LB/UB) of atoll sources. (Figure from van der Klis 1995 [78].) . 27 2-1 Rossi X-ray Timing Explorer spacecraft. . . . . . . . . . . . . . . . . . . . . . . . . . 34 3-1 ASM light curve (1.5–12 keV; 90-s exposures) covering 1996 January – 1998 May. Phase zero is indicated by vertical dotted lines. Bars along the top axis mark the times of PCA observations (labeled A–H; see section 3.2). . . . . . . . . . . . . . . 40 3-2 Light curve and two hardness ratios for the first ten full 16.55-day cycles of Cir X-1 observed with the ASM. Each light-curve point is the average count rate (I [1.5– 12 keV]) from a single 90-s ASM exposure. The hardness ratios were obtained from 1-day averages of the 90-s intensities and are defined as HR1=I [3–5 keV] / I [1.5– 3 keV] and HR2=I [5–12 keV] / I [3–5 keV]. The vertical dotted lines indicate phase zero. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42 3-3 Folded light curve of Cir X-1 from the three ASM energy bands. The data are repeated in order to show two complete cycles. Each point is from a 90-s exposure, with a typical error bar of 1–2 c/s. The folded data spans the entire time period covered by Figure 3-1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43 11 12 LIST OF FIGURES 3-4 Left: Folded Cir X-1 light curve from the full ASM energy band (1.5–12 keV) and two hardness ratios. Right: Folded light curves from the three ASM energy bands. The data are binned into 100 phase bins per cycle, averaged in each phase bin, and duplicated to show two complete cycles. Error bars on the light curves indicate the standard deviation of the mean value derived for each phase bin, and these error estimates were propagated when computing the hardness ratios. The folded data spans the entire time period covered by Figure 3-1. . . . . . . . . . . . . . . . . . . . 45 3-5 Light curves for each 16.55-d cycle of Cir X-1 observed with the PCA. For each cycle, the phase-zero date and letter of the study (see Table 3.1) are listed. The intensity is for the full PCA energy range, and a conversion factor of 1 Crab ≈ 2600 counts s−1 PCU−1 has been used. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3-6 PCA light curves in two energy bands (16-s bins), and the ratio of intensity in the hard band to that in the soft band, for the observation on 1996 April 7 (study B), spanning 15 ks. The lowest intensity in both bands during dips is well above the background level (∼30 c/s and ∼45 c/s for 2–7.4 keV and 7.4–28 keV respectively, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . not subtracted). 47 48 3-7 PCA light curves in two energy bands (16-s bins), and the ratio of intensity in the hard band to that in the soft band, for the observations on 1996 September 24 – October 8 (study D). Due to some observations with a PCU not operating, all points in this plot were derived from PCUs 0, 1, & 2 and adjusted by 5/3 for comparison with other figures in this section. Phase zeros occurred at days 346.782 and 363.328. 49 3-8 PCA light curves in two energy bands (16-s bins; 13 kilocounts/s=1.0 Crab for the combined 2-33 keV band), and the ratio of intensity in the hard band to that in the soft band, for the observation on 1997 September 20 (part of study H), spanning 7000 s. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51 4-1 PCA light curves in two energy bands (16-s bins), and the ratio of intensity in the hard band to that in the soft band, for observations 1–11 of 1996 March. Phase zero corresponds to day 148.227. The intensity is well above the background level in both bands (∼25 c/s and ∼40 c/s for 2–6.3 keV and 6.3–24 keV respectively, not subtracted) even during the dips. Observation 12 is omitted since it was carried out with only three PCUs and a different gain setting. . . . . . . . . . . . . . . . . . . . 55 LIST OF FIGURES 4-2 Sample PCA light curves (2–60 keV) comparing 2000-s segments showing (a) strong variability and dips near phase zero (phase 0.98; Obs. 1), (b) a brighter portion of the bright flaring state three days later (phase 0.15; Obs. 2), and (c) a typical observation away from phase zero (phase 0.50; Obs. 6). Background (∼100 c/s) is not subtracted . . . . . . . . . . . . . . . . . . . . . and no deadtime corrections have been made. 13 56 4-3 Power density spectra from seven RXTE orbits away from phase zero, offset downward at decade intervals. (The ordinate scale applies to the top curve.) The curves are ordered by the frequency of the 1.3–12 Hz QPO peak and are labeled with orbital phase. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58 4-4 Flat-top power level (a) and QPO frequency (b) as a function of orbital phase over a single orbital cycle. (c) Frequency of the narrow low-frequency QPO and the broad high-frequency QPO vs. the flat-top power level. (d) Cut-off frequency of the flat- topped power vs. flat-top power level. . . . . . . . . . . . . . . . . . . . . . . . . . . 59 5-1 RXTE ASM light curves (1.5–12 keV) for three 16.55 d cycles of Cir X-1 showing different flaring profiles. Each intensity point corresponds to a 90-s exposure by one of the three ASM cameras, and the hardness ratio (HR), defined as the ratio of counting rates for 5–12 keV to 3–5 keV, is shown in one-day averages. The 3–5 keV to 1.5–3 keV hardness ratio exhibits very similar behavior and is not shown here. The intensities are for Cir X-1 after background and other sources in the field of view have been subtracted. The Crab nebula yields ∼75 c/s. Vertical dashed lines indicate phase zero based on the radio ephemeris of Stewart et al. 1991. Day zero corresponds to (a) 1997 April 23.87 (b) 1996 August 2.14 and (c) 1997 February 16.69. For cycle (c), the intensity ranges (I [2.0–18 keV]) seen in the eight RXTE PCA observations (I–VIII in time order) are also shown. . . . . . . . . . . . . . . . . . . . . . . . . . . 63 5-2 PCA light curves (2–18 keV, 5 PCUs) and hardness ratios (HR = I [6.3–13 keV] / I [2.0–6.3 keV]) in 16 s time bins for the eight observations made during 1997 Febru- ary 18 – March 4. A count rate of 13 kcts/s ≈ 1.0 Crab. The data gaps in Obs. VI were longer than as shown here; the second segment of the observation has been shifted left by 4000 s and the third segment by 5000 s. These data were used to construct the hardness-intensity diagram in Figure 5-3. The association with specific regions of that diagram is indicated below each light curve. . . . . . . . . . . . . . . . . . . . . 65 14 LIST OF FIGURES 5-3 Color-color diagram (a) and hardness-intensity diagram (b) for all eight observations (I–VIII). In the CD, soft color is defined as I [4.8–6.3 keV] / I [2.0–4.8 keV] and hard color as I [13–18 keV] / I [8.5–13 keV]. In the HID, the intensity, I [2.0–18 keV], is from all five PCUs and the hardness ratio is a "broad" color: I [6.3–13 keV] / I [2.0– 6.3 keV]. Each point corresponds to 16 s of data. Background has been subtracted, but it does not affect the intensity or soft color and only slightly affects the hard color. The three insets in the CD separate overlapping points from observations I–V. The HID track for each observation has been divided into three regions (1–3) for timing analysis. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66 5-4 Averaged and rebinned power density spectra (2–32 keV) for each of the three HID regions for each observation. Poisson noise has been subtracted from each PDS (see text). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67 5-5 Averaged and rebinned power density spectra for HID region VIII-3 in three energy bands. A harmonic peak of the 7.6 Hz QPO is clearly visible in the high-energy channel (c). The broad high-frequency peak, most clear in (b), occurs near ∼100 Hz in this observation. The low-frequency noise cuts off less sharply as energy increases. 69 5-6 Centroid frequency of the QPOs versus intensity (I [2.0–18 keV]). A filled circle rep- resents the narrow QPO and a filled square represents the broad ∼4 Hz QPO (all points below 5 Hz are the broad QPO). Unfilled circles and squares indicate the approximate frequency of a knee or very weak peak that may be associated with the narrow and broad QPO respectively. Error bars on frequency measurements (filled points only) represent 90% confidence intervals for a single parameter (∆χ2=2.7). In many cases, the error bar for the QPO frequency is smaller than the plot symbol. . . 71 5-7 Rms amplitude of QPOs versus photon energy. Panel (a) shows the rms for the narrow QPO at 7.2 Hz (solid dot) and (b) for an example of the broad 4 Hz QPO (solid box). In panel (c), unfilled circles and boxes indicate the rms amplitude of a component forming a knee or very weak peak that may be associated with the narrow and broad QPOs respectively. The broad QPO points have been offset slightly to the right in energy for clarity. Errors on QPO amplitudes represent 90% confidence intervals. . . 72 5-8 Hardness-intensity diagram showing QPO properties for the regions from Figure 5-3b. The frequency of the 6.8 to 32 Hz QPO is labeled (in Hz) beside each region where it is present. Parenthesized frequencies indicate that this component was unpeaked, i.e., a knee. Letters indicate the strength of the broad 4 Hz QPO: S-strong, M-medium, and W-weak or unpeaked. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74 LIST OF FIGURES 15 5-9 Broad-color and hard-color CDs and HIDs for observations V and VI. In all four diagrams, Obs. V is in the lower left and Obs. VI in the upper right. The CD and HID tracks for Obs. V both turn upward at the left end in the broad-color diagrams (a,b) but turn downward in the hard-color diagrams (c,d). In the HIDs, presence of the 32 Hz HBO, 4 Hz NBO, and VLFN is indicated along the branches. An apparent shift of the normal branch between observations V and VI is labeled in (b). The intensity, I [2.0–18 keV], is from all five PCUs. The soft color is defined as I [4.8– 6.3 keV] / I [2.0–4.8 keV], the broad color as I [6.3–13 keV] / I [2.0–6.3 keV], and the hard color as I [13–18 keV] / I [8.5–13 keV]. Each point corresponds to 16 s of background-subtracted data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77 6-1 Light curves in four energy bands and two hardness ratios for PCA observations of Cir X-1 from 1997 June 10–20, covering a 10-day period around phase zero (φ = 0). The intensities at the beginning of these observations (day 610) are typical "quiescent" levels. Each point represents 16 s of background-subtracted data from PCUs 0, 1, and 2. Ratios of the intensities in the four bands produce soft (I [4.8–6.3 keV] / I [2.0– 4.8 keV]) and hard (I [13–18 keV] / I [6.3–13 keV]) hardness ratios. Segments labeled A, B, C & D were used for spectral studies. . . . . . . . . . . . . . . . . . . . . . . 83 6-2 Color-color and hardness-intensity diagrams from PCA observations during 1997 June 10–21 (the entire period covered by Figure 6-1). Each point represents 16 s of background-subtracted data from PCUs 0, 1, and 2. . . . . . . . . . . . . . . . . . . 84 6-3 Color-color and hardness-intensity diagrams from time segments "C" (left panels) and "D" (right panels) of Figure 6-1 (1997 June 13.625–14.125 and 1997 June 17.075– 17.600 respectively). Each point represents 16 s of background-subtracted data from PCUs 0, 1, and 2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85 6-4 Hardness-intensity diagram from time segment C (1997 June 13.625–14.125; day 12.625–13.125). The HID track has been divided into 20 regions from which power density spectra and energy spectra were constructed. . . . . . . . . . . . . . . . . . 87 6-5 Top: light curves in 3 energy channels and bottom: total 2–18 keV light curve, broad color, and HID regions for time segment C (1997 June 13.625-14.125; see Figure 6-1). Based on HID region numbers, the predominant spectral branch is identified for each portion of the data. Absorption dips (omitted from HID regions) are clearly identified by decreased intensity coupled with upward broad-color spikes. . . . . . . . . . . . . 88 16 LIST OF FIGURES 6-6 Averaged and rebinned power density spectra (2–32 keV) for each of the 20 regions along the HID track in Figure 6-4. The estimated Poisson noise level has been sub- tracted from each PDS. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90 6-7 Cir X-1 energy spectra (2.5–25 keV; PCU 0 only) for several regions on each segment of the HID track in Figure 6-4. The region numbers within the figures are ordered vertically to match the relative intensities at the low and/or high-energy ends of the spectra. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92 6-8 Spectrum A and model (histogram) consisting of a disk blackbody and blackbody (see Table 6.1). The residuals show a peak at 6–7 keV that may be due to an iron emission line. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95 6-9 Fitted energy spectra (2.5–25 keV; PCU 0 only) and model components (disk black- body and blackbody) for HID regions 1, 7, and 11 of Figure 6-4. The disk blackbody dominates at low energy and the blackbody dominates at high energy. Fit parameters are listed in Table 6.2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98 6-10 Fitted energy spectra (2.5–25 keV; PCU 0 only) and model components (disk black- body and blackbody) for HID regions 13, 15, and 20 of Figure 6-4. The blackbody only contributes to the high-energy end of the spectrum in region 13, and has faded away entirely in regions 15 and 20. Fit parameters are listed in Table 6.2. . . . . . 99 6-11 Spectra from HID regions 17 and 20 (crosses) and fitted models (histograms) consist- ing of a disk blackbody and blackbody. The ratio of residuals to error bars shows a narrow edge or line-like feature above 10 keV. . . . . . . . . . . . . . . . . . . . . . 100 7-1 Light curves in four energy channels and two hardness ratios for PCA observations of Cir X-1 from 1996 September 20–22, covering a two-day period around phase zero (φ = 0). Three dips have been identified for further study. Each point represents 16 s of background-subtracted data from PCUs 0, 1, and 2. Ratios of the intensities produce soft (I [4.8–6.3 keV] / I [2.0–4.8 keV]) and hard (I [13–18 keV] / I [6.3–13 keV]) hardness ratios. The intensity levels of the segment at day ∼347.2 (after Dip 1) are close to the level in each band during the quiescent phases of the orbit. . . . . . . . 105 7-2 Light curves in four energy channels and two hardness ratios spanning 4000 s including Dip 1. Time zero corresponds to day 347.041 in Figure 7-1. The intensity and hardness ratio points are the same as in Figure 7-1. Energy spectra were extracted from two 96- s time segments (A and B), indicated by dotted vertical lines, and four 16-s segments (a–d) indicated by diamonds. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106 LIST OF FIGURES 17 7-3 Light curves in four energy channels and two hardness ratios spanning 4000 s including Dip 2. Time zero corresponds to day 347.254 in Figure 7-1. The intensity and hardness ratio points are the same as in Figure 7-1. Energy spectra were extracted from two 304-s time segments (A and B), indicated by dotted vertical lines, and four 16-s segments (a–d), indicated by diamonds. . . . . . . . . . . . . . . . . . . . . . . . . . 107 7-4 Light curves in four energy channels and two hardness ratios spanning 4000 s including Dip 3. Time zero corresponds to day 347.325 in Figure 7-1. The intensity and hardness ratio points are the same as in Figure 7-1. Energy spectra were extracted from six 16-s time segments (a–f), indicated by diamonds. . . . . . . . . . . . . . . . . . . . 108 7-5 Color-color and hardness-intensity diagrams for the last 1716 s of data in Figure 7-4, during which the intensity gradually transitioned from the non-dip baseline to the bottom of a dip. Intensity is I [2.0–18 keV], and the hardness ratios are defined as soft color: I [4.8–6.3 keV] / I [2.0–4.8 keV], hard color: I [13–18 keV] / I [6.3–13 keV], and broad color: I [6.3–18 keV] / I [2.0–6.3 keV]. Each point represents 16 s of background- subtracted data from PCUs 0, 1, and 2. . . . . . . . . . . . . . . . . . . . . . . . . . 110 7-6 Color-color and hardness-intensity diagrams for the first three segments in Figure 7-1 (day 346.31–346.52), during which Cir X-1 was in an extended low/dip state. Intensity is I [2.0–18 keV], and the hardness ratios are defined as soft color: I [4.8–6.3 keV] / I [2.0–4.8 keV], hard color: I [13–18 keV] / I [6.3–13 keV], and broad color: I [6.3– 18 keV] / I [2.0–6.3 keV]. Each point represents 16 s of background-subtracted data from PCUs 0, 1, and 2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111 7-7 Top: Spectral fits for 96-s segments prior to Dip 1 (spectrum A) and during Dip 1 (spectrum B). Bottom: same spectra and model (see text), but with a Gaussian emission-line component included to fit the peaked residuals near 6.5 keV. The data shown are from PCU 0 only. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114 7-8 Top: Spectral fits for 304-s segments prior to Dip 2 (spectrum A) and during Dip 2 (spectrum B). Bottom: same spectra and model (see text), but with a Gaussian emission-line component included to fit the peaked residuals near 6.5 keV. The data shown are from PCU 0 only. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115 7-9 Spectral fits for four 16-s segments (a–d) during the decline into Dip 1. Only the column density of the bright component varies between the four jointly-fit curves. The data shown are from PCU 0 only. . . . . . . . . . . . . . . . . . . . . . . . . . . 119 18 LIST OF FIGURES 7-10 Spectral fits for four 16-s segments (a–d) during the decline into Dip 2. Only the column density of the bright component varies between the four jointly-fit curves. The data shown are from PCU 0 only. . . . . . . . . . . . . . . . . . . . . . . . . . . 120 7-11 Spectral fits for six 16-s segments (a–d) during the decline into Dip 3. Only the column density of the bright component varies between the six jointly-fit curves. The data shown are from PCU 0 only. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121 8-1 X-ray, IR, and radio light curves of Cir X-1 from 1996 May 7–17. Phase zero is indicated by the vertical dotted line at day 214.413 (1997 May 11.413). PCA intensity is from the 2–32 keV band (16-s bins) and ASM intensity is from 2–12 keV (90-s exposures). Errors on the PCA data are negligible. Typical error bars are shown for the ASM, IR, and radio data: errors on the ASM data are typically 2–5 c/s, on the IR data are typically about 0.03 magnitudes (worse on days 214 & 215), and on the radio data are about 30 mJy. The average quiescent level of 29 mJy has been subtracted from the radio data. (Data provided by I. Glass (IR), G. Nicolson (radio), and the RXTE /ASM team.) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126 8-2 X-ray and IR light curves of Cir X-1 from day 212.93–213.19. X-ray intensity is from the 2–32 keV PCA band (16-s bins). A typical IR error bar of +/−0.03 magnitudes is shown. (IR data courtesy of I. Glass.) . . . . . . . . . . . . . . . . . . . . . . . . 129 8-3 X-ray (PCA and ASM) and radio (5.0 GHz) light curves of Cir X-1 from 1997 May 7 – August 25. Phase zero is indicated by the vertical dotted lines. PCA intensity is from the 2–32 keV band (16-s bins) and ASM intensity is from 2–12 keV (90-s exposures). Errors on the PCA data are negligible and errors on the ASM data are typically 2– 5 c/s. For each 16.55-d cycle, the average radio flux for the phase interval 0.3 to 0.9 of that cycle was subtracted from the data; the bias removed was typically 50–60 mJy and a typical error bar is shown. (The radio data was provided by G. Nicolson, and ASM data by the RXTE /ASM team. The PCA data are from studies F and G; see Table 3.1.) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131 8-4 Infrared spectrum of Cir X-1 from ANU on 1997 June 20 (about phase 0.5), showing He I (2.06 µm) and H Brackett γ (2.165 µm) emission lines. This spectrum was compiled from the sum of 16 120-s exposures. (Figure courtesy of H. Johnston.) . . 133 8-5 Optical spectrum of Cir X-1 from AAT on 1997 June 4 (about phase 0.5), showing a prominent Hα emission line, as well as weaker He I(6678 A) and He I(7065 A) lines. (Figure courtesy of H. Johnston.) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134 LIST OF FIGURES 19 9-1 Panels (a) and (c): stop time of particles in a system with e=0.5 (and P=16.55 d) versus time of release from donor (1 M⊙), with the final outcome (recapture or entry into disk) indicated. Panels (b) and (d): normalized rate of particles entering the accretion disk versus their arrival time relative to periastron. The top panels, (a) and (b), are for an initial radial velocity from the surface of 30 km/s, and the bottom panels, (c) and (d), are for 60 km/s. In (a) and (c), the solid diagonal line has a slope of unity and indicates zero elapsed time between release and the final outcome; the horizontal dashed line indicates a final outcome occurring at periastron. . . . . . . . 141 9-2 Panels (a) and (c): stop time of particles in a system with e=0.8 (and P=16.55 d) versus time of release from donor (1 M⊙), with the final outcome (recapture or entry into disk) indicated. Panels (b) and (d): normalized rate of particles entering the accretion disk versus their arrival time relative to periastron. The top panels, (a) and (b), are for an initial radial velocity from the surface of 50 km/s, and the bottom panels, (c) and (d), are for 100 km/s. In (a) and (c), the solid diagonal line has a slope of unity and indicates zero elapsed time between release and the final outcome; the horizontal dashed line indicates a final outcome occurring at periastron. . . . . 143 9-3 Trajectory of a particle moving in an eccentric binary system (Mdon=1 M⊙, e=0.5), where the particle starts with an initial radial velocity of 60 km/s off the surface of the donor at a time t0 relative to periastron. The plot is in a non-rotating frame with the CM at the origin (+). The location of the particle, donor star, and neutron star at t0 are indicated by the diamond, asterisk, and solid dot respectively. The subsequent motion of the particle until enters the accretion disk or is recaptured is shown as a solid curve, and the motion of the donor and neutron star in their orbits during that time is indicated by a dashed or dotted ellipse segment respectively. . . . . . . . . . 144 9-4 Wind accretion rate versus orbital phase for three eccentricities and wind speeds for a 5 M⊙ donor. Each curve is normalized by its peak value; the actual wind capture fraction (relative to the mass-loss rate of the donor) decreases significantly with higher velocity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148 10-1 Results of a binary evolution calculation for a system started with the donor star just filling its critical potential lobe, with initial values: Mdon = 2 M⊙, a = 9.34 R⊙, and e = 0.6. The donor remains in contact with its critical potential lobe throughout the evolution; thus, the curve for Rlobe coincides with the curve for Rdon. A solid dot on some of the curves marks the time corresponding to an orbital period of 16.55 days. 161 20 LIST OF FIGURES 10-2 Results of binary evolution calculation for a system started with the donor underfilling its critical potential lobe, with initial values: Mdon = 1 M⊙, a = 27 R⊙, and e = 0.9. 162 10-3 Eccentricity and mass transfer rate versus time for binaries with initial values Mdon = 2 R⊙, e = 0.9, and semimajor axes (a) between 37 and 128 R⊙. . . . . . . . . . . . 164 10-4 Histograms of number of binaries versus initial Mdon and e which eventually reach P = 16.55 d and 1×10−9 M⊙/yr < M < 1×10−7 M⊙/yr. The y-axis normalization is arbitrary. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166 10-5 Histograms of the number of binaries versus various system parameters at the time when systems resemble Cir X-1, i.e., P = 16.55 d and 1×10−9 M⊙/yr < M < 1×10−7 M⊙/yr. The y-axis normalization is arbitrary. . . . . . . . . . . . . . . . . . 167 10-6 Optical and infrared absolute magnitudes (V,R,I,J,K) of the donor star in systems that resemble Cir X-1, i.e., those from Figure 10-5. The y-axis normalization of the histograms is arbitrary. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169 List of Tables 2.1 Approximate energy ranges for eight PCA bands (A–H) defined in terms of absolute (0–255) and Standard2 (0–128) channels. Gain epoch 1 includes all observations before before 1996 March 21, epoch 2 1996 March 21 – April 15, and epoch 3 after 1996 April 15. Energy values shown are derived from the full PCA with all detectors and layers added. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37 3.1 PCA observations of Cir X-1 (1996–1997). . . . . . . . . . . . . . . . . . . . . . . . . 46 4.1 March 1996 PCA observations of Cir X-1 . . . . . . . . . . . . . . . . . . . . . . . . 54 5.1 PCA observations of Cir X-1 during 1997 February 18 – March 4. . . . . . . . . . . . 62 6.1 Fit parameters for spectra A & B for four models. . . . . . . . . . . . . . . . . . . . 94 6.2 Fit parameters for HID regions 1–20 using a model consisting of a disk blackbody and a blackbody. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97 7.1 Top group: joint fit parameters of the bright and faint components of spectra from outside (A) and inside (B) dips 1 and 2. Middle group: the same, but with an added Gaussian line. Third group: joint fit parameters for the four to six 16-s spectra for dips 1, 2, and 3. The absorption parameters (N (1) H ) for the bright component are given in Tables 7.2 (A–B) & 7.4 (a-f) and the iron line parameters are given in Table 7.3. 116 7.2 Effective hydrogen column density of the bright component outside (A) and inside (B) Dips 1 and 2 (from fits with and without a ∼6.5 keV Gaussian emission line). . 116 7.3 Gaussian emission-line parameters for the joint fits of spectra outside and inside dips (A–B) and during entry into a dip (a–d, a–f). In each case, the same Gaussian line parameters were used for each spectrum in a joint fit. . . . . . . . . . . . . . . . . . 118 21 22 LIST OF TABLES 7.4 Effective hydrogen column density responsible for the variable absorption in the bright component, due to photo-electric absorption and Thompson scattering, during 16-s spectra from Dips 1, 2, and 3. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122 8.1 Multi-frequency Observations of Cir X-1 from 1996 May. Dates in parentheses were cancelled due to poor weather conditions. . . . . . . . . . . . . . . . . . . . . . . . . 127 8.2 Multi-frequency Observations of Cir X-1 from 1997 June. Dates in parentheses were cancelled due to poor weather conditions. . . . . . . . . . . . . . . . . . . . . . . . . 130 8.3 Infrared JHKL photometry measurements from 1997 June 14 & 15. Errors are +/− 0.03, except on the L magnitude the error is +/− 0.05. . . . . . . . . . . . . . 132 Chapter 1 Introduction 1.1 X-ray Astronomy X-rays are strongly absorbed by the Earth's atmosphere, making ground-based observations of astronomical X-ray sources impossible. Thus, X-ray astronomy did not begin until rockets and high-altitude balloons became available to carry instruments above a significant fraction of the at- mosphere. The first celestial X-ray source (other than the Sun), Scorpius X-1, was discovered in 1962 using a rocket-borne detector [22]. Additional X-ray sources were first detected by balloon and rocket experiments during the 1960's. Many more sources were discovered with the launch of the first astronomical X-ray satellite, Uhuru, in 1970. This was followed by a series of satellites in the 1970's, including Vela 5, Copernicus, Ariel V, SAS-3, OSO-7, OSO-8, COS-B, HEAO-1, Einstein, and Hakucho. In the 1980's Tenma, EXOSAT , Ginga , and Mir-Kvant were launched. Missions in the 1990's include Granat, ROSAT and ASCA , and most recently, RXTE and BeppoSAX. In the next few years, several major X-ray spectroscopy and imaging missions will be launched, including AXAF, XMM, and Spectrum-X- Gamma. 1.2 X-ray Binaries 1.2.1 Overview Most of the bright celestial X-ray sources are thought to consist of a neutron star or black hole accreting matter from a binary companion star (see the book X-ray Binaries [43] for a thorough review of the subject). The gravitational potential energy released when matter falls onto these 23 24 CHAPTER 1. INTRODUCTION highly compact objects produces high-luminosity radiation (of order 1037–1038 erg/s) from a small area (∼10 km radius), resulting in an effective temperature of 106–107 K, i.e., primarily X-ray wavelengths (λ ∼0.1–100 A, E ∼0.1–100 keV). The binary orbital period can be well established through Doppler studies of X-ray pulsars (see section 1.2.2 below). The period is sometimes detected through modulation of the intensity in X-ray, optical, or other frequency bands. X-ray binaries have periods as short as minutes to as long as months. When the radial component of the orbital velocity of the companion is also measured, from Doppler-shifted lines in the optical spectrum, a lower limit to the mass of the compact object (the mass function) can be derived. A more exact estimate requires knowledge of the mass of the companion (e.g. measured from Doppler shifts of pulsed X-ray emission) and the inclination angle of the orbit relative to the line of sight. Even a lower limit can be very informative, since degeneracy pressure in a neutron star is not expected to be able to support more than about 3 times the mass of the sun (M⊙). Systems with a mass function in excess of 3 M⊙ are considered strong candidate black-holes. X-ray binaries are generally classified into two groups, based on the spectral type (observed or inferred) of the companion star. High mass X-ray binaries (HMXBs) have O or B-star companions with masses from several to about 40 M⊙, while low mass X-ray binaries (LMXBs) have companions later than type A (or even a white dwarf), with masses usually about 1 M⊙ or less. The two classes are believed to be produced via different evolutionary pathways. 1.2.2 HMXBs In HMXBs, the OB star can have a substantial radiative-driven stellar wind, which removes 10−10– 10−6 M⊙/yr (1 M⊙/yr = 6.30×1025 gm/s) with a terminal velocity up to 2000 km/s. The orbit of the neutron star or black hole can bring it close enough to the OB star so that it can capture a significant fraction of the material in the wind and thus power the X-ray source. Because the compact object is immersed in the wind material, photo-electric absorption often attenuates the X-ray intensity. In some HMXBs, Roche-lobe overflow also contributes to (or dominates) the total mass transfer (see section 1.2.3 on LMXBs). Since OB stars are quite bright in optical/UV light, the optical counterparts of HMXBs are often easily identified. The spindown rate of radio (rotation-powered) pulsars indicates a high surface magnetic field of (B > ∼1012 G). Many HMXBs are X-ray pulsars and are also thought to contain a neutron star with a high magnetic field. Such objects channel accreting material along the magnetic field lines to the magnetic poles. If the spin and magnetic axes are misaligned, the X-ray beam will rotate (like a lighthouse beam). If the beam crosses our line of sight, the system will be observed as an accretion- 1.2. X-RAY BINARIES 25 powered X-ray pulsar (in contrast to rotation-powered radio pulsars). The accreted matter produces a torque that can slow or increase the spin rate (and pulse frequency) of the compact object. In many systems an accreting white dwarf can be rule out because the rate of change of the pulse period is too high for an object with a relatively high moment of inertia (compared to a much smaller neutron star) such as a white dwarf [31]. Pulsations are also assumed to require a surface, allowing black holes to be ruled out in systems that pulse. Thus, X-ray pulsars are well established as highly-magnetized accreting neutron stars. 1.2.3 LMXBs When the companion star in an X-ray binary lacks a significant stellar wind, as is the case for most LMXBs and some HMXBs, mass transfer generally occurs though overflow of the critical equipotential surface (Roche lobe) of the star. Matter transferred in this manner is driven through the inner Lagrange point between the stars and has a high specific angular momentum. The material forms an accretion disk around the compact object and loses angular momentum though viscosity until falling onto the surface (or being expelled). X-ray heating of the disk and companion dominates the optical light from the donor star, often making it difficult to determine the spectral type of the companion. Some LMXBs exhibit flares known as type I X-ray bursts. These bursts typically have rise times from less than a second to ∼10 s and slower decay times, in the range of ∼10 s to minutes. Spectral evolution during bursts suggests a sharp rise in temperature at the onset of a burst and gradual cooling during the decay. This behavior is consistent with thermonuclear ignition of accreted matter that has accumulated on the surface of a neutron star. Because a black hole has no surface, type I X-ray bursts provide strong evidence that a system contains a neutron star. Most persistently bright LMXBs are believed to contain low magnetic field neutron stars (B < ∼ 109 G) since they generally show X-ray bursts and do not show pulsations (a higher magnetic field would funnel matter along the field lines onto the poles, where the material would be burned continuously rather than in bursts, and might produce pulsations due to the non-uniform emission from the surface [30].) Neutron-star LMXBs can be classified as either "Z" or "atoll" sources. These categories are based on the correlated X-ray timing and spectral properties of these sources [27, 78] (see Fig- ure 1-1). The fast timing properties (on time scales of seconds or less) of LMXBs and black-hole candidates are typically studied by computing Fourier power density spectra (PDSs), which often show several broad-band noise components and sometimes peaks due to quasi-periodic intensity oscillations (QPOs). The spectral properties can be examined by forming ratios of count rates in 26 CHAPTER 1. INTRODUCTION different energy bands (X-ray hardness ratios or "colors"). The six LMXBs known as Z sources are: Sco X-1, GX 349+2, GX 17+2, Cyg X-2, GX 5-1, and GX 340+0. These sources typically trace a "Z" pattern in color-color and hardness-intensity diagrams (CDs/HIDs), where X-ray "color" or "hardness" refers to the ratio of intensity in one energy band to that in a lower-energy band, while atoll sources trace a single arc (see Figure 1-1). Z sources show two types of quasi-periodic oscillations: along the top "horizontal branch" (HB) of the Z pattern are 13–60 Hz horizontal-branch oscillations (HBOs) which increase in frequency moving to the right on the diagram, and on the diagonal "normal branch" (NB) are ∼6 Hz normal- branch oscillations (NBOs). In some Z sources, the normal-branch oscillations are observed to evolve into 6–20 Hz QPOs (FBOs) on the lower "flaring branch" (FB). The broad-band noise components of atoll sources also show correlations with position in the spectral diagrams. The most widely held model for these two classes of LMXBs proposes two physical differences between them: Z sources have stronger magnetic fields (1–5×109 G) and higher mass accretion MEdd ∼ 3×10−8 M⊙/yr, or luminosity LEdd ∼ 1.8×1038 erg/s) rates (near the Eddington limit of ∼108 G and luminosities between 10−3 LEdd and a few 10−1 LEdd) [77]. The suggestion that Z sources have stronger magnetic fields than atoll sources was than atoll sources (with magnetic fields < based largely on the existence of QPOs in Z sources and their general absence in atoll sources. In the magnetospheric beat-frequency model for Z-source horizontal-branch oscillations, the QPOs are the result of clumps of matter at the inner edge of the accretion disk (at the magnetospheric radius) which fall in along magnetic field lines to the surface with a frequency that is the difference between the Keplerian frequency at the magnetospheric radius and the neutron-star spin frequency [1, 37]. In this picture, the absence of such QPOs in atoll sources suggests lower fields so that the disk might extend almost to the neutron star surface. Near-Eddington accretion rates in Z sources are thought to be responsible for the QPOs on the normal and flaring branches, where a radial accretion inflow occurs with oscillations in the optical depth [20]. Recently kilohertz QPOs (∼300–1200 Hz) have been discovered in a number of Z and atoll sources with the Rossi X-ray Timing Explorer (RXTE ). These QPOs possess many interesting properties (see van der Klis [79] for a recent review). They often appear in pairs that shift together in fre- quency. In some sources, nearly coherent oscillations have been observed during type I bursts at a frequency equal to or double the separation between peaks in kHz QPO pairs. This suggests that a beat-frequency mechanism is responsible, e.g., if the higher-frequency QPO peak is related to the Keplerian frequency at the inner edge of the accretion disk and the oscillations during bursts reflect the spin frequency (or twice the spin frequency) of the neutron star, then the lower frequency QPO would occur at the beat-frequency between the former two frequencies. One such beat-frequency is 1.2. X-RAY BINARIES 27 Figure 1-1: X-ray color-color diagrams and power density spectra typical of Z sources and atoll sources. The soft color is approximately I [3–5 keV] / I [1–3 keV] and the hard color I [6.5–18 keV] / I [5–6.5 keV]. Power spectra are shown for the horizontal, normal, and flaring branches (HB/NB/FB) of Z sources, and the island state, lower banana, and upper banana (IS/LB/UB) of atoll sources. (Figure from van der Klis 1995 [78].) 28 CHAPTER 1. INTRODUCTION the magnetospheric beat-frequency model, previously applied to Z-source horizontal branch oscil- lations (see above). This model cannot be correct for both types of QPOs since sometimes HBOs and kHz QPOs are observed simultaneously; other beat-frequency models propose a different radius associated with the kHz QPOs, e.g., the sonic-point model [48]. Important constraints on neutron star masses, radii, and equations of state can be derived from these oscillations, since they may reflect neutron-star spin frequencies and frequencies of orbits very close to the neutron-star surface. The recent discovery of coherent 2.49-ms pulsations in RXTE observations of the source SAX J1808.4-3658 demonstrates the existence of millisecond X-ray pulsars [85, 13]. Millisecond X-ray pulsars are widely held to be the progenitors of millisecond radio pulsars. Angular momentum transfered to the neutron star by accreting matter is thought to be the mechanism that has spun-up millisecond radio pulsars, and such sources are expected to appear as X-ray pulsars during the mass accretion phase. 1.3 Circinus X-1 1.3.1 Intensity Cycle The X-ray binary Circinus X-1 is unique in its complex temporal and spectral variability. A 16.55 day cycle of flaring is observed in the X-ray [32] as well as optical [50], IR [23], and radio bands [83]. The high degree of stability of the period of this cycle is evidence that it is the orbital period. The periodic behavior is believed to be the result of a highly eccentric binary orbit, in which mass transfer occurs only near periastron, leading to intermittent obscuration and flaring near phase zero [52, 54, 12]. The X-ray profile and average intensity of the 16.55 day cycle has varied considerably over time scales of years. The phase zero behavior has been described as a high-to-low transition, a low-to- high transition, or complex dips and flares (see e.g., [16, 68, 54, 62]). This has lead some authors to suggest that disk precession or apsidal motion with a time scale of several years plays an important role in the observed profile [52, 54]. 1.3.2 X-ray Bursts Eight X-ray bursts detected during an EXOSAT observation of Cir X-1 somewhat resembled type I bursts, but had profiles which left their identification inconclusive [71]. Three additional X-ray bursts seen during a later EXOSAT observation provided more convincing evidence for type-I behavior, and thus demonstrated that Cir X-1 is a neutron star with a weak magnetic field [72]. This in turn 1.3. CIRCINUS X-1 29 suggests that Cir X-1 is an LMXB. Further type I bursts have not been observed from Cir X-1 since the EXOSAT discovery, possibly because the source intensity has been higher during subsequent observations. (At higher accretion rates, the material is expected to burn continuously on the surface rather than accumulate, as is necessary for type I bursts [30].) 1.3.3 Rapid X-ray Variability The rapid X-ray variability of Cir X-1 at times resembles that of both "atoll" and "Z" low-mass X-ray binaries (LMXBs) as well as black-hole candidates [54]. Quasi-periodic oscillations (QPOs) were reported at 1.4 Hz, 5–20 Hz, and 100–200 Hz in EXOSAT observations of Cir X-1 in a bright state [69, 70], but other observations at lower intensity showed no such QPOs [54]. Based on these data, it has been suggested that Cir X-1 is an atoll source that can uniquely reach the Eddington accretion rate and exhibit normal/flaring branch QPOs at 5–20 Hz [54, 77]. Similar QPOs were observed in Ginga observations of Cir X-1 [45]. A significant portion of this current project is devoted to studying the rapid X-ray variability of Cir X-1. The current results demonstrate behavior associated with all three branches of Z-sources (horizontal, normal, and flaring), including horizontal and normal branch QPOs. No evidence for atoll-source behavior is found in the current observations; this might be related to the currently bright state of the source. 1.3.4 Distance Neutral hydrogen clouds orbiting in the Galaxy produce 21-cm absorption features in the radio spectrum of sources observed through such clouds. The absorption lines are Doppler shifted by an amount equal to the radial velocity of clouds in their orbits around the Galaxy. The maximum velocity observed in these features indicates the radius of the Galactic orbit tangential to the line of sight, thus putting a lower limit on the distance to the absorbed source. HI absorption features in the spectrum of Cir X-1 extend to as much as −90 km/s, which puts a lower limit on the distance of Cir X-1 at about 80% of the distance from the Sun to the center of the Galaxy (∼8–10 kpc, where 1 kpc≡3.086×1021 cm), or 6.4–8 kpc [25, 24]. 1.3.5 Young Runaway Binary Radio images show that Cir X-1 is embedded in a synchrotron nebula which trails back toward the nearby (∼ 1/2 ◦, or 70 pc at 8 kpc) supernova remnant (SNR) G 321.9-0.3. Inside the nebula, there is a compact source at the position of Cir X-1, located at (J2000) RA 15 h 20 m 40.99 s, Dec. -57◦ 09′ 59.91′′. Jet-like radio structures emanate outward from the compact source for about 30 arcsec 30 CHAPTER 1. INTRODUCTION before curving back several arcmin toward the SNR. This suggests that the system is a runaway from the explosion that created the neutron star ∼30,000-100,000 years ago [67]. The velocity required for Cir X-1 to travel ∼70 pc in this time is 700–2200 km/s. Although this velocity is quite high for a binary system, it is not unprecedented. Furthermore, there is spectral evidence that may also support a high velocity for the system; this is discussed in Chapter 8 in conjunction with our multi-frequency observations, as are the prospects for a measurement of the angular velocity (proper motion) of Cir X-1. The possibility of a young age makes the object highly interesting from the standpoint of binary- system evolution. For example, mass transfer in a long-period LMXB (P > ∼days) is usually initiated when the donor star evolves to become a Roche-lobe filling giant star. If the primary star has only recently collapsed into a neutron star, this scenario implies that the secondary (probably less massive) star would be close behind in its evolution; this may be an unlikely scenario. Based on binary evolution calculations (Chapter 10), I show in this thesis that the system need not be unusually young to have a high eccentricity and substantial accretion rate, and in fact would probably have to be much older than the supernova remnant if the donor star has low mass. 1.3.6 Infrared and Optical Counterpart An optical counterpart of Cir X-1 was resolved from a group of three stars using images and pho- tometry in the V[0.55µm], R[0.70µm], and I[0.90µm] optical bands and J[1.25 µm], H[1.65 µm], and K[2.2 µm] infrared bands. This counterpart is highly reddened and shows variability with the 16.55-d period [50]. Each cycle shows an infrared flare beginning shortly after phase zero. IR magnitudes at peak can reach K=7.2, H=8.5, and J=9.7, while the non-flaring component has magnitudes K∼11.5, H∼12.1, J∼13.3 [23, 24]. The magnitude of the non-flaring component in optical bands is I∼18.4, R∼19.6, and V∼21.5 [50]. The non-flaring V−K color is ∼10 mag which would suggest a late type (low-mass) companion. However, the optical spectrum shows no atmospheric features that would identify the spectral type of the companion star, suggesting that disk emission dominates over the light of the companion [17]. An early type (high-mass) star would have to be reddened by AV ∼ 11 mag. Based on the minimum X-ray absorption column observed (NH ∼ 1022), the interstellar extinction can be estimated to be AV ∼ 4 mag, favoring a low-mass star. However, other methods result in AV ∼ 5 − 11 mag, so that an O or B-type dwarf or giant (but not supergiant) cannot be ruled out [24]. Past X-ray behavior has suggested that Cir X-1 is a low mass X-ray binary, and the results in this thesis strengthen this X-ray evidence by demonstrating Z-source behavior (Chapters 5 and 6). 1.4. OBJECTIVES AND OVERVIEW OF THIS PROJECT 31 1.3.7 Radio Ephemeris Based on the onset times of radio flares (defined to be phase zero) observed at Hartebeeshoek Radio Astronomical Observatory (HartRAO) between 1978 and 1988, G. Nicolson calculated the following ephemeris equation, reported by Stewart et al. [68]: JD0 = 2443076.87 + (16.5768 − 0.0000353N )N. (1.1) Although radio flares have been too weak during the past decade to determine an updated radio ephemeris, equation 1.1 has proven to be be an excellent predictor of the current X-ray behavior (see Chapter 3). It should be noted, however, that phase zero no longer necessarily corresponds to the onset of radio flares. Furthermore, if the cycle is indeed orbital, the exact time of periastron relative to phase zero is not known. In this thesis, all references to the phase of Cir X-1 cycles are based on equation 1.1, unless otherwise noted. The quadratic term in equation 1.1 implies that the period is currently decreasing by about two minutes per year ( P =4.26×10−6=134 s/yr). This would imply a characteristic time scale (P/2 P ) of only ∼5000 y for the period to change by an amount equal to the current period, assuming the period changed at a constant rate. If the period truly is changing at such a rapid rate, it would further support the notion that Cir X-1 is very young. However, the radio and X-ray data are not yet able to confirm that a quadratic term is required. The change of period in equation 1.1 amounts to a shift of phase zero of about 25 minutes per year or about 4 hours over a decade. For convenience, the zero point of equation 1.1 can be shifted by 423 cycles to 1995 De- cember 31.04, immediately after the December 30 launch of RXTE , by making the substitution N ′ = N − 423: JD0 = 2450082.54 + (16.54694 − 0.0000353N ′)N ′. (1.2) Thus, the current period will be quoted as 16.55 d throughout this thesis. 1.4 Objectives and Overview of this Project The M.I.T. Center for Space Research was heavily involved in the pre-launch planning and instru- mentation for the Rossi X-ray Timing Explorer satellite (RXTE ), and continues to participate in the operations and scientific endeavors of that observatory now that it is in orbit. The capabilities and scientific instruments of RXTE are described in Chapter 2. Briefly, the main strengths of RXTE are: excellent time resolution, large collecting area, broad spectral coverage, flexible scheduling, and ability to monitor the X-ray sky to track the intensity of known sources and detect transient 32 phenomena. CHAPTER 1. INTRODUCTION A main objective of this project was to apply the unique capabilities of RXTE toward some of the outstanding problems of Circinus X-1 and to relate the results to X-ray binaries in general. Since the overall intensity of Cir X-1 is known to evolve over a period of years, we first sought to establish the current state of the source and to study the nature of its cycle profile. Chapter 3 gives a summary of the long-term behavior of Cir X-1 observed with RXTE as well as an overview of specific studies we carried out through observations of specific cycles. A major outstanding issue for Cir X-1 is its unique status among LMXBs as a possible high- M atoll source that can at times show some types of Z-source behavior. Thus, several studies were carried out to explore the characteristics of the rapid variability in Cir X-1, as well as how those timing properties relate to other properties such as orbital phase (Chapter 4) or spectral/intensity state (Chapters 5 and 6). Chapters 5 and 6 also demonstrate how the energy spectrum evolves during various stages of the flaring state. The variability of Cir X-1 has been described as both flares and dips. Dips are important to understand for several reasons: (1) they might be caused by some part of the mass transfer process, (2) they help probe the local environment of the system, and (3) they can be confused with other variability if not properly identified. Detailed spectral analysis of dips in the intensity of Cir X-1 is presented in Chapter 7. Since the periodic flaring activity is also observed at other wavelengths, we organized two multi- frequency campaigns to study correlated variability. These observations are presented in Chapter 8. An additional objective of this project was to explore how the eccentric orbit of Cir X-1 affects mass transfer and the evolution of the system. In Chapter 9, I show the results of simple computer simulations of mass transfer due to a stellar wind or Roche-lobe overflow, for systems with an eccentric orbit. An eccentric binary evolution code was also developed (Chapter 10) to explore possible evolutionary paths for systems similar to Cir X-1. The conclusions based on these studies are discussed in Chapter 11, as are several promising future studies of Cir X-1. Chapter 2 The Rossi X-ray Timing Explorer 2.1 Overview The Rossi X-ray Timing Explorer (RXTE , see Figure 2-1) is a NASA orbiting astrophysical observa- tory designed to provide temporal and spectral information about celestial X-ray sources [10]. The satellite was launched on a Delta II rocket on 1995 December 30. Its primary targets are compact objects in our galaxy, such as white dwarfs, neutron stars, and stellar-mass black holes, as well as active galactic nuclei, which may contain super-massive black holes. The strengths of RXTE are its high time resolution, large collecting area, broad spectral coverage, and ability to detect and respond quickly to transient phenomena. These capabilities are achieved through three instruments: the Proportional Counter Array (PCA), the High-Energy X-ray Timing Experiment (HEXTE), and the All-Sky Monitor (ASM). In addition, a specialized on-board computer, the Experiment Data System (EDS), serves to pre-analyze and compress data from the ASM and PCA to maximize use of telemetry. Each instrument is briefly described below. For more details see The RXTE Technical Appendix [61]. 2.2 Proportional Counter Array The PCA (built by the Goddard Space Flight Center of NASA) consists of five Proportional Counter Units (PCUs) with a total open area of ∼6250 cm2. This large collecting area allows photons to be detected at a rate high enough to study variability as short as milliseconds. The detector electronics record the arrival time of each photon to an accuracy of 1 µs, and the EDS is also capable of processing data with 1 µs resolution (see section 2.5). The PCUs are sensitive to X-ray photons 33 34 CHAPTER 2. THE ROSSI X-RAY TIMING EXPLORER Figure 2-1: Rossi X-ray Timing Explorer spacecraft. 2.3. HIGH-ENERGY X-RAY TIMING EXPERIMENT 35 with energies between 2 keV and 90 keV, and can detect sources down to a flux level of a few mCrab, where 1.0 Crab ≈ 1060 µJy at 5.2 keV. The intrinsic energy resolution (∆E/E, FWHM) of the PCA is about 18% at 6 keV and 10% at 20 keV. The PCUs have no imaging capability, but collimators restrict the field of view to ∼ 1◦. Since the number of bright X-ray sources in the sky numbers less than 100, the 1◦ field of view is small enough to exclude other bright sources in all but the most crowded fields (i.e., towards the Galactic center). Each PCU detector consists of a chamber of xenon gas (and a small quantity of methane) with three layers of high-voltage anode wires. The anodes closest to the detector walls, and a fourth layer on the bottom form an anti-coincidence detector used to reject events due to particles which enter through the walls of the detector. X-ray photons enter through the front window of the chamber and eject a photo-electron from an atom in the gas. The electron is accelerated toward an anode wire, creating ion pairs along its path. The multiplied electron signal is read out by the anode electronics. A propane chamber between the main chamber and the detector window is only sensitive to low- energy photons (< ∼3 keV). Anode wires in the propane layer are also used in the anti-coincidence logic to screen out electron-induced events. All data from the PCA is sent to the EDS for processing before being telemetered to the ground (see section 2.5). The capabilities of the PCA are particularly well-suited for studies of the rapid variability in X- ray binaries, and we have made use of these capabilities in our study of Cir X-1. Data can be obtained at high time resolution (fractions of a millisecond) simultaneously in multiple energy bands to study the variability components (such as the quasi-periodic oscillations) and their energy-dependence (Chapters 4, 5, and 6). All PCA data are also telemetered as 16-s spectra with 129 channels, allowing more detailed spectral studies to be made on time scales shorter than or comparable to most changes in the state of the source (Chapters 6 and 7). 2.3 High-Energy X-ray Timing Experiment The HEXTE detectors (built by the University of California, San Diego) are sensitive to X-ray photons in the 20–200 keV band, extending the full capability of RXTE to cover two orders of magnitude in photon energy (2–200 keV). They are co-aligned with the PCA detectors and are also collimated to produce a ∼ 1◦ field-of-view. The HEXTE consists of two independent clusters of four NaI(Tl)/CsI(Na) phoswich scintillation detectors. Photons entering such a detector undergo multiple interactions in the NaI scintillation crystal, and the resulting signal is transferred through the CsI crystal to a photo-multiplier tube. The CsI crystal also serves to reject photons that only partially deposit their energy and events that are induced by particles. The two detector clusters are 36 CHAPTER 2. THE ROSSI X-RAY TIMING EXPLORER alternately rocked on and off-source to provide continued background monitoring. Each phoswich detector has a net collecting area of 200 cm2 and an energy resolution of about 13% at 60 keV. HEXTE data is obtained in parallel with all PCA observations. For Cir X-1, the HEXTE count rates are relatively low since the X-ray spectrum of Cir X-1 is soft, resulting far more counts in the PCA than HEXTE. Furthermore, the analysis of HEXTE data is complicated by the rocking detectors and complex calibration issues. Thus, the data we obtained for Cir X-1 from HEXTE is not incorporated into this thesis. 2.4 All-Sky Monitor The ASM (built by the Center for Space Research at M.I.T.) consists of three scanning shadow cameras (SSCs) which regularly monitor all bright (> ∼10 mCrab) celestial X-ray sources (about 10 times per day), providing long-term light curves of their intensity in three energy bands between 2 keV and 12 keV [42]. Each SSC contains a position-sensitive proportional counter mounted below a wide-field collimator (∼ 6◦× ∼ 90◦ FWHM). The PSPCs each contain eight carbon-coated quartz fibers which are used as anode wires in a chamber filled with Xe and CO2. The anode electronics read out signals from both ends of each anode and can determine the position along the wire by comparing the relative strength of the signals due to the resistance of the anodes. A mask covering the outer window of the collimator is coded with rectangular slit openings which cast X-ray shadows on the detector and allow positions to be determined (typical error boxes of 3′ × 12′). The three cameras are mounted on a tripod at one end of the satellite (see Figure 2-1) and are rotated together, stopping for 90-s dwells every 6◦. A sensitivity of about 30 mCrab (3σ) is achieved for each such 90-s exposure. The fields-of-view of two of the cameras point in the same direction perpendicular to the rotation axis and cross at a ∼ 24◦ angle (resulting in positional error boxes that cross and can be used for further refinement of positions). The third camera points along the rotation axis. Several of the anodes lost functionality shortly after launch, but this has not significantly affected the overall performance. For Cir X-1 the ASM has proved invaluable in understanding the general spectral and intensity profile of the 16.55-d cycle. The ASM also plays an important role by placing other observations (e.g., our PCA observations) in context with the rest of the cycle and by aiding plans for future observations. With continued successful performance, the detailed long-term light curve will allow us to track how the source behavior evolves between its various states (see section 1.3). So far, the ASM light curve of Cir X-1 has shown the source to maintain the same high baseline intensity level for the entire duration of the RXTE mission (see Chapter 3), with no evidence for evolution on a 2.5. EXPERIMENT DATA SYSTEM 37 Channel Energy (keV) Abs 0–9 10–13 14–17 18–23 24–35 36–49 50–87 88–249 Std2 0–5 6–9 10–13 14–19 20–31 32–45 46–66 67–127 Epoch 1 0–2.6 2.6–3.6 3.6–4.7 4.7–6.3 6.3–9.5 9.5–13.2 13.2–24 24–71 Epoch 2 0–3.0 3.0–4.2 4.2–5.5 5.5–7.4 7.4–11.1 11.1–15.6 15.6–28 28–84 Epoch 3 0–3.5 3.5–5.0 5.0–6.5 6.5–8.7 8.7–13.1 13.1–18.3 18.3–33 33–99 Band A B C D E F G H Table 2.1: Approximate energy ranges for eight PCA bands (A–H) defined in terms of absolute (0–255) and Standard2 (0–128) channels. Gain epoch 1 includes all observations before before 1996 March 21, epoch 2 1996 March 21 – April 15, and epoch 3 after 1996 April 15. Energy values shown are derived from the full PCA with all detectors and layers added. few-year time scale as has been suggested based on previous observations (see section 1.3). 2.5 Experiment Data System The EDS (also built by the CSR at M.I.T.) is a computer containing eight microprocessor Event Analyzers (EAs), of which six are dedicated to handling PCA data while the other two are dedicated to the ASM. For each 90-s dwell, one of the ASM EAs produces position histograms of the X-ray shadow patterns on each wire of each detector in the three energy channels mentioned above. The second ASM EA produces time-series histograms of "good" and background count rates, as well as pulse height histograms useful in monitoring the gain of the detectors based on an on-board calibration source. Two of the PCA EAs process data using standard modes: one that provides count rates for each layer of the five PCU detectors and other rates every 1/8 s (Standard1) and another which provides 129-channel spectra every 16 s (Standard2). The other four PCA EAs can be set to process data in parallel using a wide variety of modes in order to fit desired information into the telemetry constraints. Most data modes fall into two categories: event modes which send down the time and energy of every event and binned modes which send down the number of photons detected in selectable energy and time bins. Many data modes process data from within standard sub-bands of the entire PCA energy range. These sub-bands are defined in terms of the 256 "absolute" channels of the PCA. Table 2.1 shows the channel and energy ranges for eight standard bands. The gain of the PCA detectors was changed twice early in the mission, resulting in a different channel-to- energy relationship for each of three epochs. Although the relationship is not strictly linear, a single 38 CHAPTER 2. THE ROSSI X-RAY TIMING EXPLORER channel-to-energy conversion factor is a good approximation for each epoch: through 1996 March 21, 0.27 keV/chan; 1996 March 21 – 1996 April 15, 0.32 keV/chan; after 1996 April 15, 0.38 keV/chan. All light curves, hardness ratios, and Fourier power spectra in this thesis have been binned into one or several of these standard bands. The four PCA Event Analyzers not devoted to standard modes are typically configured, through modes selected by the observer based on science goals, to provide optimal time resolution and energy resolution within the available telemetry. For Cir X-1, we generally chose two or three "single-bit" modes (single-channel binned modes) which covered absolute channels 0–35 of Table 2.1 (up to 13 keV in epoch 3) using various combinations of energy bands A–E. These configurations provided a time resolution of 61 µs or 122 µs. Due to lower count rates in the higher channels, the energy and time of every photon with energy above channel 35 (bands F–H in Table 2.1) could be recorded with an event mode having 32 or 64 channels covering absolute energy channels 36–249 and time resolution of 16 µs or 64 µs. Sometimes a binned mode was used to provide 32 low-energy channels at 0.5 s resolution to allow the possibility of creating spectra on time scales shorter than the 16 s provided by the Standard2 mode. Thus, all Cir X-1 data were obtained with high time resolution in multiple energy bands for rapid variability studies (Chapters 4, 5, and 6) and good energy resolution every 16-s or less for spectral studies (Chapters 6 and 7). Chapter 3 Overview of RXTE Observations 3.1 ASM Observations of Circinus X-1 3.1.1 General Features of the ASM Light Curves The RXTE All-Sky Monitor (see section 2.4) has provided 90-s intensity measurements of most bright X-ray sources in three energy channels (1.5–3 keV, 3–5 keV, and 5-12 keV) about 12 times daily since early 1996 [42]. The intensity for each source in the ASM field of view, and the diffuse X-ray background, is obtained from fits to the ASM mask patterns of each 90-s exposure. Thus background has, in effect, been subtracted from the source light curves. The ASM light curve of Cir X-1 for the entire RXTE mission to date (through 1998 May) is shown in Figure 3-1. (A gap of ∼40 days occurred early in the mission due to instrumental problems, and a gap of 2–3 weeks has occurred annually, near days 420 and 780, due to sun-angle constraints.) These observations show Cir X-1 in a sustained bright state with a baseline intensity level very similar to that of the Crab nebula (1.0 Crab ≈ 75 ASM counts/s; 1060 µJy at 5.2 keV). Each cycle shows strong flaring wherein the intensity increases by a factor of three or more. Phase zero of each cycle (based on equation 1.1) is indicated in Figure 3-1. The consistent alignment of brief dips with phase zero and subsequent flaring shortly after phase zero demonstrate that the decade-old radio ephemeris is very successful in predicting the current X-ray behavior. The flaring state typically begins during the day following phase zero, reaching as high as 3.7 Crab (corresponding to a luminosity at 8 kpc that is several times the Eddington limit) and typically lasting 2–5 days. The profile of the flaring state is actually quite variable, sometimes showing a secondary flare or an extended main flare lasting most of the cycle (see Chapter 5). Many cycles observed with the ASM show brief dips below the 1 Crab level, usually immediately 39 40 CHAPTER 3. OVERVIEW OF RXTE OBSERVATIONS Figure 3-1: ASM light curve (1.5–12 keV; 90-s exposures) covering 1996 January – 1998 May. Phase zero is indicated by vertical dotted lines. Bars along the top axis mark the times of PCA observations (labeled A–H; see section 3.2). 3.1. ASM OBSERVATIONS OF CIRCINUS X-1 41 before or shortly after phase zero, but occasionally occurring later in the cycle as well. Previous satellites have also observed intensity dips from Cir X-1 near phase zero. For example, Brandt et al. [12] recently presented evidence for strong absorption during a phase-zero intensity transition in ASCA observations of Cir X-1. More extensive RXTE PCA observations of dips will be discussed in Chapter 7. 3.1.2 Cycling Hardness Ratios The light curve and hardness ratios of counts in adjacent energy channels for the first ten 16.55-d cy- cles of Cir X-1 observed with the ASM are shown in Figure 3-2. These data show significant spectral evolution during the non-flare phases. The hardness ratios were found to generally increase (harden) from about phase 0.2 until phase 0.6-0.85. Beyond this point the hardness ratios usually flatten, decrease, or dip and rise again before phase zero. Near phase zero, the ratios vary dramatically (not apparent in the one-day averages shown) but generally become quite low during the flaring phases (0 < ∼φ< ∼ 0.2). The hardening in the two colors (0.2 < ∼φ< ∼ 0.7) is usually due to both an increase in the 5–12 keV count rate and a decrease in the 1.5-3 keV count rate. 3.1.3 16.55-day Period The onset of flaring relative to phase zero can vary by of order one day. Over the ∼50 cycles observed by the ASM, this allows the period to be determined to within about 1/50 d = 0.02 d. Since this uncertainty is similar to the change in period of 0.03 d over 20 yr implied by equation 1.1, it is not yet possible with the ASM to confirm or reject a period change of that magnitude. However, to obtain the best-fitting period I used a 64-times oversampled Fourier power density spectrum from the ASM data in Figure 3-1 (and fitting the peak corresponding to the ∼16.55-d period with a Gaussian). This method gave a best-fitting constant period of 16.555 d. Within the 0.02 d uncertainty, this is consistent with equation 1.1, with or without the quadratic term. The possibility of using the phase-zero dips as "markers" for the period is explored in the next section (using the average folded light curve). Hopefully the ASM will continue to operate for several more years, potentially allowing the period to be determined with the accuracy necessary to test the constancy of the period. 3.1.4 Folded Cycle Profile ASM light curves of Cir X-1 folded at the 16.55-d period (using equation 1.1; Figure 3-3) show the rapid rise and more gradual decline of the profile. The scatter of points during the "flaring" state demonstrates the high degree of variability at those phases. The 5–12 keV band exhibits flaring 42 CHAPTER 3. OVERVIEW OF RXTE OBSERVATIONS Figure 3-2: Light curve and two hardness ratios for the first ten full 16.55-day cycles of Cir X- 1 observed with the ASM. Each light-curve point is the average count rate (I [1.5–12 keV]) from a single 90-s ASM exposure. The hardness ratios were obtained from 1-day averages of the 90-s intensities and are defined as HR1=I [3–5 keV] / I [1.5–3 keV] and HR2=I [5–12 keV] / I [3–5 keV]. The vertical dotted lines indicate phase zero. 3.1. ASM OBSERVATIONS OF CIRCINUS X-1 43 Figure 3-3: Folded light curve of Cir X-1 from the three ASM energy bands. The data are repeated in order to show two complete cycles. Each point is from a 90-s exposure, with a typical error bar of 1–2 c/s. The folded data spans the entire time period covered by Figure 3-1. 44 CHAPTER 3. OVERVIEW OF RXTE OBSERVATIONS above a baseline that dips below the quiescent level. This is related to the spectral softening seen during flaring in Figure 3-2. Although dips typically occur anywhere between phases −0.03 and +0.10, they occur most consistently in a narrow range immediately before phase zero. There is a clear absence of points at the baseline level in the narrow phase range 0.985–0.989; however, the total number of points in a phase range of that width is typically only about 30, so sparse coverage may play a role in the width of this gap. The data shown in Figure 3-3 for the three ASM bands were also binned into 100 phase bins and averaged within each bin. These average folded profiles, as well as those for the the total ASM energy band and the two hardness ratios, are shown in Figure 3-4. The average profile has a very narrow dip at phase −0.012 (±0.004), followed by a sharp rise over the first 1.5 d of the cycle and a much more gradual decline. The hardness ratios are anti-correlated with the total intensity. Similar intensity and hardness profiles were seen in folded Ginga ASM data from 1987 [75]. The folded profile from the highest energy ASM band (5–12 keV) does not exhibit the large increase in flux characteristic of the other two bands and shows only increased variability during the flaring state, due to the scatter of points above and below the baseline (see Figure 3-3). We attempted to use the narrow dip at phase −0.012 in the average folded curves to constrain the best period. Constant periods of 16.535–16.555 d all showed similar dips, and the dips were deepest (and most similar to Figure 3-4) using a period of 16.540–16.545. Although the range of periods showing the narrow dip is lower than the earlier radio periods by about 0.02–0.04 d, we have no a priori knowledge of what shape (if any) this feature should have, and thus must use caution when interpreting periods determined by maximizing its depth (or other property). 3.2 PCA Observations of Circinus X-1 3.2.1 Overview We have proposed and carried out an extensive observation campaign for Cir X-1 with the RXTE PCA, collecting over 800 kiloseconds of data, as shown in Table 3.1. The project consisted of eight separate "studies" which used a variety of strategies for different 16.55-d cycles: We sampled several cycles with brief observations (typically about 6000 s), distributed at different phases, to study how the detailed timing and spectral characteristics evolve with orbital phase and relate to the spectral evolution indicated by the cycling hardness ratios of the ASM (see above). High-efficiency observations (i.e., ∼60% coverage) during two long time segments (2 and 7 days) including phase-zero were used to study the highly complex dipping and flaring behavior associated with that portion of the cycle. In order to study correlated behavior at different wavelengths, we coordinated two 3.2. PCA OBSERVATIONS OF CIRCINUS X-1 45 Figure 3-4: Left: Folded Cir X-1 light curve from the full ASM energy band (1.5–12 keV) and two hardness ratios. Right: Folded light curves from the three ASM energy bands. The data are binned into 100 phase bins per cycle, averaged in each phase bin, and duplicated to show two complete cycles. Error bars on the light curves indicate the standard deviation of the mean value derived for each phase bin, and these error estimates were propagated when computing the hardness ratios. The folded data spans the entire time period covered by Figure 3-1. 46 CHAPTER 3. OVERVIEW OF RXTE OBSERVATIONS Study Dates A B C 1996 Mar. 5 - 21 1996 Apr. 7 1996 May 7 - 14 Observing Time 76 ksec 10 ksec 54 ksec D E F G 1996 Sept. 20 - Oct. 8 190 ksec 1997 Feb. 18 - Mar 4 1997 May 16 - 21 1997 June 4 - 22 48 ksec 13 ksec 396 ksec H 1997 Sept. 18 - Oct. 4 50 ksec 837 ksec Description 12 observations over one 16.6-d cycle 1 observation during phase-zero dips 7 observations between phase 0.78-0.16, coord. with HartRAO (radio) & SAAO (IR) 60% coverage for 2 days (phase 0.97-0.09), 13 observations over remainder of 16.6-d cycle 8 samples over one 16.6-d cycle 13 1-ksec observations, twice daily 56% coverage for 7 days (phase 0.93-0.33), sampling observations before and after; portions coordinated with AAT (optical), ANU (IR), HartRAO (radio), & SAAO (IR) 8 samples over one 16.6-d cycle Table 3.1: PCA observations of Cir X-1 (1996–1997). multi-frequency campaigns to provide optical, infrared, and radio observations simultaneous with our X-ray coverage. The light curves, from the full PCA energy range, for each of the studies are shown in Figure 3-5 as a function of phase for comparison between cycles. Standard 20-kilosecond light curves of all these observations are collected in Appendix B. The time of each PCA observation is also indicated on the 2.3-year ASM light curve in Figure 3-1. 3.2.2 Observation Summaries The intensity and activity level of Cir X-1 had been observed to change over a period of years (see section 1.3). We thus did not know what to expect from Cir X-1 once RXTE launched. By the end of 1996 February, it was already clear from the very early ASM data that Cir X-1 was currently bright and active. Thus, we requested that our proposed PCA/HEXTE observations of Cir X-1 be activated, and began observations in early 1996 March. The eight studies (A–H) of our project are described briefly here. 3.2.2.1 Study A: 1996 March 5–21 The first study consisted of twelve observations, each ∼6000 s in duration. They were separated by ∼2-day intervals to sample one complete 16-d cycle (see Figure 3-5A). The first observation of study A occurred during the half day before phase zero and showed significant dipping behavior. The second observation occurred shortly after phase zero, and the source was very bright (up to 3.2. PCA OBSERVATIONS OF CIRCINUS X-1 47 Figure 3-5: Light curves for each 16.55-d cycle of Cir X-1 observed with the PCA. For each cycle, the phase-zero date and letter of the study (see Table 3.1) are listed. The intensity is for the full PCA energy range, and a conversion factor of 1 Crab ≈ 2600 counts s−1 PCU−1 has been used. 48 CHAPTER 3. OVERVIEW OF RXTE OBSERVATIONS Figure 3-6: PCA light curves in two energy bands (16-s bins), and the ratio of intensity in the hard band to that in the soft band, for the observation on 1996 April 7 (study B), spanning 15 ks. The lowest intensity in both bands during dips is well above the background level (∼30 c/s and ∼45 c/s for 2–7.4 keV and 7.4–28 keV respectively, not subtracted). ∼2.6 Crab) and highly variable. Nine observations carried out between phases 0.2 and 0.8 showed a quite steady flux, and a final observation shortly before the next phase zero showed increased activity again. These data are presented in Chapter 4, with a focus on the evolution of the timing and spectral properties during the non-flaring phases. 3.2.2.2 Study B: 1996 April 7 This study consisted of a single 10-ks observation carried out on 1996 April 7, approximately one full cycle after the end of the initial monitoring campaign (see Figure 3-6). This observation also showed significant dipping below the ∼1.0 Crab baseline (∼13 kilocounts/s) during the half day before phase zero. The dips are chaotic and accompanied by dramatic spectral evolution (as indicated by the hardness ratio). Similar dipping behavior from study D (see below) was studied in detail and is presented in Chapter 7. This observation occurred during PCA gain epoch 2 (which defines the channel-to-energy conversion factor, see Table 2.1). All following observations of Cir X-1 occurred during gain epoch 3. 3.2. PCA OBSERVATIONS OF CIRCINUS X-1 49 Figure 3-7: PCA light curves in two energy bands (16-s bins), and the ratio of intensity in the hard band to that in the soft band, for the observations on 1996 September 24 – October 8 (study D). Due to some observations with a PCU not operating, all points in this plot were derived from PCUs 0, 1, & 2 and adjusted by 5/3 for comparison with other figures in this section. Phase zeros occurred at days 346.782 and 363.328. 3.2.2.3 Study C: 1996 May 7–14 In 1996 May, the first of two multi-frequency campaigns was carried out, with coordinated infrared, radio, and RXTE observations. These observations spanned several days on either side of phase zero, and showed an increase in intensity in all three frequency regimes during the day following phase zero. These multi-frequency observations are discussed in more detail in Chapter 8. 3.2.2.4 Study D: 1996 Sept. 20 – Oct. 8 In 1996 September, a high efficiency observation was carried out, providing 60% coverage of the 48-hour period beginning 0.5 days before phase zero. These data showed extensive dipping both before and after phase zero and are the focus of Chapter 7, in which the spectral evolution of dips is studied in detail. Following this extensive phase zero observation, 13 sampling observations were made over the remainder of the 16.55-d cycle. These are shown in Figure 3-7 and demonstrate a gradual hardening of the spectrum with orbital phase, as well as more dramatic spectral evolution during dips in the last observation. 50 CHAPTER 3. OVERVIEW OF RXTE OBSERVATIONS 3.2.2.5 Study E: 1997 Feb. 18 – Mar. 4 Eight 6-kilosecond observations were carried out during 1997 February 18 – March 4, sampling a complete cycle at approximately 2-day intervals (see Figure 3-5E). The intensity of this cycle declined from its peak value gradually over ∼10 days. These observations are used in Chapter 5 to study the evolution of timing properties as a function of intensity and spectral state. 3.2.2.6 Study F: 1997 May 16 – May 21 In 1997 May, a series of short (1 ks) public observations of Cir X-1 were carried out between phases 0.35 and 0.7 (the data were originally part of another observer's program, but were made publicly available due to a scheduling conflict). Although these data were not part of our proposed observations, they fit well with our other sampling studies and were thus incorporated into the project. The intensity level during each observation was relatively steady, but the average level varied between 1.0 and 1.6 Crab. 3.2.2.7 Study G: 1997 June 4–22 Our most extensive set of observations were carried out during 1997 June. These included a 7- day period of 56% coverage that included dips before phase zero and most of a very active flaring state. This excellent coverage allowed us to observe continuous transitions between the various spectral/intensity states we found in the earlier 1997 February–March data. These results are presented in Chapter 6. Some of our June PCA observations were coordinated with with radio, optical, and infrared observations made by collaborators in South Africa and Australia. The results from this second multi-frequency campaign are presented in Chapter 8 along with the previous campaign (study C). 3.2.2.8 Study H: 1997 Sept. 18 – Oct. 4 A final set of eight 6-kilosecond observations were carried out during 1997 September–October to again sample a complete 16-d cycle. These observations all showed the intensity to be near the 1.0 Crab baseline, except for the sample at phase 0.06, where the intensity reached 2.5 Crab. The soft and hard light curves and hardness ratio for the flaring observation are shown in Figure 3-8. There is strong variability on time scales of hundreds of seconds during the climb to 2.5 Crab. The hardness ratio is generally anti-correlated with intensity during the first segment and then correlated with intensity during the remainder of the observation. This bi-modal behavior can be understood as relating to different branches in a hardness-intensity diagram. Both types of behavior are associated 3.2. PCA OBSERVATIONS OF CIRCINUS X-1 51 Figure 3-8: PCA light curves in two energy bands (16-s bins; 13 kilocounts/s=1.0 Crab for the combined 2-33 keV band), and the ratio of intensity in the hard band to that in the soft band, for the observation on 1997 September 20 (part of study H), spanning 7000 s. with the spectral/intensity states that are discussed in Chapters 5 and 6 in the context of studies E and G. 52 CHAPTER 3. OVERVIEW OF RXTE OBSERVATIONS Chapter 4 X-ray Timing and Spectral Evolution vs. Orbital Phase This chapter has been published (in a slightly modified form) in the September 20, 1996 issue of The Astrophysical Journal (Vol. 469) [62]. 4.1 Overview We have carried out a study of Cir X-1 (study A, see Figure 3-5) through detailed sampling over a single 16.55-d intensity cycle with the Rossi X-ray Timing Explorer (RXTE ) Proportional Counter Array (PCA). We report here the current state of the source and focus primarily upon the evolution of the emission characteristics away from the flaring activity, i.e., at phases 0.2 < ∼φ< ∼ 0.9 (based on the 1991 radio ephemeris of Stewart et al. [68]). Heretofore, these phases have not been systematically sampled with an instrument of large aperture. In the eccentric binary scenario, the source would be relatively remote from the secondary during these phases. During the non-flaring phases, Cir X-1 remained unusually bright (∼1.0 Crab) and relatively steady. The Fourier power density spectrum of the source was observed to vary with strong correla- tions among low-frequency flat-topped power (< ∼ 1–10 Hz), a QPO peak centered at 1.3–12 Hz, and a broad QPO peak centered at ∼20 Hz up to ∼100 Hz. As orbital phase increased within the cycle, the rms amplitude of the flat-topped power generally (but not monotonically) decreased, while the QPO features generally evolved toward higher frequency. The PCA spectrum was observed to generally harden during the non-flaring phases, consistent with results from the RXTE All-Sky Monitor (see section 3.1). 53 54 CHAPTER 4. X-RAY TIMING AND SPECTRAL EVOLUTION VS. ORBITAL PHASE Julian Date Orbital Obs.1 −2450000.52 Phase3 0.98 0.15 0.28 0.39 0.46 0.50 147.9 150.7 152.8 154.6 155.8 156.6 1 2 3 4 5 6 Julian Date Orbital Phase 0.61 0.63 0.69 0.75 0.78 0.94 Obs. −2450000.5 7 8 9 10 11 12 158.3 158.7 159.7 160.7 161.1 163.8 Table 4.1: March 1996 PCA observations of Cir X-1 4.2 Observations and Results In 1996 March, Cir X-1 was sampled with the PCA (effective area ∼7000 cm2, see section 2.2) twelve times during one 16.55-day orbital period. The observation times and orbital phases are summarized in Table 4.1. The low and high-energy light curves for observations 1–11 are shown in Figure 4-1, along with the hardness ratio derived from those intensities. Observation 12 used a different PCA gain setting (gain epoch 2, see Table 2.1 on page 37) and only 3 PCUs, so it is not included in Figure 4-1. The hardness ratio generally increases during the non-flaring phases as phase increases up to ∼0.7, beyond which the ratios are more steady. This confirms the evolution observed in the ASM data (see section 3.1). Three detailed sample light-curve segments are shown in Figure 4-2. In the full PCA energy band (2–60 keV), Cir X-1 was found to be extremely active, remaining bright (> ∼ 1.0 Crab) throughout the 16.55-d cycle (1 Crab ≈ 13,000 c/s). In the vicinity of phase zero (Obs. 1, 2, & 12), dramatic flares and dips occurred (Figure 4-2a,b). In Obs. 2 (phase 0.15), the count rate climbed to more than 2.5 Crab (Figure 4-2b). Away from phase zero (0.2 < ∼φ< ∼ 0.8; Obs. 3–11), the count rate remained fairly steady (within 15%) at about 1.0 Crab. Figure 2c shows a sample light curve for phase 0.5, which is typical of these observations. The source is relatively steady but exhibits flickering greatly in excess of Poisson statistics. Power density spectra (PDS) were produced from PCA data covering 2–8.6 keV in 61-µs time bins. For each 96-min RXTE orbit, a continuous observation of 1 to 4 ks was divided into 64-s segments, and the PDS was calculated for each segment. All PDS from the segments of a single continuous observation were averaged together, weighted by the total counts. The resulting average 1Each observation lasted 1–5 hr and produced 1.4–10 ks of data. Dates and phases listed are the centroids of the observations. 2JD-2450000.5=147 corresponds to 1996 March 5. 3Orbital phase based on the radio ephemeris of Stewart et al. 1991 [68]. Due to faint radio flares in recent years, this ephemeris is still currently in use as the best available (G. Nicolson 1996, private communication). 4.2. OBSERVATIONS AND RESULTS 55 Figure 4-1: PCA light curves in two energy bands (16-s bins), and the ratio of intensity in the hard band to that in the soft band, for observations 1–11 of 1996 March. Phase zero corresponds to day 148.227. The intensity is well above the background level in both bands (∼25 c/s and ∼40 c/s for 2–6.3 keV and 6.3–24 keV respectively, not subtracted) even during the dips. Observation 12 is omitted since it was carried out with only three PCUs and a different gain setting. 56 CHAPTER 4. X-RAY TIMING AND SPECTRAL EVOLUTION VS. ORBITAL PHASE Figure 4-2: Sample PCA light curves (2–60 keV) comparing 2000-s segments showing (a) strong variability and dips near phase zero (phase 0.98; Obs. 1), (b) a brighter portion of the bright flaring state three days later (phase 0.15; Obs. 2), and (c) a typical observation away from phase zero (phase 0.50; Obs. 6). Background (∼100 c/s) is not subtracted and no deadtime corrections have been made. 4.2. OBSERVATIONS AND RESULTS 57 power spectrum from each RXTE orbit was then logarithmically rebinned, with dead-time-corrected Poisson noise subtracted [88] and normalized to give the fractional rms amplitude squared per Hz (after Belloni & Hasinger [5]). The PDS from seven orbits are shown in Figure 4-3, offset downward by one-decade intervals. These samples show significant low-frequency flat-topped power which cuts off above 1-10 Hz and also a QPO peak which sits on the high-frequency edge of the flat-topped power at 1.3–12 Hz (compared to 1.4 Hz and 5-17 Hz QPO reported near phase zero by [69, 70]. The PDS in Figure 4-3 are ordered by the frequency of the QPO peak. All the PDS during the non-flaring phases exhibit these features. The flat-topped power level systematically decreases as the cutoff and QPO frequencies increase (note that each flat-top level is more than one decade below the previous level in Figure 4-3). Above the cut-off frequencies, the curves all track together with roughly the same power. A few PDS not illustrated here have a steep low-frequency component below 0.1-0.4 Hz. A broad high-frequency QPO is also apparent in Figure 4-3 centered at ∼20 Hz to ∼100 Hz or more. For observations in which the 1.3–12 Hz QPO frequency is below ∼8 Hz, the high-frequency broad peak is always clearly present; otherwise, it is sometimes marginally detectable above 100 Hz. Note that this high-frequency QPO peak moves to higher frequencies with the low-frequency QPO peak. All the PDS in the non-flaring phases were analyzed to further study these trends. The flat- topped component was fit with a zero-centered Lorentzian profile, the two peaked features were also each fit with a Lorentzian, and the high-frequency tail was fit with a power law. Although the results of the fits indicate that this model is crude (via large values of the associated chi-squared statistic), the fits were adequate to determine approximate centroid frequencies and power levels. Errors (rms) for QPO centroid frequencies were estimated from 90% confidence intervals of the fits or 20% of the FWHM of the peak, whichever was greater. Errors for the flat-top power level were estimated to be ∼5%. As the orbital phase increased, the power level of the flat-topped noise generally decreased (Figure 4-4a) while the narrow QPO frequency increased (Figure 4-4b), in accord with the trends noted above. The trends are not monotonic though; a large excursion occurs at φ = 0.4 in both curves. Remarkably, the correlation between these two quantities (flat-topped level and narrow QPO frequency) in Figure 4-4c is very strong (and negative). Moreover, the frequency of the 20–100 Hz peak scales as ∼13 times the frequency of the 1.3–12 Hz peak (Figure 4-4c) with a weak tendency for the frequency ratio to decrease (from ∼16 to ∼11) with increasing frequency. The frequency difference between the two peaks changes by more than a factor of four over this range, from less than 20 Hz to more than 80 Hz. The cutoff frequency of the flat-topped component (as measured by the width of the best fitting zero-centered Lorentzian) also moved to higher frequency with the 58 CHAPTER 4. X-RAY TIMING AND SPECTRAL EVOLUTION VS. ORBITAL PHASE Figure 4-3: Power density spectra from seven RXTE orbits away from phase zero, offset downward at decade intervals. (The ordinate scale applies to the top curve.) The curves are ordered by the frequency of the 1.3–12 Hz QPO peak and are labeled with orbital phase. 4.3. DISCUSSION 59 Figure 4-4: Flat-top power level (a) and QPO frequency (b) as a function of orbital phase over a single orbital cycle. (c) Frequency of the narrow low-frequency QPO and the broad high-frequency QPO vs. the flat-top power level. (d) Cut-off frequency of the flat-topped power vs. flat-top power level. narrow QPO peak (Figure 4-4d). These two frequencies agree within 20% for all observations. 4.3 Discussion The 1.4 and 5-17 Hz QPO reported by Tennant et al. [69, 70] were found during observations near phase zero. It was suggested that the QPO were bimodal, with the 1.4 Hz QPO independent of luminosity and the 5–17 Hz component correlated with the estimated 2–10 keV unabsorbed energy flux [69, 70]. Our observations show a gradual evolution of the QPO from 1.3 to 12 Hz over the non-flaring phases. Some of the phase zero EXOSAT observations which exhibited 5–17 Hz QPO 60 CHAPTER 4. X-RAY TIMING AND SPECTRAL EVOLUTION VS. ORBITAL PHASE also had a QPO peak at 100–200 Hz [69]. This low/high frequency QPO pair is likely to be the evolving QPO pair we have observed, but extended to higher frequencies. The correlation between the flat-topped power level and the cut-off frequency (Figure 4-4d) is similar to the effect observed by Belloni and Hasinger [6] in the the black hole candidate Cyg X-1 and also observed in the neutron star 4U 1608-522 [87] by Yoshida et al. In fact, several properties of 4U 1608-522 show correlations similar to those we have observed in Cir X-1. In two of the three observations of 4U 1608-522 by Yoshida et al. [87], QPO were seen (at 0.4 and 2.0 Hz respectively) just above the knee in the power spectrum, and a broader peaked noise feature was observed at higher frequencies of 1.0 and 5.1 Hz respectively [87]. We note that the ratio of the two peaks is the same in both observations, just as the two peaks in our Cir X-1 observations move to higher frequency with a nearly constant frequency ratio (although the ratio is different for the two sources). In both sources, the spectrum generally (but not always) hardens as these these features move out to higher frequencies and the flat-topped power level decreases. The RXTE observations of Cir X-1 have (1) confirmed a general hardening of the spectrum during the non-flaring phases 0.2–0.7, (2) demonstrated a trend (with moderate scatter) for the power density features to shift to higher frequency and the amplitude of the low-frequency power to decrease with increasing orbital phase during the non-flaring phases, and (3) demonstrated extremely tight correlations among the QPO features and the flat-topped power. In the eccentric-orbit model, these characteristics may reflect the evolving state of the accretion disk while it is being depleted with little or no replenishment. For example, the leveling of the hardness ratios after phase 0.6-0.8 may reflect a time constant (of order 10 days) of the disk structure. Chapter 5 QPOs Associated with Spectral Branches This chapter is scheduled to be published (in a slightly modified form) in the October 10, 1998 issue of The Astrophysical Journal (Vol. 506) [63]. 5.1 Overview We present Rossi X-ray Timing Explorer (RXTE ) All-Sky Monitor observations of the X-ray binary Circinus X-1 which illustrate the variety of intensity profiles associated with the 16.55 d flaring cycle of the source. We also present eight observations of Cir X-1 made with the RXTE Proportional Counter Array over the course of a cycle wherein the average intensity of the flaring state decreased gradually over ∼12 days (study E, see Figure 3-5). This unusually slow transition allows us to demonstrate how the time-variability properties of the source are related to its intensity and its spectral properties (In Chapter 4, we characterized only the quiescent phases of a typical cycle). Fourier power density spectra for these observations show a narrow quasi-periodic oscillation (QPO) peak which shifts in frequency between 6.8 Hz and 32 Hz, as well as a broad QPO peak that remains roughly stationary at ∼4 Hz. We identify these as Z-source horizontal and normal branch oscillations (HBOs/NBOs) respectively. Color-color and hardness-intensity diagrams (CDs/HIDs) show curvilinear tracks for each of the observations. The properties of the QPOs and very low frequency noise allow us to identify segments of these tracks with Z-source horizontal, normal, and flaring branches which shift location in the CDs and HIDs over the course of the 16.55 d cycle. These results contradict a previous prediction, based on the hypothesis that Cir X-1 is a high- M 61 62 CHAPTER 5. QPOS ASSOCIATED WITH SPECTRAL BRANCHES Obs. Julian Date1 Phase Mean Intensity I II III IV V VI VII VIII 2450497.90 2450499.98 2450501.69 2450503.66 2450505.80 2450507.31 2450509.35 2450511.62 0.10 0.23 0.33 0.45 0.58 0.67 0.79 0.93 (Crab)2 2.3 1.8 1.6 1.5 1.2 1.3 1.1 1.0 Table 5.1: PCA observations of Cir X-1 during 1997 February 18 – March 4. atoll source, that HBOs should never occur in this source [54, 77]. 5.2 Observations The RXTE ASM has provided 2–12 keV light curves of Cir X-1 since 1996 February (see Figure 3-1). Throughout the ASM observations, the baseline intensity of Cir X-1 has remained near 1.0 Crab. However, the profile of each 16.55-d flaring cycle can vary considerably. The variety of intensity profiles is illustrated in Figure 5-1, which shows ASM light curves and hardness ratios for three individual cycles. In many cycles, after 3–5 days in the flaring state, the intensity is quite steady for the remainder of the cycle (e.g., Figure 5-1a). In addition to the main flaring episode, some cycles show a mid-phase flare (not always at the same phase) to as high as 2 Crab (Figure 5-1b). Occasionally, the flaring state begins after phase zero and continues for most of the cycle with a gradually decreasing intensity (Figure 5-1c). Despite the variety of intensity profiles, all cycles observed with the ASM show the general pattern of spectral hardening mentioned in section 2.4. During the half day before phase zero, and continuing intermittently for up to two days, brief dips occur in many cycles (perhaps in all cycles, since the ASM coverage is incomplete). These dips are seen as isolated low points in the ASM light curves of Figure 5-1. Eight PCA observations (∼6 ksec each) were carried out at roughly two-day intervals (Table 5.1) during 1997 February 18 – March 4 to sample the 16.55 day cycle shown in Figure 5-1c. The very gradual decline of the flaring-state intensity in this cycle serendipitously provided an opportunity to study intensity-related source properties. All five proportional counter units (PCUs) of the PCA operated normally during each observation, except during the first few minutes of the first 1Midpoint of 2–3 hr observation (∼6 ksec of data per observation). 21.0 Crab ≈ 13,000 counts/s (2–32 keV, all 5 PCUs) 5.2. OBSERVATIONS 63 Figure 5-1: RXTE ASM light curves (1.5–12 keV) for three 16.55 d cycles of Cir X-1 showing different flaring profiles. Each intensity point corresponds to a 90-s exposure by one of the three ASM cameras, and the hardness ratio (HR), defined as the ratio of counting rates for 5–12 keV to 3–5 keV, is shown in one-day averages. The 3–5 keV to 1.5–3 keV hardness ratio exhibits very similar behavior and is not shown here. The intensities are for Cir X-1 after background and other sources in the field of view have been subtracted. The Crab nebula yields ∼75 c/s. Vertical dashed lines indicate phase zero based on the radio ephemeris of Stewart et al. 1991. Day zero corresponds to (a) 1997 April 23.87 (b) 1996 August 2.14 and (c) 1997 February 16.69. For cycle (c), the intensity ranges (I [2.0–18 keV]) seen in the eight RXTE PCA observations (I–VIII in time order) are also shown. 64 CHAPTER 5. QPOS ASSOCIATED WITH SPECTRAL BRANCHES observation when only three PCUs were on. All intensities for that period have been adjusted by a factor of 5/3, but these data are not used in color-color and hardness-intensity diagrams due to gain differences between detectors. Figure 5-2 shows the light curves and hardness ratios (with 16 s time resolution) for each of the PCA observations (I–VIII in time order), made as the intensity declined from 2.5 Crab to 1.0 Crab. On time-scales of hundreds of seconds, the observations made at high intensities show strong variability, while observations at 1.0 Crab (∼13 kilocounts/s) show quite steady count rates. As expected from the ASM hardness ratios, the PCA hardness ratio gradually increases from a low value during the early observations when the source was in the flaring state to a factor four higher as it reached the quiescent level. The relationship between intensity and spectral changes is discussed in detail below. 5.3 Analysis and Results 5.3.1 Color-color and Hardness-intensity Diagrams For the eight PCA observations of 1997 February–March, 16 s intensity and hardness-ratio measure- ments were used to construct color-color and hardness-intensity diagrams (CDs/HIDs, Figure 5-3). The hardness ratios were defined as the ratio of count rates in selected energy bands: a soft color (I [4.8–6.3 keV] / I [2.0–4.8 keV]) and hard color (I [13–18 keV] / I [8.5–13 keV]) for the CD, and a broad color (I [6.3–13 keV] / I [2.0–6.3 keV]) for the HID. The evolution from flaring to quiescent state produced a large range of colors and intensities over the entire cycle. In contrast, each individ- ual observation yielded a localized cluster or track within the CDs and HIDs. The spectral branches for each of the observations are easier to distinguish in the HID than the CD. The long tracks in the HID associated with observations I–V show the color changes associated with the large intensity variations during the flaring phases. The intensity variations are smaller for observations VI–VIII, but significant color changes do occur during these observations as well. The choice of energy bands used in constructing these diagrams can affect the appearance of spectral tracks. For observations showing a single branch, only the length and slope of the branch is affected. Observations V and VI each show two branches. The orientation of these branches is discussed in more detail below. The tracks in the CD and HID are reminiscent of the correlated spectral/intensity behavior of Z and atoll class LMXBs (see Figure 1-1), which also show correlations of temporal properties with position along tracks or branches in CDs and HIDs [27]. Thus, we have investigated how the temporal properties of Cir X-1 are related to position in the CD or HID. For this purpose, we divided the 5.3. ANALYSIS AND RESULTS 65 Figure 5-2: PCA light curves (2–18 keV, 5 PCUs) and hardness ratios (HR = I [6.3–13 keV] / I [2.0– 6.3 keV]) in 16 s time bins for the eight observations made during 1997 February 18 – March 4. A count rate of 13 kcts/s ≈ 1.0 Crab. The data gaps in Obs. VI were longer than as shown here; the second segment of the observation has been shifted left by 4000 s and the third segment by 5000 s. These data were used to construct the hardness-intensity diagram in Figure 5-3. The association with specific regions of that diagram is indicated below each light curve. 66 CHAPTER 5. QPOS ASSOCIATED WITH SPECTRAL BRANCHES Figure 5-3: Color-color diagram (a) and hardness-intensity diagram (b) for all eight observations (I–VIII). In the CD, soft color is defined as I [4.8–6.3 keV] / I [2.0–4.8 keV] and hard color as I [13– 18 keV] / I [8.5–13 keV]. In the HID, the intensity, I [2.0–18 keV], is from all five PCUs and the hardness ratio is a "broad" color: I [6.3–13 keV] / I [2.0–6.3 keV]. Each point corresponds to 16 s of data. Background has been subtracted, but it does not affect the intensity or soft color and only slightly affects the hard color. The three insets in the CD separate overlapping points from observations I–V. The HID track for each observation has been divided into three regions (1–3) for timing analysis. 5.3. ANALYSIS AND RESULTS 67 Figure 5-4: Averaged and rebinned power density spectra (2–32 keV) for each of the three HID regions for each observation. Poisson noise has been subtracted from each PDS (see text). HID track for each of the eight observations into three regions (Figure 5-3b). The choice of numbers for each region was motivated in part by the timing results discussed below, but the numbers serve mainly as reference labels rather than as meaningful quantities (such as Z "rank number"). 5.3.2 Power Density Spectra Fourier power density spectra (PDSs) were computed using 16 s segments with 244 µs (2−12 s) time bins. This was done for both the full 2–32 keV energy range and for four energy channels: 2.0–4.8 keV, 4.8–13 keV, 13–18 keV, and 18–32 keV. The Leahy-normalized power spectra [40] were converted to the fractional rms normalization by dividing by the background-subtracted count rate in the selected band. The expected Poisson level, i.e. the level of white noise due to counting statistics, was estimated taking into account the effects of deadtime [51, 88, 89] and subtracted from each PDS; this method tends to slightly underestimate the actual Poisson level. For each of the 24 HID 68 CHAPTER 5. QPOS ASSOCIATED WITH SPECTRAL BRANCHES regions defined in Figure 5-3, an average PDS was calculated from the power spectra corresponding to points in that region. The PDSs were then logarithmically rebinned. The average PDS (2–32 keV) for each HID region is shown in Figure 5-4. During the extended active state (observations I-VI), a broad peak is often observed near 4 Hz; this feature is prominent in PDSs from observations III–VI, weak in observation II, and indistinguishable from a flat-topped component in observation I (see below). A strong narrow QPO feature is seen at frequencies from 6.8 to 13 Hz in observations VII and VIII. In some cases, especially at higher photon energy (see Figure 5-5), a harmonic peak is observed at twice the frequency of this QPO. A weak narrow QPO feature is present at frequencies above 20 Hz in regions II-1 and VI-1. A sharp "knee" is present at similar frequencies in regions II-2, II-3, III-1, IV-1, VI-2, VI-3, and possibly I-1. Broad high- frequency noise is sometimes seen, e.g. at ∼100 Hz in observation VIII (Figure 5-5b). There is an underlying red continuum spectrum of noise in all of the regions of the HID, but the shape and low frequency slope of the continuum vary over a wide range. The narrow QPO peaks and the low frequency noise in the PDSs from the "quiescent" 1 Crab observations (VII and VIII) resemble previously observed PDSs (Chapter 4, [62, 54, 69]). Those PDSs also contained narrow QPO features, with centroid frequencies in the range 1.3–20 Hz, and similarly shaped low-frequency noise (e.g. see Figure 4-3). The broad high-frequency component that we detect in the present observations is similar to the 20–100 Hz QPO seen in earlier PCA observations (Chapter 4, [62]) and to the 100–200 Hz QPO observed with EXOSAT [69]. The weak narrow QPO feature above 20–30 Hz in regions II-1 and VI-1 occurs near the knee of the low-frequency noise component. This similarity to the LFN and prominent QPO at lower frequency in observations VII and VIII suggests these higher frequency oscillations are produced by the same physical process as the lower frequency QPOs. In observations II and VI, this QPO feature is visible in region 1 as a small peak that fades in region 2 and becomes only a "knee" in region 3 (see Figure 5-4). Thus we assume that this knee is related to the QPO. Similar knees are present in regions III-1, IV-1, and possibly I-1. We include a narrow QPO component in fits of PDSs which show a knee above 20 Hz, but identify these cases as "unpeaked" in the discussion below. Likewise, although no peak appears in the PDSs from observation I, a broad noise component has roughly constant power below about 4 Hz and drops off above that frequency, forming a "knee" which might indicate the presence of the 4 Hz QPO component. The PDS for region I-1 somewhat resembles those of regions III-1 and IV-1, in that all show a break in the power spectrum near 4 Hz and a second knee or change in slope near 30 Hz. We include a broad QPO component in fits of the PDSs for observation I, but we identify these cases as "unpeaked". The PDSs were fit with models comprising both broad-band and QPO components: a power-law 5.3. ANALYSIS AND RESULTS 69 Figure 5-5: Averaged and rebinned power density spectra for HID region VIII-3 in three energy bands. A harmonic peak of the 7.6 Hz QPO is clearly visible in the high-energy channel (c). The broad high-frequency peak, most clear in (b), occurs near ∼100 Hz in this observation. The low- frequency noise cuts off less sharply as energy increases. 70 CHAPTER 5. QPOS ASSOCIATED WITH SPECTRAL BRANCHES for the very low frequency noise (VLFN), an exponentially cut-off power-law for the broad low- frequency noise, a Lorentzian for the broad QPO near 4 Hz, Lorentzians for the narrow QPO and its first harmonic, a broad Lorentzian for the high-frequency peak, and a second power-law to fit the residual Poisson noise at high frequency. The model for each PDS consisted of two to five of these components, depending on which components were necessary for an acceptable fit. The frequency of the harmonic (when present) of the narrow QPO was fixed at twice the fundamental frequency. For the fits of the PDSs from the four narrower energy channels, the QPO centroid frequencies were fixed at the values determined from the 2–32 keV PDSs. There were generally not enough counts to obtain useful PDS fits for the 18–32 keV channel. For use in performing the fits, we estimated the variance of each power in each binned and averaged PDS by calculating the sample variance of the powers in the individual PDSs that were averaged to obtain each point, and dividing the result by the number of the powers used in computing the sample variance. The centroids of the narrow variable-frequency QPO and the ∼4 Hz QPO were measured accu- rately whenever a clear peak was visible. However, in cases where these components are weak or unpeaked, the centroids were less well-constrained. The centroid of the broad high-frequency peak and the cut-off frequency of the LFN were often poorly constrained. Figure 5-6 shows the frequency of the broad and narrow QPOs versus intensity (2–18 keV). The frequency of the narrow peak is generally correlated with intensity, starting at 6.8 Hz at 1 Crab and reaching 32 Hz at 1.3 Crab. At higher intensity, this QPO is sometimes present above 20 Hz and is often unpeaked (i.e., a knee). In observations III–VI, the broad QPO is clearly present at 3.3 to 4.3 Hz. This QPO component was included in the fits of the PDSs from observations I and II, and the resulting frequencies (2.1–4.5 Hz) are shown as unfilled squares (indicating a weak peak or a knee) above 20 kcts/s in Figure 5-6. The ratio of the width of the narrow QPO peak to its centroid frequency (∆ν/ν) is about 0.15 when at 6.8 to 13 Hz. At higher frequency this QPO becomes broader, with ∆ν/ν ∼ 0.4. When the broad QPO near 4 Hz is strong, we find that ∆ν/ν ∼ 1, and when it is weak ∆ν/ν ∼ 2 to 3. Figure 5-7 illustrates the dependence of the rms amplitude of the QPOs upon photon energy. Typical values for the rms amplitude of the 6.8–13.1 Hz QPO at 2–4.8 keV, 4.8–13 keV, and 13– 18 keV are 4%, 5%, and 8% respectively (Figure 5-7a), indicating a weak trend of increasing rms amplitude at higher photon energy. The amplitude of the broad QPO increases significantly at higher photon energy. For clearly peaked 4 Hz QPOs, the rms amplitude is typically about 3%, 8%, and 18% in these three energy bands (Figure 5-7b). The rms values vary considerably when these components are weak or unpeaked but their amplitudes still generally increase with energy (Figure 5-7c). 5.3. ANALYSIS AND RESULTS 71 Figure 5-6: Centroid frequency of the QPOs versus intensity (I [2.0–18 keV]). A filled circle represents the narrow QPO and a filled square represents the broad ∼4 Hz QPO (all points below 5 Hz are the broad QPO). Unfilled circles and squares indicate the approximate frequency of a knee or very weak peak that may be associated with the narrow and broad QPO respectively. Error bars on frequency measurements (filled points only) represent 90% confidence intervals for a single parameter (∆χ2=2.7). In many cases, the error bar for the QPO frequency is smaller than the plot symbol. 72 CHAPTER 5. QPOS ASSOCIATED WITH SPECTRAL BRANCHES Figure 5-7: Rms amplitude of QPOs versus photon energy. Panel (a) shows the rms for the narrow QPO at 7.2 Hz (solid dot) and (b) for an example of the broad 4 Hz QPO (solid box). In panel (c), unfilled circles and boxes indicate the rms amplitude of a component forming a knee or very weak peak that may be associated with the narrow and broad QPOs respectively. The broad QPO points have been offset slightly to the right in energy for clarity. Errors on QPO amplitudes represent 90% confidence intervals. 5.4. DISCUSSION 73 5.3.3 Temporal Behavior versus Position on Spectral Branches The outlines of the HID regions of Figure 5-3b are reproduced in Figure 5-8 with labels summarizing the observed QPO properties. As the hardness ratio decreases and the intensity increases along the HID tracks for observations VIII, VII, and VI, the frequency of the narrow QPO feature increases from 6.8 Hz to 32 Hz. The feature is rather weak and knee-like in observation VI, but it appears to have a width consistent with the width of the prominent QPO peak in observations VII and VIII. A similar weak and somewhat knee-like feature is also present in observation II, where it increases in frequency from 22 Hz to 30 Hz as the intensity increases. The PDSs from observations I, III, and IV all show a knee above 30 Hz at the high-intensity, hard end of their HID tracks; these knees may be related to the narrow QPO features seen in the other observations. The broad 4 Hz QPO is not present in the "quiescent" observations (VII and VIII). This QPO is strongest in portions of the intermediate-intensity observations (III–VI) and is weakly present in the soft, high-intensity observations (II and possibly I). Very low frequency noise dominates the power spectrum of regions V-3 and III-3. Both of these regions appear to begin upturned branches at the low-intensity, soft end of branches showing the more pronounced 4 Hz QPOs. 5.4 Discussion The combined temporal and spectral-branch properties of the observations presented here suggest Z-like behavior. We identify the 6.8–32 Hz QPOs as horizontal-branch oscillations (HBOs), the 4 Hz QPO as normal-branch oscillations (NBOs), and the strong VLFN as flaring-branch behavior (see discussion below). These identifications of characteristic time-variability patterns then help to identify the tracks in the HID as horizontal, normal, and flaring branches (HB/NB/FB), where each 6 ks observation of Cir X-1 appears to have captured a snapshot of portions of one or two of the branches. The spectral branches appear to shift around as the flaring gradually subsides, rather than forming a stable Z pattern. It is likely that the shapes of the spectral branches become distorted somewhat during these large shifts. We now describe the inferred properties of each of the spectral branches in more detail. 5.4.1 Horizontal Branch HID regions VIII, VII, and VI-1 show a narrow QPO peak or knee at 6.8–7.6 Hz, 11.3–13.1 Hz, and 32 Hz respectively. This frequency range overlaps the 13–60 Hz range of typical horizontal branch QPOs [78]. The associated low-frequency noise and harmonic peak are also typical of horizontal 74 CHAPTER 5. QPOS ASSOCIATED WITH SPECTRAL BRANCHES Figure 5-8: Hardness-intensity diagram showing QPO properties for the regions from Figure 5-3b. The frequency of the 6.8 to 32 Hz QPO is labeled (in Hz) beside each region where it is present. Parenthesized frequencies indicate that this component was unpeaked, i.e., a knee. Letters indicate the strength of the broad 4 Hz QPO: S-strong, M-medium, and W-weak or unpeaked. 5.4. DISCUSSION 75 branch power spectra. The broad high frequency peak in Cir X-1 may be related to the high frequency noise component often observed on the horizontal branch. The HID track for observation VI shows the narrow QPO at 32 Hz on a roughly horizontal segment (region VI-1) and a knee at 37 Hz on the right end of this segment (region VI-2). The apex of region VI-2 brings a transition to the 4 Hz QPO, which is dominant on the downward branch of this track (region VI-3). This is very similar to the HB/NB transition in Z sources. When Cir X-1 is in "quiescence" in observations VII and VIII, the "horizontal branch" turns upward and becomes vertical in the HID. For comparison, RXTE PCA observations of Cir X-1 from 1996 March 10–19 which show a narrow QPO peak at 1.3–12 Hz [62] (Chapter 4) are almost entirely confined to the 12.3–14.7 kcts/s (2–21 keV) intensity range. The HID tracks for those observations lie along a nearly vertical line, and probably represent sections of the "horizontal" branch. Observation II may also be on part of the HB, since a weak narrow QPO appears to evolve into a knee and increase in frequency from 22 to 30 Hz as the intensity increases. However, the broad QPO is also weakly visible in PDSs for this observation. The fact that observations II, VI, VII, and VIII all show little variation of the hard color used in Figure 5-3a suggests that observation II may be associated with the other HB observations. 5.4.2 Normal Branch The 4 Hz QPO is observed when the source intensity rises above the "quiescent" 1-Crab level (∼13 kcts/s). It is roughly stationary in frequency (3.3–4.3 Hz when clearly peaked) and broader than the HBO. The feature is easily seen in observations III–VI; at these times the location in the HID moves along diagonal tracks. The ∼4 Hz frequency and motion along diagonal tracks in the HID is consistent with the 4–7 Hz NBOs observed at nearly constant frequency on the NB of typical Z sources [27]. We therefore identify the broad 4 Hz QPO as a normal branch oscillation, and the diagonal tracks for observations III–VI as shifted normal branches. The broad QPO component may be also present in the highest intensity observations, as a weak feature in observation II and in the form of a break near the 4 Hz QPO frequency in observation I. We also note that at the top of the normal branch (regions I-1, III-1, IV-1, VI-2) a knee above 30 Hz is present in addition to the NBO component. A similar broad 4 Hz QPO is present in observations from 1996 March 5–6 (Obs. 1 in Table 4.1) made immediately before phase zero of the cycle showing the 1.3–12 Hz narrow QPO. 76 CHAPTER 5. QPOS ASSOCIATED WITH SPECTRAL BRANCHES 5.4.3 Flaring Branch Beyond the left apex of the normal branch a short upturned branch is observed in HID region V-3 and possibly III-3. The PDS for these regions are dominated by very low frequency noise, which is typical for flaring branches, and no QPO peaks are obviously apparent. We note that in the well- established Z sources neither NBOs nor HBOs are present on the flaring branch, except for Sco X-1 and GX 17+2, in which the NBO evolves into a 6–20 Hz QPO (see [78] and references therein). The left end of the spectral track for observation V bends upward in the HID shown in Figure 5- 3b, but bends downward in the CD in Figure 5-3a. This behavior is demonstrated more clearly in Figure 5-9, which shows CDs and HIDs for observations V and VI. When a broad color (I [6.3– 13 keV] / I [2.0–6.3 keV]) is used as the ordinate of the diagrams (Figure 5-9a,b), the track for observation V turns upward on the left end. When a harder color (I [13–18 keV] / I [8.5–13 keV]) is used as the ordinate (Figure 5-9c,d), this branch turns downward. The CD and particularly the HID version based on the harder color show the most clear similarity to canonical Z diagrams, with the temporal behavior of observations V and VI being generally consistent with horizontal, normal, and flaring branches. The broad-color HID (Figure 5-9b) shows evidence for a shift of the normal branch that does not show up in the other three diagrams of that figure. 5.4.4 Relation to Other Sources Our observations reveal spectral branches which shift in the CD and HID as Cir X-1 evolves from a soft, high-intensity state to a hard, lower-intensity state. The ASM light curves and hardness ratios (Figure 5-1) show that this evolution occurs periodically with the 16.55 day cycle, thus suggesting that the CD/HID shifts may also be periodic. Shifts of the "Z" pattern in CDs and HIDs have been observed in the so-called Cyg-like Z sources: Cyg X-2 [35]), GX 5-1 [34], and GX 340+0 [33]. However, the shifts do not occur periodically in those sources, nor do they have the magnitude of the shifts observed in Cir X-1. The flaring branch of Cir X-1 turns upward when a soft or broad color is used on the vertical axis. When a harder color is used, this branch turns downward but then bends to the left. In the Cyg-like Z sources, the flaring branch sometimes turns upward or starts toward higher intensity and then loops back to lower intensity [35, 34, 33, 55]). In some cases, these sources are observed to "dip" while on the flaring branch [34, 55, 84]), with tracks which turn down and then to the left, similar to that of Cir X-1 in Figure 5-9c. The left end of the horizontal branch in Cir X-1 turns upward and becomes vertical at low intensity (Figure 5-8). On this section of the branch, HBO frequencies are low: 6.8–13 Hz in 5.4. DISCUSSION 77 Figure 5-9: Broad-color and hard-color CDs and HIDs for observations V and VI. In all four diagrams, Obs. V is in the lower left and Obs. VI in the upper right. The CD and HID tracks for Obs. V both turn upward at the left end in the broad-color diagrams (a,b) but turn downward in the hard-color diagrams (c,d). In the HIDs, presence of the 32 Hz HBO, 4 Hz NBO, and VLFN is indicated along the branches. An apparent shift of the normal branch between observations V and VI is labeled in (b). The intensity, I [2.0–18 keV], is from all five PCUs. The soft color is defined as I [4.8–6.3 keV] / I [2.0–4.8 keV], the broad color as I [6.3–13 keV] / I [2.0–6.3 keV], and the hard color as I [13–18 keV] / I [8.5–13 keV]. Each point corresponds to 16 s of background-subtracted data. 78 CHAPTER 5. QPOS ASSOCIATED WITH SPECTRAL BRANCHES observations VII and VIII and 1.3–12 Hz in the earlier 1996 March observations. A similar effect was reported in GX 5-1 [44, 34], in which the HB turns upward at the low-intensity end while HBOs are observed at relatively low frequency (13–17 Hz). Lewin et al. [44] suggested that other Z sources might show such an upward turn of the HB if their intensities and QPO frequencies became sufficiently low. Recent RXTE observations of Cyg X-2 [65] show a long vertical extension of the horizontal branch in an HID. The 5–20 Hz narrow QPO was detected with EXOSAT at an intensity similar to the quiescent level observed by RXTE . We note that absorption dips are responsible for much of the structure seen in the CD shown for that observation; however, the HIDs show that the narrow QPO occurred on an upturned left end of a horizontally oriented track as in our data (see Figures 2–4, 8, & 10 in [54]). At higher intensity during the same observation, the narrow QPO was not present, and we note that some of the high-intensity PDSs show hints of a broad peak near 4 Hz. We thus conclude that the behavior observed by EXOSAT during that observation is related to the Z-like behavior we observe with RXTE . Most of the other EXOSAT observations took place when Cir X-1 was significantly lower in intensity than the "quiescent" level of the current observations. The CDs and HIDs for these EXOSAT observations did not show tracks which could clearly be identified as Z or atoll. Their power spectra were generally dominated by VLFN, typical of atoll sources in the banana state, and sometimes also showed a broad red noise component resembling atoll high-frequency noise [54]. However, these power-spectral shapes are not unique to atoll sources: power spectra for black hole candidates in the high state are dominated by VLFN, as are those of the current observations on the low-intensity end of the normal branch and on the flaring branch (i.e., regions III-3 and V-3). Cir X-1 was expected to never show HBOs since atoll-like behavior was taken as evidence that the magnetic field is not strong enough to allow the magnetospheric beat frequency mechanism (MBFM) to operate [77, 54]. (However, it is also possible that the HBOs are not produced by the MBFM.) The results presented here demonstrate both HBOs and NBOs in Cir X-1 and show no evidence for atoll behavior. Since the atoll-like behavior observed with EXOSAT occurred at lower intensity than in the present observations, it is possible that they do represent a different state of the source. If Cir X-1 actually can show atoll behavior as well as the Z-like behavior shown here, then we would have new clues to the differences between the two types of sources. Such observations would challenge the hypothesis that differences in both M and magnetic field distinguish these two classes. 5.5. SUMMARY 5.5 Summary 79 Our results from an analysis of RXTE observations of Cir X-1 reveal behavior similar to that of Z sources, and, in particular, allow us to identify temporal and spectral signatures of the horizontal, normal, and flaring branches. The spectral variability of Cir X-1 is seen to correspond to tracks in a HID which are similar in direction to the typical direction in the HID of Z sources in general, but the locations of the tracks corresponding to each branch move from observation to observation in a systematic manner. To be specific, in the current observations of Cir X-1 the horizontal branch is characterized by the presence of relatively narrow 6.8–32 Hz QPO features in the PDS. The track in the HID of the horizontal branch is horizontal at the high intensity end and becomes vertical at the low intensity end, where the source is "quiescent", i.e., has an intensity near 1 Crab and is characterized by a relatively low degree of variability on time scales longer than 1 s. The normal branch is characterized by broad 4 Hz QPOs, and by motion in the HID which generally falls along tracks which run diagonally from hard high-intensity locations to soft low-intensity locations. There are also time intervals when the PDS is dominated by very low frequency noise. We identify these intervals as excursions onto the flaring branch. The large amplitude intensity variations associated with the active/flaring state of Cir X-1 can be divided into three categories: (1) motion across the horizontal portion of the horizontal branch and along the normal and flaring branches, (2) shifts of the spectral branches, and (3) absorption dips. While our RXTE observations have allowed us to recognize and distinguish these different types of variability, there is still much to be understood about the physical mechanisms responsible. 80 CHAPTER 5. QPOS ASSOCIATED WITH SPECTRAL BRANCHES Chapter 6 Correlated Timing and Spectral Behavior 6.1 Overview The eight observations presented in Chapter 5 (study E) were characterized by well-separated spec- tral tracks in color-color and hardness-intensity diagrams. Based on timing characteristics, the spec- tral branches were identified with Z-source horizontal, normal, and flaring branches (HB/NB/FB). However, due to the short duration of the observations and the large shifts of the tracks between them, the complete spectral-branch pattern for could only be inferred from fragmented pieces. Those results indicated that the transition between the quiescent and flaring states of Cir X-1 is at least partially related to motion around the "Z" track. In 1997 June, we carried out an extensive RXTE campaign to study in detail the transition to active state, including 7 days of 56% coverage beginning before the phase zero dips (study G, see Figure 3-5). The excellent coverage provided by those observations allowed us to find portions of the transition which clearly demonstrate the full spectral track of Cir X-1. The evolution of the timing characteristics and energy spectrum were studied as a function of position along the track. We found a continuous evolution of the power density spectrum between the different states defined by the spectral branches. We also found that each branch of the spectral track is associated with a specific type of evolution of the energy spectrum (e.g., pivoting about 7 keV in one case and increasing at low energy while remaining constant at high energy in another case). We explored various physical models for the energy spectrum and parameterized the evolution of the spectrum in terms of a two-component model consisting of a multi-temperature "disk blackbody" and a higher-temperature isothermal blackbody. 81 82 CHAPTER 6. CORRELATED TIMING AND SPECTRAL BEHAVIOR 6.2 1997 June PCA Observations The PCA light curves and hardness ratios for the 1997 June observations are shown in Figure 6-1. These data show only moderate variability before phase zero until entering a phase of significant dipping during the half day before phase zero. The hardness ratios show that significant spectral evolution (both hardening and softening) occurs during these dips. Similar dip behavior from another set of PCA observations was studied in detail and is presented in Chapter 7. By phase zero (day 611.5), the main dipping episode ends, and the climb to the flaring state begins. While the intensity increases by more than a factor of three in the lowest energy band (2–4.8 keV), the intensity between 6.3 keV and 13 keV does not climb at all, and above 13 keV the intensity actually decreases by a factor of about 10 over the first 1.5 days following phase zero. This anti-correlation of the low and high-energy intensity during the transition results in decreased hardness ratios after phase zero, as is observed by the ASM (section 2.4). After a relatively smooth transition toward high total intensity during the first day following phase zero, the intensity becomes highly variable (i.e., the "active" or "flaring" state) for the remainder of the observations (about a week). 6.3 Complete Spectral Track The color-color and hardness-intensity diagrams (CDs/HIDs) for all data in Figure 6-1 are shown in Figure 6-2. These data cover a significant portion (10 d) of an entire 16.55 d cycle. The dips seen in Figure 6-1 appear as prominent light tracks with two sharp bends in the CD (initially toward the right of the main arc-shaped locus) and one sharp bend in the HID (I < 2.3 kilocounts/s/PCU). I show in Chapter 7 that tracks with these shapes are indicative of highly variable absorption of a bright spectral component and only moderate, fixed absorption of a fainter component. Having identified absorption dip signatures, I will now focus on non-dip spectral behavior (presumably more directly related to the X-ray source) for the remainder of this chapter. Most of the data fall along a single arc-shaped locus in the CD and a more complicated curved structure in the HID. This behavior is quite similar to that shown for the eight brief sampling observations from 1997 February–March presented in Chapter 5 (see the CD and HID in Figure 5- 3). However, the high-efficiency coverage of the current observations has filled in many of the gaps that occurred between the tracks of the earlier observations. In order to examine the detailed structure with the overall locus of CD/HID points, the data were divided into shorter time segments (of order hours) and plotted separately in CDs and HIDs. In general, each segment produced a fragmented track or tracks, similar to the 1997 February–March 6.3. COMPLETE SPECTRAL TRACK 83 Figure 6-1: Light curves in four energy bands and two hardness ratios for PCA observations of Cir X-1 from 1997 June 10–20, covering a 10-day period around phase zero (φ = 0). The intensities at the beginning of these observations (day 610) are typical "quiescent" levels. Each point represents 16 s of background-subtracted data from PCUs 0, 1, and 2. Ratios of the intensities in the four bands produce soft (I [4.8–6.3 keV] / I [2.0–4.8 keV]) and hard (I [13–18 keV] / I [6.3–13 keV]) hardness ratios. Segments labeled A, B, C & D were used for spectral studies. 84 CHAPTER 6. CORRELATED TIMING AND SPECTRAL BEHAVIOR Figure 6-2: Color-color and hardness-intensity diagrams from PCA observations during 1997 June 10–21 (the entire period covered by Figure 6-1). Each point represents 16 s of background-subtracted data from PCUs 0, 1, and 2. 6.3. COMPLETE SPECTRAL TRACK 85 Figure 6-3: Color-color and hardness-intensity diagrams from time segments "C" (left panels) and "D" (right panels) of Figure 6-1 (1997 June 13.625–14.125 and 1997 June 17.075–17.600 respectively). Each point represents 16 s of background-subtracted data from PCUs 0, 1, and 2. 86 CHAPTER 6. CORRELATED TIMING AND SPECTRAL BEHAVIOR observations. In Figure 6-3, the CDs and HIDs are shown for two time segments (labeled "C" and "D" in Figure 6-1), during which the source traced out a significant portion of several connected branches. These time segments were each about 12 hours in duration and separated by several days. Tracks of other time segments generally each resembled some portion of the entire pattern shown in Figure 6-3, but often with a shifted position in the diagrams. The location of the tracks for segments C and D in the HID for the total 10 d observation set is such that the "hole" produced by the looped branches is visible between 3 and 4 kilocounts/s/PCU in Figure 6-2. The data from the segment C included some absorption dips, resulting in tracks moving off the right side of the CD and the left side of the HID (and far beyond the limits of the plot in both cases). The HID patterns are consistent with the shape of the full "Z" track as inferred by the fragmented tracks in Chapter 5. For segment C, there is a large upturned left extension of the horizontal branch, the horizontal portion of the HB, the diagonal normal branch, and a flaring branch that turns above rather than below the NB. Similar tracks are also present in the HID for segment D, except there is only a small hint an upward turn at the left end of the HB. The plots for the two segments have the same limits and show a significant shift of the HB and upper NB between the two time segments. The HID for segment D is remarkably similar to that derived from RXTE PCA observations of the Z source Cyg X-2 [65]. The Cyg X-2 HID shows a very prominent vertical extension of the HB. The branches in Figure 6-3 are less well separated in the color-color diagrams, as noted in Chap- ter 5. However, the flaring branch clearly turns above the normal branch in the current diagrams, and the upturned left extension of HID horizontal branch of segment C appears to also have an in- creased slope in the CD. The upturned flaring branch is very similar to the flaring branch observed in the color-color diagram for the Z source GX 349+2 in recent RXTE PCA observations [90]. Twenty regions along the HID track of time segment C, shown in the lower left panel of Figure 6- 3, were selected for further timing and spectral analysis. These regions are shown in Figure 6-4 and labeled with increasing numbers from the left HB, through the NB, to the FB. We will refer to this label as the "rank number" after similar work done by other authors on the standard Z sources; however, the numbers at the various apexes have no special significance. It should also be noted that region 6 does not fit in with the monotonic increase of rank number around the spectral track, and may be an indication of a shifted horizontal portion of the HB. The light curves, hardness ratio, and HID regions versus time for the data in Figure 6-4 (time segment C) are shown in Figure 6-5. During this half-day segment, the source generally moves toward higher rank number around the spectral track as the observation progresses. Thus, the data have been divided into four sub-segments which predominantly correspond to each portion of the 6.3. COMPLETE SPECTRAL TRACK 87 Figure 6-4: Hardness-intensity diagram from time segment C (1997 June 13.625–14.125; day 12.625– 13.125). The HID track has been divided into 20 regions from which power density spectra and energy spectra were constructed. 88 CHAPTER 6. CORRELATED TIMING AND SPECTRAL BEHAVIOR Figure 6-5: Top: light curves in 3 energy channels and bottom: total 2–18 keV light curve, broad color, and HID regions for time segment C (1997 June 13.625-14.125; see Figure 6-1). Based on HID region numbers, the predominant spectral branch is identified for each portion of the data. Absorption dips (omitted from HID regions) are clearly identified by decreased intensity coupled with upward broad-color spikes. 6.4. EVOLUTION OF THE POWER DENSITY SPECTRUM 89 HID track: the horizontal and vertical portions of the HB, the NB, and the FB. Brief absorption dips occur in all but the flaring branch during this particular data set; these are easily identified by sharp intensity dips coupled with a spiked increase in broad color. The different branches are characterized by the following characteristics, excluding the behavior associated with the dips: The upturned left portion of the HB evolves relatively smoothly, with a slight increase in 2–6.3 keV intensity and a decrease of almost a factor of two in the 13–18 keV band. The horizontal portion of the HB shows a substantial increase in soft intensity and on average shows relatively steady hard intensity. The normal branch shows increased variability; it is bright in the soft bands but shows a decrease in the hard band. The NB/FB transition occurs at lower intensity in all bands compared to most of the NB. The flaring branch itself is then produced by high-variability "mini-flares" or bursts above the NB/FB apex level. Although the HID regions were defined in Figure 6-4 such that obvious absorption tracks were avoided, one brief dip, on day 612.98, occurred from region 12 on the normal branch and placed a few points artificially across regions 9, 7, and 5. These points are easily identified in Figure 6-5 and are thus not included in subsequent timing and spectral analysis. Likewise, the highest mini-flares on the flaring branch actually extend beyond region 20 and cross regions 8 and 9. In fact a few such points can even be seen above region 10 in Figure 6-4. The FB points that fell into HB regions can also be clearly identified as points with rank numbers of 8 or 9 in the FB portion of Figure 6-5. These are not included in subsequent timing and spectral analysis. 6.4 Evolution of the Power Density Spectrum Fourier power density spectra (PDSs) were computed for each 16 s of time segment C (1997 June 13.625–14.125). Each transform used 216 244-µs (2−12 s) time bins and covered the full 2–32 keV energy range. The expected Poisson level, i.e. the level of white noise due to counting statistics, was estimated taking into account the effects of deadtime [51, 88, 89] and subtracted from each PDS; this method tends to slightly underestimate the actual Poisson level. For each of the 20 HID regions defined in Figure 6-4, an average PDS was calculated from the power spectra corresponding to points in that region. The PDSs were then logarithmically rebinned and are shown in Figure 6-6. The general features of the power spectra are similar to those observed in previous PCA obser- vations (see Figures 4-3 and 5-4), although here a continuous evolution is observed between each timing state in a single observation. The narrow QPO is observed to evolve from 12–25 Hz moving down the upturned left extension of the HB (regions 1–5 and 7; region 6 may be a shifted version of 8). Across the horizontal portion of the HB (8–11), the narrow QPO remains close to 30 Hz and 90 CHAPTER 6. CORRELATED TIMING AND SPECTRAL BEHAVIOR Figure 6-6: Averaged and rebinned power density spectra (2–32 keV) for each of the 20 regions along the HID track in Figure 6-4. The estimated Poisson noise level has been subtracted from each PDS. 6.5. EVOLUTION OF THE ENERGY SPECTRUM 91 fades into a knee, while the broad QPO gradually rises up near 4 Hz. The broad QPO is present all along the normal branch (regions 8–16) but is most prominently peaked in the middle of the branch. On the flaring branch (regions 18–20), no QPOs are present and the power spectrum shows only strong very low frequency noise. 6.5 Evolution of the Energy Spectrum The Standard2 data mode of the PCA produces 129-channel energy spectra every 16 s. A parallel background file was constructed using the "pcabackest" program provided with the FTOOLS analy- sis package (version 4.0) and three background model files provided by the PCA instrument team at NASA/GSFC (pca bkgd q6 e03v01.mdl, pca bkgd xray e03v02.mdl, pca bkgd activ e03v03.mdl). Average energy spectra (and background spectra) were constructed for each of the 20 HID regions, separately for each of the five PCUs. A 1% systematic error estimate was added to each channel of the spectra to account for calibration uncertainties. Representative energy spectra for each segment of the HID track are shown in Figure 6-7. They demonstrate how the spectrum evolves along each branch. One spectrum in each panel is included in the next (since it is derived from the intersection of branches), allowing the comparison to be boot-strapped around the entire HID track. The region numbers in the figures are ordered vertically to aid the comparison. The spectrum is hardest in region 1, at the top of the vertical extension of the horizontal branch. Motion down the branch (softening, regions 1–7) is due to pivoting of the spectrum about ∼7 keV, i.e., increasing intensity below ∼7 keV and decreasing intensity above that energy. Motion to the right across the horizontal portion of the HB (regions 8–11) is due to continued increasing low-energy intensity, but with a nearly constant spectrum above 12 keV. The numerator of the hardness ratio (I [6.3–13 keV] / I [2–6.3 keV]) does not extend very far into this band of nearly constant flux and furthermore is dominated by photons with energies at the low end of the 6.3–13 keV band where the intensity increases by a factor more similar to that of the denominator. So, although the 2.5–25 keV spectrum softens significantly (e.g., comparing the intensity at 5 keV to 18 keV), the hardness ratio in Figure 6-7 decreases only slightly from region 7 to region 11. While moving down the normal branch (regions 11–17), the spectrum decreases across the entire 2.5–25 keV band, but most strongly at high energy (thus further softening). Towards the bottom of the NB, the spectrum begins to show an abrupt steepening in slope above 10 keV. Motion up the FB (regions 18–20) is produced by increasing intensity above ∼4 keV and a relatively constant spectrum below that energy (thus hardening). The change in slope above 10 keV evolves into a 92 CHAPTER 6. CORRELATED TIMING AND SPECTRAL BEHAVIOR Figure 6-7: Cir X-1 energy spectra (2.5–25 keV; PCU 0 only) for several regions on each segment of the HID track in Figure 6-4. The region numbers within the figures are ordered vertically to match the relative intensities at the low and/or high-energy ends of the spectra. 6.6. SELECTION OF SPECTRAL MODELS 93 much more prominent step feature, with structure more complicated than merely a change in slope (see below). 6.6 Selection of Spectral Models In order to explore possible models for use in fitting the spectra from the HID regions, two high- quality spectra were constructed from long steady segments (17–19 ks) at the beginning of the 1997 June observations (time segments A and B in Figure 6-1, from days 609.93–610.16 and 610.66–610.90 respectively). Based on timing properties measured throughout these data sets, time segment A falls on the vertical portion of the HB (strong narrow QPO at 8.4–11.5 Hz) and time segment B falls near the HB/NB apex (weak narrow QPO above 30 Hz and/or the broad 4 Hz QPO). Variability in both of these segments was limited to less than 10% in all energy bands between 2.5–18 keV. We thus constructed a single (averaged) spectrum for each segment, and we will refer to them as spectrum A and spectrum B. Errors in these spectra are dominated by the 1% systematics at all energies up to ∼20 keV. Based on a study of the PCA response matrices using Crab nebula data, R. Remillard has recommended limiting fits to 2.5–25 keV and using only PCUs 0, 1, and 4 (pvt. comm., currently available at http://lheawww.gsfc.nasa.gov/users/keith/ronr.txt). Spectra from each of these detectors are fit separately. Fit parameters reported are the average values for PCUs 0, 1, and when possible 4 (see below), and errors are conservatively estimated from the entire range allowed by 90% confidence intervals from each of the detectors. PCU 4 consistently gives lower normalizations for fitted spectral components, so normalizations and flux values from that detector are not included when computing the average values and their errors. Furthermore, spectrum B was not constructed for PCU 4 since that detector was turned off during part of time segment B. Several single-component models were fit to spectra A and B (all models included an additional component for interstellar absorption). Blackbody and power-law models fit very poorly in both cases, as did a multi-temperature "disk blackbody" spectrum (summed emission from various radii of an accretion disk [49, 46]; model "diskbb" in XSPEC), with reduced χ2 (χ2 r) values of 22–545. A thermal bremsstrahlung model (emission due to acceleration of electrons by protons or ions in an optically thin plasma cloud) provided a better fit to spectrum B (χ2 r = 4.0), but fit spectrum A poorly (χ2 r = 34). A relatively good fit was achieved for both spectra with a modified bremsstrahlung model (see Table 6.1), which includes the self-Comptonization of bremsstrahlung photons to higher energy due to interactions with electrons in an optically thick plasma cloud [38] (model "compLS" in XSPEC). 94 CHAPTER 6. CORRELATED TIMING AND SPECTRAL BEHAVIOR NH /1022 1 (cm−2) 2.80+0.22 −0.23 3.94+0.18 −0.21 Model Components2 Self-Comptonized Bremsstrahlung T (keV) 2.78+0.04 −0.05 2.89+0.12 −0.11 Optical depth 11.05+0.47 −0.36 7.20+0.71 −0.69 Cutoff Power Law 5 norm 4.54+0.24 −0.23 9.67+0.37 −0.43 Blackbody 6 Photon index 0.27+0.94 −0.27 1.95+0.74 −0.86 −0.65+0.72 −0.30 −0.34+0.46 −0.60 Ecut (keV) 1.73+0.58 −0.24 1.67+0.31 −0.27 Blackbody 6 T (keV) R (km) 1.16+0.02 −0.02 1.12+0.01 −0.02 Disk Blackbody 7 24.13+0.38 −0.37 30.46+0.92 −0.40 Tin (keV) 1.81+0.08 −0.06 1.54+0.04 −0.04 Rin cos1/2 θ (km) 9.39+0.64 −0.64 15.63+0.93 −0.84 0.00+0.02 −0.00 0.01+0.21 −0.01 1.44+0.26 −0.27 2.49+0.23 −0.23 T norm (keV) R (km) 2.24+1.53 −0.47 5.97+2.70 −2.29 2.44+0.10 −0.05 2.29+0.09 −0.07 Blackbody 6 4.40+0.34 −0.87 4.61+0.67 −0.73 T (keV) R (km) 2.33+0.02 −0.03 2.20+0.03 −0.03 Blackbody 6 5.38+0.16 −0.15 5.59+0.27 −0.25 T (keV) 2.47+0.04 −0.04 2.28+0.05 −0.05 R (km) 4.10+0.25 −0.25 4.66+0.37 −0.37 Flux/10−8 3 erg cm−2 s−1 4 χ2 r 2.73+0.02 −0.02 3.04+0.02 −0.02 1.16–1.55 1.51–2.02 2.73+0.02 −0.02 3.04+0.02 −0.02 0.79–1.04 0.79–1.35 2.72+0.02 −0.02 3.04+0.02 −0.02 2.51–2.82 1.42–1.78 2.73+0.02 −0.02 3.04+0.02 −0.02 0.79–1.06 0.75–1.31 A B A B A B A B Table 6.1: Fit parameters for spectra A & B for four models. A number of two-component models were also fit to these two spectra. A model using a disk blackbody and power law did not fit well (χ2 r=3–5), mainly because the spectrum does not flatten to a single slope at high energy. A hard (high-temperature, > ∼2 keV) blackbody is often used to fit the spectra of LMXBs thought to contain a neutron star, where emission near the surface might produce a high-temperature blackbody with small effective area. A power law with high-energy exponential cutoff plus a hard blackbody fit well (see Table 6.1) but parameters for the cutoff power law were poorly constrained since the cutoff energy (Ecut ≈ 1.7 keV ) was below the PCA bandpass. Two blackbodies (∼1.1 keV and ∼2.2 keV) fit moderately well (see Table 6.1), but required neg- ligible interstellar absorption. The low absorption is inconsistent with previous measurements from ASCA and ROSAT (both sensitive below 2 keV where the absorption is most easily constrained) which estimated the interstellar column density to be NH =(1.8–2.4)×1022 cm−2 [12, 56]. 1Absorption column density (Hydrogen atoms per cm2). 2 Errors quoted are 90% confidence limits for a single parameter (∆χ2 = 2.7). 3Total 2.5–25 keV flux. 4 Reduced χ2 = χ2/dof , where dof = degrees of freedom = the number of spectral bins (52–54/spectrum) minus the number of fit parameters (4–6). 5High-energy exponential cut-off. 6Blackbody radius assumes a distance of 8 kpc. 7 For the disk blackbody, the inner radius of the accretion disk (times cos1/2 θ, where θ is the disk inclination angle and θ = 0 is parallel to the line of sight) is given for a distance of 8 kpc. 6.6. SELECTION OF SPECTRAL MODELS 95 Figure 6-8: Spectrum A and model (histogram) consisting of a disk blackbody and blackbody (see Table 6.1). The residuals show a peak at 6–7 keV that may be due to an iron emission line. A soft disk blackbody, with temperatures at the inner edge of the disk of 1.5–1.8 keV, plus the hard blackbody fit both spectra quite well (see Table 6.1). This gave absorption column densities roughly consistent with the ASCA and ROSAT values. Other two-component models could also produce similar fits to those describe above. Thus, the spectrum cannot be uniquely deconvolved into separate components. However, for purposes of fitting the HID regions, the model composed of a disk blackbody and blackbody was used. This model is physically motivated (a component from an accretion disk and a harder component from near the surface of a neutron star) and provides good fits with interstellar column densities roughly consistent with previously measured values. The spectral fit for spectrum A for a disk blackbody plus blackbody is shown in Figure 6-8. This plot illustrates the good fit of the model over the entire 2.5–25 keV band. Peaked residuals at 6–7 keV suggest the presence of an emission line, probably iron Kα. Very similar residuals appear in most of the fits discussed above for both spectra A and B. Addition of a Gaussian line to the models does in fact improve the fits in almost all cases; however, the best-fitting line often has an extremely 96 CHAPTER 6. CORRELATED TIMING AND SPECTRAL BEHAVIOR large Gaussian width (σ > 1 keV). The energy resolution of the PCA is about 1 keV FWHM at 6 keV; thus it is difficult to place reliable constraints on the parameters (such as centroid and width) of a narrow component such as an emission line. We have not included an emission line in the fits reported in Table 6.1. The presence of an emission line near 6.4 keV is discussed in more detail in Chapter 7 in conjunction with spectra of absorption dips, which show the line more prominently. 6.7 Fits to Spectra from 20 HID Regions A disk blackbody plus blackbody was fit to the spectra from each of the 20 HID regions. The resulting fit parameters are listed in Table 6.2, and several representative fits are shown in Figures 6-9 and 6- 10. We estimated the distance to Cir X-1 to be 8 kpc (see section 1.3) in converting blackbody and disk blackbody normalizations to radii. The spectra along the horizontal branch (regions 1–11) were all fit relatively well, with residuals similar to those for spectra A and B above, suggesting a possible emission line from iron. These spectra all show column densities of 1.8–2.3×1022 cm−2, consistent with the ASCA and ROSAT values discussed above. The temperatures of both components (∼1.3 keV for the inner disk and ∼2.0 keV for the hard blackbody) are quite stable on the HB, with only slight evidence for cooling down the vertical portion of the branch (i.e., higher temperatures in regions 1–4). The pivoting spectrum on the vertical portion of the HB may be related to these temperature changes. The inner radius of the disk blackbody component (times a factor of order unity involving the inclination angle of the disk) increases from 19 to 33 km, while the radius of hard blackbody remains between 3 and 4 km. Thus, it is mainly changes in the (soft) disk blackbody radius that produce the HB track. These size scales are consistent with the hypothesis that these components arise from emission close to a neutron star, which has a radius of order 10 km. From region 1 to 11, the total 2.5–25 keV flux increases monotonically (with the exception of region 6) from 2.89–4.35×10−8 erg cm−2 s−1 (corresponding to 1.2–1.8 times the Eddington luminosity limit for a 1.4 M⊙ neutron star at 8 kpc). Moving down the normal branch, the quality of the fits decrease, as indicated by the increasing χ2 values in Table 6.2. The absorption column density gradually decreases by a factor of two, but may be related to the decreasing fit quality. The inner radius and temperature of the disk blackbody change only slightly on the normal branch. In contrast, the hard blackbody begins to fade on the upper portion of the normal branch (regions 12–14), as indicated by a decreasing radius for the emission area. In fact, by the middle of the normal branch, the hard blackbody has faded entirely and fits have lower χ2 values without it. Thus, the hard blackbody is omitted from the fits for regions 15–20. The 2.5–6 keV residuals continue to appear similar to those on the HB, but the 6.7. FITS TO SPECTRA FROM 20 HID REGIONS 97 HID 1 NH /1022 2 region (cm−2) 1.83+0.28 −0.32 1.97+0.30 −0.28 2.04+0.27 −0.35 2.18+0.37 −0.32 2.29+0.33 −0.27 2.35+0.30 −0.24 2.31+0.24 −0.25 2.27+0.29 −0.25 2.18+0.21 −0.24 2.01+0.23 −0.24 1.98+0.19 −0.22 1.78+0.22 −0.26 1.63+0.21 −0.25 1.53+0.18 −0.17 1.48+0.15 −0.16 1.43+0.19 −0.17 0.94+0.17 −0.15 0.56+0.16 −0.18 0.64+0.17 −0.16 0.81+0.15 −0.19 T (keV) 2.16+0.09 −0.07 2.15+0.09 −0.10 2.08+0.09 −0.08 2.07+0.18 −0.17 2.02+0.11 −0.14 2.02+0.11 −0.13 2.03+0.09 −0.11 2.01+0.12 −0.14 2.01+0.14 −0.14 2.01+0.12 −0.11 1.93+0.08 −0.08 1.94+0.17 −0.12 2.12+0.22 −0.19 2.34+0.53 −0.40 Tin (keV) 1.45+0.05 −0.05 1.41+0.04 −0.05 1.38+0.04 −0.04 1.36+0.05 −0.06 1.31+0.03 −0.05 1.34+0.03 −0.04 1.30+0.02 −0.03 1.29+0.02 −0.03 1.30+0.03 −0.03 1.31+0.02 −0.03 1.30+0.02 −0.02 1.29+0.02 −0.02 1.28+0.02 −0.02 1.26+0.01 −0.02 1.22+0.00 −0.00 1.19+0.00 −0.00 1.20+0.00 −0.00 1.26+0.01 −0.01 1.29+0.01 −0.01 1.33+0.01 −0.01 Rin cos1/2 θ 3 (km) 19.29+1.31 −1.22 20.87+1.46 −1.29 22.24+1.30 −1.16 23.46+2.27 −1.85 27.22+1.97 −1.74 26.21+1.33 −1.19 28.67+1.20 −1.05 30.06+1.70 −1.48 30.68+1.15 −1.07 31.35+1.31 −1.10 32.90+1.24 −1.17 33.41+1.51 −1.34 34.02+1.09 −0.92 33.94+1.28 −1.06 35.30+0.51 −0.49 35.88+0.50 −0.48 32.69+0.23 −0.53 28.22+0.49 −0.48 27.02+0.35 −0.36 26.25+0.38 −0.37 R 4 (km) Flux/10−8 5 erg cm−2 s−1 3.65+0.50 −0.52 3.71+0.68 −0.64 4.00+0.56 −0.52 3.89+1.27 −1.18 3.78+0.97 −0.94 3.31+0.58 −0.53 3.30+0.65 −0.57 3.34+0.91 −0.83 3.36+0.85 −0.69 3.02+0.86 −0.67 3.52+0.67 −0.54 3.09+0.82 −0.72 1.51+0.70 −0.47 0.84+0.77 −0.46 2.89+0.03 −0.02 2.99+0.02 −0.02 2.96+0.02 −0.02 3.02+0.03 −0.02 3.21+0.02 −0.02 3.37+0.02 −0.02 3.27+0.02 −0.02 3.46+0.03 −0.02 3.75+0.02 −0.02 4.15+0.02 −0.03 4.39+0.03 −0.03 4.35+0.03 −0.03 4.21+0.03 −0.03 3.98+0.03 −0.03 3.68+0.02 −0.02 3.32+0.02 −0.02 3.00+0.02 −0.01 2.94+0.02 −0.02 3.08+0.02 −0.02 3.24+0.02 −0.02 6 χ2 r 1.41–2.14 1.40–1.70 1.32–1.81 2.21–3.47 2.51–3.24 1.41–1.55 2.22–2.76 2.41–2.70 2.29–2.44 1.82–2.56 2.55–3.05 2.87–3.06 3.74–4.66 4.19–5.11 5.15–6.34 5.91–7.52 5.87–7.16 9.59–11.39 9.36–10.45 14.67–17.01 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 Table 6.2: Fit parameters for HID regions 1–20 using a model consisting of a disk blackbody and a blackbody. peaked ∼6.5 keV residual on the left HB becomes broader and more complicated for higher rank numbers. On the flaring branch (regions 18–20), the fit quality decreases further, accompanied by very low values for the absorption column density. A significant contribution to the high χ2 values on the flaring branch is due to the feature above 10 keV that was mentioned above in section 6.5. This feature actually begins to develop on the lower NB, as early as region 14 and is illustrated by two spectra shown Figure 6-11. The residuals shown for each bin have been divided by the size of the error bar for that bin. The feature resembles a narrow line or absorption edge but its location above 10 keV rules out even hydrogen-like iron as a possible source. A number of other spectral models were fit to the HID-region spectra, and all failed to satis- 1 Errors quoted are 90% confidence limits for a single parameter (∆χ2 = 2.7). 2Absorption column density (Hydrogen atoms per cm2). 3 Inner radius of the accretion disk (times cos1/2 θ, where θ is the disk inclination angle and θ = 0 is parallel to the line of sight) assuming a distance of 8 kpc. 4Blackbody radius assuming a distance of 8 kpc. 5Total 2.5–25 keV flux. 6 Reduced χ2 = χ2/dof , where dof = degrees of freedom = the number of spectral bins (52–54/spectrum) minus the number of fit parameters (4–6). 98 CHAPTER 6. CORRELATED TIMING AND SPECTRAL BEHAVIOR Figure 6-9: Fitted energy spectra (2.5–25 keV; PCU 0 only) and model components (disk blackbody and blackbody) for HID regions 1, 7, and 11 of Figure 6-4. The disk blackbody dominates at low energy and the blackbody dominates at high energy. Fit parameters are listed in Table 6.2. 6.7. FITS TO SPECTRA FROM 20 HID REGIONS 99 Figure 6-10: Fitted energy spectra (2.5–25 keV; PCU 0 only) and model components (disk black- body and blackbody) for HID regions 13, 15, and 20 of Figure 6-4. The blackbody only contributes to the high-energy end of the spectrum in region 13, and has faded away entirely in regions 15 and 20. Fit parameters are listed in Table 6.2. 100 CHAPTER 6. CORRELATED TIMING AND SPECTRAL BEHAVIOR Figure 6-11: Spectra from HID regions 17 and 20 (crosses) and fitted models (histograms) consisting of a disk blackbody and blackbody. The ratio of residuals to error bars shows a narrow edge or line- like feature above 10 keV. 6.8. SUMMARY 101 factorally fit the lower portion of the track (rank number 14 and greater). In addition no single or double-component model was found that could reproduce the unusual feature above 10 keV. For example, a broken power law (model with two different slopes above and below a cutoff energy) fails to account for the more step-like nature of the feature. 6.8 Summary Although color-color and hardness-intensity diagrams for Cir X-1 are complicated by absorption dips and shifting spectral tracks, the source clearly traces out a complete Z-source track during short time segments (hours). Power spectra taken from regions of the HID track show continuous evolution from the narrow QPO (increasing in frequency from 12–30 Hz) on the horizontal branch, to the broad 4 Hz QPO on the normal branch, to only very low frequency noise on the flaring branch. Flaring-branch light curves show high levels of variability due to "mini-flares". The energy spectrum on the horizontal branch is fit well with a two-component model consisting of a soft disk blackbody and a harder blackbody (presumably from closer to the surface). In this model, motion along the HB is mainly associated with an increasing inner radius of the disk (in- creasing disk blackbody normalization). This would imply that, as the luminosity increases across the HB, the inner edge of the disk is pushed further away from the surface. It is not clear how this is related to the increasing QPO frequency, which would typically be expected to require a decreasing radius if the QPOs were related to Keplerian motion at the inner edge of the disk. Energy spectra on the normal branch indicate that the hard blackbody fades away, leaving only the disk blackbody. On the lower NB, a feature in the spectrum develops above 10 keV. This feature becomes more prominent on the flaring branch. Although, there is still much to be understood about the physical mechanisms responsible for the spectral and timing evolution of Cir X-1 around the HID diagram, these results have demonstrated the specific timing and spectral properties associated with each branch. 102 CHAPTER 6. CORRELATED TIMING AND SPECTRAL BEHAVIOR Chapter 7 Absorption Dips 7.1 Overview RXTE All-Sky Monitor light curves of Cir X-1 show that dips occur near phase zero of the 16.55- day cycle of the source (Chapter 3). RXTE Proportional Counter Array observations carried out between 1996 September 20–22 provided 60% observing efficiency for 48 hours around phase zero (study D, see Figure 3-5). These observations showed significant dipping activity during much of those two days. We have studied the dramatic spectral evolution associated with the dips, and show that the spectrum throughout the dips is well-fit by variable and at times heavy absorption of a bright component, plus a faint component that is only weakly absorbed. We also show that an iron emission line near 6.5 keV present during dips maintains a constant absolute flux level outside the dips as well, indicating that the line is associated with the faint component. We suggest that these results are consistent with a model in which the bright component is reprocessed by a scattering medium, only slightly modifying the spectral shape, but adding an iron line due to fluorescence. These results are also generally consistent with a spectral study based on ASCA observations of a low-to-high transition in Cir X-1 [12], where the spectrum was fit with a partial-covering model. 7.2 Dips in RXTE Light Curves Light curves from the ASM and PCA show that the intensity of Cir X-1 drops below the 1 Crab baseline near phase zero of most (possibly all) cycles. These low-intensity episodes, which we refer to as "dips", are brief compared to the 16.55-d cycle, lasting as short as seconds (e.g., Figure 6-5) and as long as half a day (e.g., Figure 7-1). They often show abrupt transitions (rise or fall times 103 104 CHAPTER 7. ABSORPTION DIPS much smaller than the total duration), and are characterized by specific spectral evolution which we describe in detail below. During the strongest dips observed in PCA observations, the intensity dropped to levels more than a factor of 10 below the baseline level (2–90 keV). Dips in the ASM light curves (2–12 keV) reached similar levels (i.e., as low as ∼80 mCrab). The dips most often occurred during the half day before and after phase zero, but also sometimes continued through day 1.5 in ASM data and day 2.5 in PCA data. On a couple of occasions, dips also occurred at later phases in the ASM light curve, e.g., near day 4 of the cycle. Dips occurred in all six cycles of Cir X-1 for which we obtained PCA observations during the half day before phase zero (see Figure 3-5). All-Sky Monitor light curves of Cir X-1 show intensity points below the 1 Crab baseline in about half of the cycles observed (see section 2.4). Because ASM coverage is incomplete (about ten to twelve 90-s exposures per day), it is possible that dips might actually occur in all cycles. The PCA observations (study D) made during 1996 September 20–22 (Figure 7-1) covered a 48-hour period around phase zero (with ∼60% observing efficiency) and showed significant dipping for about half of the observed time. These observations provided the opportunity to study dipping behavior in detail, and are thus the focus of this chapter. Dips in other cycles have also been examined and show similar behavior to that presented below. The observations shown in Figure 7-1 began about a half day before phase zero. The intensity of Cir X-1 was very low (an extended dip) during the first half day of the observations, and then returned to the much higher "quiescent" level shortly after phase zero. During the half day immediately following phase zero, the intensity underwent shorter intermittent dips. The strongest dips all reached similar minimum levels of about 200–350 counts/s/PCU (2–18 keV) after background (∼10 c/s/PCU) was subtracted. A second episode of significant extended dipping occurred from about 0.5 to 0.9 days after phase zero, followed by flaring-state activity for the remaining half day of the observation. The flaring state is characterized by (1) a gradual increase of the average intensity in the low-energy PCA bands (which then dominate the total PCA count rate) accompanied by a significant drop in the high-energy bands and (2) individual flares on shorter timescales (less than a few hours) which show increased intensity in all PCA energy bands. The spectral and timing evolution during the flaring state of Cir X-1 was discussed in Chapters 5 and 6. 7.3 Evolution of Hardness Ratios During Dips Three dips in Figure 7-1 have been identified for further study (Dips 1–3). A 4000-s segment of Figure 7-1 is shown for each of these dips in Figures 7-2, 7-3, and 7-4. The light curves in multiple 7.3. EVOLUTION OF HARDNESS RATIOS DURING DIPS 105 Figure 7-1: Light curves in four energy channels and two hardness ratios for PCA observations of Cir X-1 from 1996 September 20–22, covering a two-day period around phase zero (φ = 0). Three dips have been identified for further study. Each point represents 16 s of background-subtracted data from PCUs 0, 1, and 2. Ratios of the intensities produce soft (I [4.8–6.3 keV] / I [2.0–4.8 keV]) and hard (I [13–18 keV] / I [6.3–13 keV]) hardness ratios. The intensity levels of the segment at day ∼347.2 (after Dip 1) are close to the level in each band during the quiescent phases of the orbit. 106 CHAPTER 7. ABSORPTION DIPS Figure 7-2: Light curves in four energy channels and two hardness ratios spanning 4000 s including Dip 1. Time zero corresponds to day 347.041 in Figure 7-1. The intensity and hardness ratio points are the same as in Figure 7-1. Energy spectra were extracted from two 96-s time segments (A and B), indicated by dotted vertical lines, and four 16-s segments (a–d) indicated by diamonds. 7.3. EVOLUTION OF HARDNESS RATIOS DURING DIPS 107 Figure 7-3: Light curves in four energy channels and two hardness ratios spanning 4000 s including Dip 2. Time zero corresponds to day 347.254 in Figure 7-1. The intensity and hardness ratio points are the same as in Figure 7-1. Energy spectra were extracted from two 304-s time segments (A and B), indicated by dotted vertical lines, and four 16-s segments (a–d), indicated by diamonds. 108 CHAPTER 7. ABSORPTION DIPS Figure 7-4: Light curves in four energy channels and two hardness ratios spanning 4000 s including Dip 3. Time zero corresponds to day 347.325 in Figure 7-1. The intensity and hardness ratio points are the same as in Figure 7-1. Energy spectra were extracted from six 16-s time segments (a–f), indicated by diamonds. 7.4. EVOLUTION OF ENERGY SPECTRA DURING DIPS 109 PCA energy channels show that the transitions into and out of dips occur more rapidly at low energy that in higher energy bands (best demonstrated by the gradual ingress of Dip 3 in Figure 7-4), thus hardness ratios initially increase (harden) as the denominator falls more quickly than the numerator. Low energy count rates are also the first to reach a fixed bottom level. This causes hardness ratios to decrease (soften) since the denominator approaches a constant while the numerator continues to decrease. The intensity and hardness ratios from Dip 3 (the last 1716 s of data in Figure 7-4) have been used to produce color-color and hardness-intensity diagrams (CDs/HIDs), shown in Figure 7-5. The spectral evolution during the dip produces tracks with dramatic bends in both the CD and the HID. A similar CD and HID has also been produced for the first three segments of Figure 7-1 (day 346.31–346.52), during which the intensity was mostly in an extended low/dip state (Figure 7-6). In the HIDs of Figures 7-5 and 7-6, dips produce motion to the left (toward lower intensity) and initially upward (harder), but turn downward (softer) when the count rate in the denominator reaches its bottom level. In the CDs (most prominently in Figure 7-6), since two hardness ratios are employed (hard and soft color), motion is initially to the right and upward (harder in both colors but mostly in the lower channels). When the lowest channel reaches the bottom level, the track turns to the left but continues upward. When intermediate channels stop decreasing and only the highest channels still drop, the track finally turns downward. The intensity of the second lowest energy band rapidly drops by a factor of two (at time 3500 s in Figure 7-4) just before settling to its lowest level; this resulted in a gap in the the middle CD branch in Figure 7-5. Evolution of the position in CDs and HIDs apart from dips is interpreted in Chapters 5 and 6 as motion around a "Z" track as well as shifts of the entire "Z" pattern. The exact position of a dip track in the diagrams depends on its starting point, i.e., the baseline intensity and spectrum. Although starting from different positions, all dips we have observed in RXTE light curves of Cir X-1 produce the same general shape in CDs and HIDs (but not always complete tracks if dips are weak). 7.4 Evolution of Energy Spectra During Dips 7.4.1 Spectral Model Cold (neutral) matter most effectively absorbs photons at lower energies, so obscuration by cold matter produces dips that are more gradual at high energy than at low energy, as we observe in Cir X-1. However the fact that strong dips in Cir X-1 appear to "bottom-out" well above background suggests the presence of a faint component unaffected by the dips. Thus, in modeling spectra of dips in Cir X-1, at least two components are necessary: (1) a bright component modified by variable and 110 CHAPTER 7. ABSORPTION DIPS Figure 7-5: Color-color and hardness-intensity diagrams for the last 1716 s of data in Figure 7-4, during which the intensity gradually transitioned from the non-dip baseline to the bottom of a dip. Intensity is I [2.0–18 keV], and the hardness ratios are defined as soft color: I [4.8–6.3 keV] / I [2.0– 4.8 keV], hard color: I [13–18 keV] / I [6.3–13 keV], and broad color: I [6.3–18 keV] / I [2.0–6.3 keV]. Each point represents 16 s of background-subtracted data from PCUs 0, 1, and 2. 7.4. EVOLUTION OF ENERGY SPECTRA DURING DIPS 111 Figure 7-6: Color-color and hardness-intensity diagrams for the first three segments in Figure 7-1 (day 346.31–346.52), during which Cir X-1 was in an extended low/dip state. Intensity is I [2.0– 18 keV], and the hardness ratios are defined as soft color: I [4.8–6.3 keV] / I [2.0–4.8 keV], hard color: I [13–18 keV] / I [6.3–13 keV], and broad color: I [6.3–18 keV] / I [2.0–6.3 keV]. Each point represents 16 s of background-subtracted data from PCUs 0, 1, and 2. 112 CHAPTER 7. ABSORPTION DIPS at times very heavy absorption and (2) a faint component seen through a constant and relatively low column density. For simplicity, we assume that the spectral shape of the two components are similar, as might be the case if the faint component is due to scattering (e.g., by a corona, stellar wind, etc.) of the main bright component back into the line of sight from other initial emission directions. (Although the terms "direct" and "scattered" components would be more physical, we will continue to use the terms "bright" and "faint" since they do not tie the discussion to a particular assumed model.) From a study of the energy spectrum during non-dip observations (Chapter 6), we found that the 2.5–25 keV spectrum of Cir X-1 was generally fit well with a model consisting of a multi- temperature "disk blackbody" plus a hard blackbody. Thus, we adopt this model for the bright component during dips and constrain the faint component to use the same model parameters, but with the flux multiplied by a constant factor f which is less than unity. The absorption column density for the faint component is assumed to remain constant throughout a dip, but the column density for the bright component is allowed to vary to produce the dips. The light curves during strong dips show a significant reduction in intensity even at high energy (13–18 keV), indicating very high column densities (NH > 1024 cm−2). Photo-electric absorption is the dominant process responsible for dips at low energy, but the photo-electric cross-section decreases with energy. Thus, above about 10 keV, Thomson scattering (with an approximately energy-independent cross section) becomes important. Both photo-electric and Thomson cross- sections are included in the absorption calculation for the bright component. The complete model used can then be expressed as: F =hexp(−σphNH (1)) exp(−σT hNH (1)) + exp(−σphNH (2))fi M, (7.1) where F is the observed flux, σph and σT h are the photo-electric and Thomson cross-sections, (2) are the effective hydrogen column densities of the bright and faint components (1) and NH NH respectively, f is the ratio of the flux of the faint component to that of bright component, and M is the disk blackbody plus blackbody model. 7.4.2 Spectral Fitting Strategy Because dip spectra are composed of multiple components (variably absorbed and faint manifes- tations of the disk blackbody and blackbody) we make use of joint fits of spectra from inside and outside dips to simultaneously constrain the parameters of the various components. The combined χ2 value obtained when the model is fit to multiple spectra is minimized by varying the model pa- 7.4. EVOLUTION OF ENERGY SPECTRA DURING DIPS 113 rameters, but allowing only the absorption column density of the bright component to vary between spectra. Since the disk blackbody and blackbody parameters typically evolve on timescales of hours (see Chapter 6), it is important to compare spectra during a dip to spectra immediately outside the dip. In order to perform joint spectral fits, we extracted PCA energy spectra from time segments before and during the three dips shown in Figures 7-2, 7-3, and 7-4. The same procedure for constructing and fitting PCA spectra presented in sections 6.5 and 6.6 is used here as well (comparing results from PCUs 0, 1, and 4). 7.4.3 Energy Spectra Inside and Outside Dips We compiled spectra from a 96-s time segments immediately prior to Dip 1 and from another 96-s segment during the lowest portion of Dip 1 (segments labeled "A" and "B" in Figure 7-2). Likewise, we selected a 304-s segment from before Dip 2, and other segment of the same duration from the lowest portion of that dip (segments "A" and "B" in Figure 7-3). These spectra are shown in Figures 7-7 and 7-8 (top panels), along with fitted curves using the model described above (equation 7.1). The best-fitting parameters for the bright and faint component of each fit are listed in the first two (2)) of the faint component is given in the table. The (1)) of the bright component differs for spectra A and B of each fit and is thus lines of Table 7.1. The column density (NH column density (NH given separately in Table 7.2. The disk blackbody and blackbody parameters for both dips are similar to those obtained for non-dip spectra in Chapter 6 (temperatures of 1.5 keV and 2.2 keV respectively). The flux of the faint component is 10–11% that of the bright component (f ). The column density of the faint component is very low, and consistent with zero. This low value is inconsistent with estimates of the interstellar column density measured with ASCA and ROSAT (NH =1.8–2.4×1022 cm−2) [12, 56]. This may indicate that the spectral shape of the faint component is not exactly the same as that of the bright component. If scattering is responsible for the faint component, then this process might distort the spectrum slightly toward lower energy, thereby giving the appearance of reduced interstellar absorption. The use of different spectral models for the bright and faint components becomes unproductive due to the large number of parameters involved, so we allow the column density of the faint component to account for small difference in spectral shape. 114 CHAPTER 7. ABSORPTION DIPS Figure 7-7: Top: Spectral fits for 96-s segments prior to Dip 1 (spectrum A) and during Dip 1 (spec- trum B). Bottom: same spectra and model (see text), but with a Gaussian emission-line component included to fit the peaked residuals near 6.5 keV. The data shown are from PCU 0 only. 7.4. EVOLUTION OF ENERGY SPECTRA DURING DIPS 115 Figure 7-8: Top: Spectral fits for 304-s segments prior to Dip 2 (spectrum A) and during Dip 2 (spectrum B). Bottom: same spectra and model (see text), but with a Gaussian emission-line com- ponent included to fit the peaked residuals near 6.5 keV. The data shown are from PCU 0 only. 116 CHAPTER 7. ABSORPTION DIPS Disk Blackbody 3 Tin Rin cos1/2 θ Blackbody 4 R T Faint Component Joint fit1 1A–B 2A–B 1A–B 2A–B 1a–d 2a–d 3a–f Iron Line 2 no no yes yes yes yes yes (keV) 1.51+0.10 −0.17 1.51+0.04 −0.06 1.23+0.22 −0.15 1.35+0.06 −0.05 1.18+0.19 −0.16 1.26+0.14 −0.14 1.07+0.08 −0.09 (km) (keV) (km) 16.94+3.96 −2.20 17.54+1.25 −0.87 25.32+8.56 −7.03 22.45+1.83 −1.69 27.04+9.39 −6.33 25.12+7.66 −5.20 42.66+8.34 −6.60 2.20+0.62 −0.42 2.32+0.16 −0.15 1.79+0.33 −0.16 2.10+0.12 −0.09 1.72+0.22 −0.12 2.08+0.25 −0.15 1.74+0.09 −0.09 3.77+3.33 −2.55 3.47+0.86 −0.76 7.87+3.34 −4.14 5.23+0.92 −0.93 8.22+2.69 −3.03 5.46+1.08 −1.08 9.23+1.41 −1.32 f 5 0.107+0.005 −0.006 0.103+0.003 −0.003 0.096+0.007 −0.006 0.092+0.002 −0.002 0.103+0.009 −0.011 0.117+0.013 −0.014 0.054+0.005 −0.005 N (2) H /1022 6 (cm−2) 0.00+0.07 −0.00 0.03+0.28 −0.03 0.00+0.10 −0.00 0.00+0.10 −0.00 0.00+0.44 −0.00 0.15+0.70 −0.15 0.00+0.14 −0.00 7 χ2 r 3.89–4.07 7.58–7.84 1.44–1.62 1.87–2.15 0.99–1.20 0.80–1.12 1.27–1.34 joint fit parameters of the bright and faint components of spectra from Table 7.1: Top group: outside (A) and inside (B) dips 1 and 2. Middle group: the same, but with an added Gaussian line. Third group: joint fit parameters for the four to six 16-s spectra for dips 1, 2, and 3. The absorption parameters (N (1) H ) for the bright component are given in Tables 7.2 (A–B) & 7.4 (a-f) and the iron line parameters are given in Table 7.3. Spectrum N (1) H /1022 (cm−2) Spectrum N (1) H /1022 (cm−2) No Gaussian 1A 1B 1A 1B 2.83+0.57 −0.43 176+12 −13 2A 2B With Gaussian 2A 2B 3.81+0.71 −0.81 184+15 −14 3.50+0.29 −0.34 306+21 −19 4.30+0.26 −0.26 283+16 −14 Table 7.2: Effective hydrogen column density of the bright component outside (A) and inside (B) Dips 1 and 2 (from fits with and without a ∼6.5 keV Gaussian emission line). 1 In joint fits (using the model given in equation 7.1), only the column density (N (1) H ) of the bright component is allowed to vary between spectra (see Tables 7.2 and 7.4). Errors quoted are 90% confidence limits for a single parameter (∆χ2 = 2.7). 2 A Gaussian emission line (∼6.5 keV) was included in the final five fits (see Table 7.3). 3 For the disk blackbody, the inner radius of the accretion disk (times cos1/2 θ, where θ is the disk inclination angle and θ = 0 is parallel to the line of sight) is given for a distance of 8 kpc. 4The blackbody radius is given assuming a distance of 8 kpc. 5 The bright and faint components are assumed to have the same disk blackbody and blackbody parameters, but the faint component is scaled down by a factor f . 6 Absorption column density on the faint component (Hydrogen atoms per cm2). 7 Reduced χ2 = χ2/dof , where dof = degrees of freedom = the number of spectral bins (∼54/spectrum) minus the number of fit parameters. 7.4. EVOLUTION OF ENERGY SPECTRA DURING DIPS 117 The column density of the bright component outside dips (spectra 1A and 2A; see Table 7.2) is about 3×1022 cm−2, just slightly above the interstellar value. The column density during the low segments of dips (spectra 1B and 2B) is extremely high: 184×1022 cm−2 for spectrum 1B and 283×1022 cm−2 for spectrum 2B. This high column density value is required to produce the observed reduction in flux at 20 keV and is sufficient to render the absorbed component totally negligible at lower energies. Thus, the flux in spectra 1B and 2B below 5 keV is due entirely to the faint component. 7.4.4 Iron Emission Line The fits for both dips show similar residuals for the spectra from outside and inside the dip (Figures 7- 7 and 7-8). These residuals are also similar to those obtained in fits of spectra from the horizontal branch or upper normal branch in the study in Chapter 6. In particular, the fits of these spectra all show peaked residuals near 6.5 keV. In the two spectra from the bottom of dips (spectra 1B and 2B), the feature is prominently visible as a "bump" in the spectra themselves (see Figures 7-7 and 7-8). The parameters of this feature are best constrained when only the faint component is present. The line component is probably from iron Kα emission, which occurs at 6.4 keV for neutral iron and higher energy for ionized iron. The fact that these residuals in the dip spectra are very similar in absolute flux level to the residuals immediately outside the dips strongly suggests that the line feature is associated with the faint component. For example, iron fluorescence might occur during the scattering process. Based on the similar peaked residuals inside and outside the dips, a Gaussian emission line component was added near 6.5 keV (see the bottom panels of Figures 7-7 and 7-8). The addition of the line does reduce the peaked residuals near 6.5 keV, but does not improve the residuals (of similar strength but indicating data below the model) between 4–5 keV. The results of the joint fits, now including the Gaussian line, are listed in the second group in Tables 7.1 and 7.2, showing that the other model parameters are generally consistent with the previous fits within the 90% confidence limits. Table 7.1 also shows that a significant improvement in the reduced χ2 value was achieved in both cases. We will continue to include a Gaussian line in all subsequent fits. The parameters for the Gaussian line are shown in Table 7.3. Although the best-fitting centroid energy of the line is close to 6.6 keV, the Gaussian width is large (probably due to the relatively coarse energy resolution of the PCA: ∼1 keV FWHM at 6 keV), so that we cannot confirm whether the line is from neutral or partially ionized iron or whether it is actually composed of multiple unresolved narrow lines. 118 CHAPTER 7. ABSORPTION DIPS E (keV) Joint Fit 1A–B 6.59+0.12 −0.13 2A–B 6.58+0.06 −0.07 6.59+0.16 1a–d −0.15 6.45+0.16 2a–d −0.20 6.46+0.23 3a–f −0.27 σ normalization (keV) (photons cm−2 s−1) 0.18+0.28 −0.18 0.29+0.10 −0.13 0.13+0.29 −0.13 0.47+0.29 −0.32 0.39+0.32 −0.39 0.018+0.007 −0.004 0.016+0.002 −0.002 0.015+0.002 −0.002 0.025+0.012 −0.008 0.012+0.004 −0.003 Table 7.3: Gaussian emission-line parameters for the joint fits of spectra outside and inside dips (A–B) and during entry into a dip (a–d, a–f). In each case, the same Gaussian line parameters were used for each spectrum in a joint fit. 7.4.5 Spectral Evolution During Dip Transitions Through its large collecting area, the PCA provides good-quality spectra of bright sources such as Cir X-1 every 16 s. This enables us to study the detailed evolution of the spectrum during the transitions between the high and low states outside and inside dips. We selected four to six 16-s segments from the ingress of Dips 1, 2 and 3 (indicated as diamonds on Figures 7-2, 7-3, and 7-4). The 16-s spectra from each of these dips are shown in Figures 7-9, 7-10, and 7-11, along with the model curves from joint fits of the 4–6 spectra of each dip. As expected from the light curves of the dips (Figures 7-2, 7-3, and 7-4), the intensity in each case initially decreases at low energy, but then reaches a fixed level while the intensity at higher energy continues to decrease. Intermediate spectra in each figure show a "step" near 7.1 keV, which cannot be accounted for solely by the emission line, but is naturally fit by the iron K absorption edge of the absorption component and indicates a column density of NH > few×1023. As discussed above, the intensity at the bottom of the dip is significantly reduced (relative to outside the dip) even at 20 keV due to Thomson scattering, The fit parameters for the bright and faint components in each joint fit of the 16-s spectra (1a–d, 2a–d, and 3a–f) are shown in Table 7.1. The flux in the faint component relative to the bright component ranges from 5% and 12% in these fits, and the column density for the faint component is again consistent with zero. The Gaussian line parameters for these fits are listed in Table 7.3, and indicate a line centered at 6.4–6.6 keV. The best-fitting normalization (absolute flux) of the line differs by a factor of two among the three dips but is consistent with no change within the errors. For each set of spectra, the variable absorption column density of the bright component is shown in Table 7.4. In each case the column density increases from a relatively low value (NH = 4–9×1022) at the start of a dip to very high values (NH > 1024) at the lowest part of the dips. 7.4. EVOLUTION OF ENERGY SPECTRA DURING DIPS 119 Figure 7-9: Spectral fits for four 16-s segments (a–d) during the decline into Dip 1. Only the column density of the bright component varies between the four jointly-fit curves. The data shown are from PCU 0 only. 120 CHAPTER 7. ABSORPTION DIPS Figure 7-10: Spectral fits for four 16-s segments (a–d) during the decline into Dip 2. Only the column density of the bright component varies between the four jointly-fit curves. The data shown are from PCU 0 only. 7.4. EVOLUTION OF ENERGY SPECTRA DURING DIPS 121 Figure 7-11: Spectral fits for six 16-s segments (a–d) during the decline into Dip 3. Only the column density of the bright component varies between the six jointly-fit curves. The data shown are from PCU 0 only. 122 CHAPTER 7. ABSORPTION DIPS Spectrum 1a 1b 1c 1d N (1) H /1022 (cm−2) 4.16+1.10 −0.75 25.8+1.6 −1.3 43.6+1.7 −2.1 129+8 −8 Spectrum 2a 2b 2c 2d N (1) H /1022 (cm−2) 8.84+1.37 −1.09 25.8+1.4 −1.5 64.0+2.4 −2.5 242+83 −52 Spectrum 3a 3b 3c 3d 3e 3f N (1) H /1022 (cm−2) 8.31+1.09 −0.96 15.7+1.0 −1.1 33.2+1.4 −1.5 66.7+1.6 −1.5 127+8 −7 > 2×104 Table 7.4: Effective hydrogen column density responsible for the variable absorption in the bright component, due to photo-electric absorption and Thompson scattering, during 16-s spectra from Dips 1, 2, and 3. These fits show that the model developed above, where only the absorption column density on the bright component varies between all spectra of a given dip, is quite successful in reproducing the evolution of the spectrum throughout the dips. 7.5 Discussion The spectral evolution during dips in Cir X-1 produces curved tracks in CDs and HIDs. By identi- fying these tracks we can exclude absorption dip behavior from analysis intended to study Z-source spectral tracks, as we have done in Chapter 6. Similar absorption-dip behavior was recently observed in RXTE observations of the black-hole candidates GRO J1655-40 and 4U 1630-47 [36, 74]. Dips in these sources show evidence for a faint unabsorbed component and produced curved tracks in CDs and HIDs similar to those we observe in Cir X-1. Spectral analysis from an ASCA observation of Cir X-1 during an intensity transition also showed evidence for an unabsorbed component [12]. A partial-covering model was used to fit the energy spectrum (with two blackbodies), but it was noted that the model could be interpreted as a direct component plus a scattered component. The ASCA spectra also showed an iron K edge due to absorption with column densities near 1024 cm−2. Thus our results are generally consistent with the ASCA results, but we are able to demonstrate how the continuum spectrum evolves throughout dips. An iron line at 6.4 keV was seen in the low-state ASCA spectrum, but no line was detected at a similar flux level in the high-state spectrum [12]. Our analysis shows that an iron emission line appears to be associated with the faint component, and is present outside dips at the same flux level. 7.5. DISCUSSION 123 Since scattering by an extended cloud of material is probably responsible for the faint component, it is likely that the iron Kα emission is also produced as a result of the scattering process. The fact that absorption dips in Cir X-1 mainly occur within a day of phase zero and are then followed by significant flaring, suggests that the dips are associated with the mass transfer process. Thus, absorption might be due to the mass transfer stream itself or due to a bulge on the disk produced by the addition of matter. The intermittent nature of the PCA light curves near phase zero (Figure 7-1) indicate that we are observing the fine structure (temporal or geometrical) of the obscuring material. 124 CHAPTER 7. ABSORPTION DIPS Chapter 8 Multi-frequency Observations 8.1 Overview The 16.55-d X-ray period of Cir X-1 is also seen in the radio, infrared, and optical bands (see Chapter 1). We organized two multi-frequency campaigns, in 1996 May and 1997 June, to study correlated variability in different frequency bands. Such observations have the potential to constrain emission mechanisms and provide information about the mass accretion rate. For example, in Z source LMXBs the optical flux is believed to result from higher energy photons being reprocessed in the accretion disk. Thus optical intensity is taken as a measure of mass accretion rate, and has been used to argue that mass accretion rate increases along Z source spectral tracks from the horizontal branch, down the normal branch, and onto the flaring branch [28, 2]. 8.2 May 1996 Campaign The first campaign (see Table 8.1) occurred during 1996 May 7–14, covering a week including phase zero on May 11.413 UT (JD 2450214.913). Radio observations were made from South Africa with the 28-m dish at Hartebeeshoek Radio Astronomical Observatory (HartRAO) by George Nicolson. Infrared observations were made with the 1.9-m telescope at the South African Astronomical Obser- vatory (SAAO) by Ian Glass. In addition to the ongoing All-Sky Monitor observations, seven RXTE PCA (and HEXTE) observations were made with a combined exposure time of 54 kiloseconds. 125 126 CHAPTER 8. MULTI-FREQUENCY OBSERVATIONS Figure 8-1: X-ray, IR, and radio light curves of Cir X-1 from 1996 May 7–17. Phase zero is indicated by the vertical dotted line at day 214.413 (1997 May 11.413). PCA intensity is from the 2–32 keV band (16-s bins) and ASM intensity is from 2–12 keV (90-s exposures). Errors on the PCA data are negligible. Typical error bars are shown for the ASM, IR, and radio data: errors on the ASM data are typically 2–5 c/s, on the IR data are typically about 0.03 magnitudes (worse on days 214 & 215), and on the radio data are about 30 mJy. The average quiescent level of 29 mJy has been subtracted from the radio data. (Data provided by I. Glass (IR), G. Nicolson (radio), and the RXTE /ASM team.) 8.2. MAY 1996 CAMPAIGN 127 Frequency Band Radio IR X-ray X-ray Observatory Type of Observation HartRAO 8.5 GHz (3 samples/d) SAAO (1.9 m) RXTE RXTE 8.5 GHz & 5.0 GHz K-band photometry ASM PCA Dates ongoing May 10–15 May (7),8–12,(13) ongoing May 7–14 Table 8.1: Multi-frequency Observations of Cir X-1 from 1996 May. Dates in parentheses were cancelled due to poor weather conditions. 8.2.1 Radio (HartRAO) Radio intensity measurements of Cir X-1 at 8.5 GHz (3.5 cm) are routinely carried out as often as sev- eral times daily at HartRAO. In recent years the periodic flares have been weak, i.e. no flares above 100 mJy, compared to 200 mJy – 2 Jy between 1976–1986 (1 Jansky = 1 Jy = 10−26 W m−2 Hz−1). In fact, radio flares since 1986 have not been strong enough to update the radio ephemeris. In conjunction with our multi-frequency campaign, HartRAO observations were made 12–20 times per day at both 8.5 GHz and 5.0 GHz from 1996 May 10–15 and 5–7 times per day at 8.5 GHz for several days prior to and after that period. Each measurement takes about 8 minutes. The observations are made as drift scans, i.e., the telescope is stationary during a scan, which begins pointing ahead of the source. Because Cir X-1 lies in the Galactic plane, there is strong confusion from other sources such as H II regions and the supernova remnant with which Cir X-1 is associated. A quiescent average scan is subtracted from the data and then fit the telescope beam pattern at the position of Cir X-1. This method gives a residual flux relative to the quiescent background rather than the absolute flux, but is sufficient to monitor the flaring behavior. The 8.5 GHz data from JD-2450000.5 = 210–219 are shown in Figure 8-1. The error on a typical point is about 30 mJy (1 σ). The average flux from observations between phase 0.2 to 1.0 is 29 mJy, and has been subtracted from the data. There is no evidence for variability during that phase interval. Between days 215.5–218.0 (phase ∼0.05–0.2) the flux is higher. After subtracting the 29 mJy quiescent level from the data, then 66% of the data points lie between 2–5 sigma above the noise of 19 mJy. This is consistent with weak flaring observed in other cycles from earlier in 1996, but flaring may be occurring earlier in the cycle than the corresponding period in 1995, when the flaring was between phases 0.2–0.3. 128 CHAPTER 8. MULTI-FREQUENCY OBSERVATIONS 8.2.2 Infrared Photometry (SAAO) The infrared observations at SAAO provided K band (2.2 µm) photometry measurements 4–32 times per night for five nights from 1996 May 8–12 (see Figure 8-1). An aperture of 9 arcsec diameter was used. The K magnitude was 11.2 on May 8–9 (day 211–212) and gradually brightened by about 1 magnitude during May 10–12 (day 213–215) to 10.15 (K = 0 mag corresponds to a flux density of 6.3 × 10−23 W m−2 Hz−1, and other magnitudes are related to flux through m2 − m1 = −2.5 log[F2/F1]). The quiescent and flaring magnitudes are similar to those obtained by Glass during 1980–1993, but are somewhat lower than the high values (K=9.5 to 7.7) observed in the late 1970's [24]. Observing conditions during the first two and last nights were good and the errors are 0.03 mag or better. The data suggest a possible precursor event during day 214; however the observing conditions were not as good during the third and fourth nights. When the conditions were bad, Cir X-1 was measured relative to star J from the image in Whelan et al. [83], which was used as a standard taken to be K=9.82. 8.2.3 Correlated Multi-frequency Variability The X-ray and radio light curves in Figure 8-1 reach a peak value during the second day after phase zero, while the IR reaches a peak value only 0.5 d after phase zero. The PCA and IR data both show evidence for increasing variability before phase zero. Furthermore, both show a gradual rise in intensity during the period from day 213.0 to 213.15, shown in more detail in Figure 8-2. Fourier power spectra from each of the five PCA segments in Figure 8-2 show a narrow quasi-periodic oscillation peak at 26.0 Hz, 24.2 Hz, 25.4 Hz, 31.54 Hz, and 37.61 Hz (in time order of the five segments). This timing behavior indicates that increasing X-ray intensity corresponds to motion to the right or down the horizontal branch in a hardness-intensity diagram (see Chapters 5 & 6 and [63]). Thus, this segment shows IR flux increasing from the horizontal branch toward the normal branch, similar to the optical trend in Z sources [28, 2]. 8.3 June 1997 Campaign The second campaign (see Table 8.2) occurred during 1997 June 4–22, covering a half cycle before and after phase zero on June 12.506 UT (JD 2450612.006). Radio observations were made from South Africa at Hartebeeshoek Radio Astronomical Observatory (HartRAO) by George Nicolson. Infrared photometry measurements were made with the 1.9-m telescope at the South African Astronomical Observatory (SAAO) by Ian Glass. IR spectroscopy at the the ANU 2.3 m telescope in Siding Springs, Australia and optical spectroscopy at the Anglo-Australian Telescope (AAT, 3.9 m) were 8.3. JUNE 1997 CAMPAIGN 129 Figure 8-2: X-ray and IR light curves of Cir X-1 from day 212.93–213.19. X-ray intensity is from the 2–32 keV PCA band (16-s bins). A typical IR error bar of +/−0.03 magnitudes is shown. (IR data courtesy of I. Glass.) 130 CHAPTER 8. MULTI-FREQUENCY OBSERVATIONS Frequency Band Radio IR Optical X-ray Observatory Type of Observation HartRAO 8.5 GHz & 5.0 GHz SAAO (1.9 m) K-band photometry ANU (2.3 m) K-band spectroscopy AAT (3.9 m) spectroscopy RXTE ASM Dates ongoing June (10–14),15–16 June (12),20 June 4,(12) ongoing PCA (samples) June 4,6,8,18,19,20 PCA (∼continuous) June 11–17 Table 8.2: Multi-frequency Observations of Cir X-1 from 1997 June. Dates in parentheses were cancelled due to poor weather conditions. carried out by Rob Fender, Kinwah Wu, and Helen Johnston. In addition to the ongoing All-Sky Monitor observations, extensive RXTE PCA (and HEXTE) observations were made, with 56% coverage over 7 days around phase zero, and shorter observations before and after that week. The X-ray observations for this campaign are discussed in Chapter 6. 8.3.1 Radio (HartRAO) In conjunction with our multi-frequency campaign, coverage during the ongoing radio (8.5 GHz and 5.0 GHz) monitoring at HartRAO was increased. The 5.0 GHz radio data for 1997 May – August are shown in Figure 8-3, along with the ASM light curve and the PCA data from the 1997 June observations (study G) and the brief observations from 1997 May (study F). As in the previous year, the radio intensity was faint with only minimal variability (a constant DC bias level has again been subtracted). However, in at least two cycles (starting at days 611 and 661), including the cycle that was the main focus of this multi-frequency campaign, there does appear to be weak evidence for flaring at about the same time as the X-ray flaring observed with the PCA and ASM. 8.3.2 IR Photometry (SAAO) Infrared photometry observations (see section 8.2.2) were scheduled at SAAO for seven nights, from 1997 June 10–16 (JD-2450000.5 = 609–615). Due to poor weather conditions, only limited data were obtained, shown in Table 8.3. On June 14, the JHK magnitudes are similar to typical peak values observed since 1980. On June 15, the K magnitude is about 0.5 mag lower, but still well above the typical quiescent value of K≈11.5. 8.3. JUNE 1997 CAMPAIGN 131 Figure 8-3: X-ray (PCA and ASM) and radio (5.0 GHz) light curves of Cir X-1 from 1997 May 7 – August 25. Phase zero is indicated by the vertical dotted lines. PCA intensity is from the 2–32 keV band (16-s bins) and ASM intensity is from 2–12 keV (90-s exposures). Errors on the PCA data are negligible and errors on the ASM data are typically 2–5 c/s. For each 16.55-d cycle, the average radio flux for the phase interval 0.3 to 0.9 of that cycle was subtracted from the data; the bias removed was typically 50–60 mJy and a typical error bar is shown. (The radio data was provided by G. Nicolson, and ASM data by the RXTE /ASM team. The PCA data are from studies F and G; see Table 3.1.) 132 CHAPTER 8. MULTI-FREQUENCY OBSERVATIONS J H 11.97 10.68 JD 2450614.35 .42 .48 2450615.41 L 8.98 K 9.88 9.88 9.85 10.34 Table 8.3: Infrared JHKL photometry measurements from 1997 June 14 & 15. Errors are +/− 0.03, except on the L magnitude the error is +/− 0.05. 8.3.3 IR Spectroscopy (ANU) Infrared K-band spectroscopy was carried out at the ANU 2.3 m telescope at Siding Spring, Australia. Two nights were scheduled, near phase 0.0 on June 12 and near phase 0.5 on June 20. Observations on the first night were cancelled due to bad weather, but conditions were more favorable on June 20 and good quality K-band spectra of Cir X-1 were obtained (see Figure 8-4). The spectrum shows H Brackett γ and HeI emission lines. Both lines are red-shifted by about +430 km/s, and there is a hint of asymmetry on the blue (shorter wavelength) wing. 8.3.4 Optical Spectroscopy (AAT) Optical spectroscopy at the Hα region (6100–7200 A) was carried out with the AAT 3.9 m telescope. Observations were scheduled on two nights, near phase 0.5 on June 4 and near phase 0.0 on June 12. Like the ANU observations on June 12, the AAT observations on that night were cancelled due to bad weather. The observations on June 4 were successful, and six good-quality spectra were obtained. The average spectrum, normalized by the continuum, is shown in Figure 8-5. This spectrum shows a prominent Hα emission line, as well as weaker He I(6678 A) and He I(7065 A) lines. The Hα line is asymmetric, as suggested by previous observations [17, 47, 53, 83], and can be fit well by two Gaussians lines: a broad component blue-shifted from the rest wavelength by -300 km/s and a narrow component red-shifted from the rest wavelength by +375 km/s. The two narrow He I lines are red-shifted by an amount similar to the narrow Hα line, and thus all three show a similar red-shift to the lines in the IR spectrum in section 8.3.3. All three narrow lines have widths of about 9.5 A, or ∼400 km/s, while the width of the broad Hα component is 46 A, or 2100 km/s. 8.3.5 Comparison of the AAT Optical Spectrum to HST Results The optical spectrum obtained with AAT during our 1997 June campaign is comparable to the spectrum of Cir X-1 obtained with the Hubble Space Telescope (HST) Faint Object Spectrograph (FOS) in 1995 June by Mignani, Caraveo, & Bignami [47]. The HST spectrum also showed an 8.3. JUNE 1997 CAMPAIGN 133 Figure 8-4: Infrared spectrum of Cir X-1 from ANU on 1997 June 20 (about phase 0.5), showing He I (2.06 µm) and H Brackett γ (2.165 µm) emission lines. This spectrum was compiled from the sum of 16 120-s exposures. (Figure courtesy of H. Johnston.) 134 CHAPTER 8. MULTI-FREQUENCY OBSERVATIONS Figure 8-5: Optical spectrum of Cir X-1 from AAT on 1997 June 4 (about phase 0.5), showing a prominent Hα emission line, as well as weaker He I(6678 A) and He I(7065 A) lines. (Figure courtesy of H. Johnston.) 8.4. CONCLUSIONS 135 asymmetric Hα emission line, and was likewise fit by a broad and a narrow Gaussian line. The broad line is centered on the Hα rest wavelength and has a width of 11.4 A. The narrow component is red-shifted by 380 km/s and has a width of 3.32 A. Based on the HST results, the authors suggested that the broad line originates in the accretion disk, with temperatures of order ∼20000 K responsible for its large width. They further suggest that the narrow line arises from a "hot spot" where the accretion stream from the companion impacts the disk. Thus the narrow line, observed at phase zero with HST, was expected to be observed at varying blue/red-shifts at other orbital phases. The interpretation suggested for the HST spectrum is not compatible with the additional results obtained with AAT, which show the narrow line at almost the same wavelength as HST, but at phase 0.5. Furthermore, it is the broad component which has shifted, from the rest wavelength to −300 km/s blue-shifted. Clearly, further observations at various orbital phases are needed. If the narrow Hα component, and the other optical and IR lines, prove to be fixed at about +400 km/s, then that velocity might reflect the radial velocity of the system, providing further support for the hypothesis that Cir X-1 is a high-velocity runaway from SNR 321.9-0.3. Measuring the velocity of the shifting broad component might help map the binary orbit, about which very little is securely known. 8.4 Conclusions The X-ray light curves of Cir X-1 show significant variability on all time scales down to seconds (and even shorter time scales including the QPOs). Unfortunately, most observations at other wavelengths typically provide measurements on timescales of hours, allowing only coarse correlations to be stud- ied. Yet even those measurement can be quite valuable. For instance, our multi-frequency campaign from 1996 has shown that according to the ephemeris equation currently in use (equation 1.1), the X-ray, IR, and radio bands all show enhanced emission within a day of phase zero (all peaking after phase zero). The 1997 radio data also show weak evidence for flares after phase zero, coincident with those in the X-ray band. The correlated increase in IR and X-ray intensity over several hours shown in Figure 8-2 demonstrates the need for observations with improved time-resolution. Clearly a return of the radio intensity to its previous bright levels would be very useful in improving the orbital ephemeris equation (i.e., testing the change in period implied by the current quadratic term.) Bright radio flares would also be able to be studied in much greater detail than the current marginal detections, allowing short-term correlation with other wavelengths to be explored. Delays of specific events between wavebands would provide information on how photons are being 136 CHAPTER 8. MULTI-FREQUENCY OBSERVATIONS reprocessed from one band into another. We continue to maintain a Target of Opportunity proposal with RXTE to provide X-ray coverage, should bright radio flares return. As mentioned above, one promising follow-up project to these campaigns is to study the shifting optical/IR emission lines. If such work can provide information about the binary orbit (through Doppler mapping), then the X-ray data can be re-examined with more consideration of the geometry of the system. This work may begin as early as this summer. A promising radio-interferometry imaging project is the attempt to detect proper motion of Cir X-1 away from the nearby supernova remnant (SNR), as is expected in the runaway binary scenario (see Chapter 1). Based on the 700–2000 km/s velocity implied by its distance from the SNR, and assuming a distance of 8 kpc, suggests that proper motion might be of order 1′′ in 20 yr [67], which could be detected in only a few years of precise measurements. Confirmation of association with the SNR, and thus of a young age, would be highly significant to our understanding of the evolution of the system. Preliminary work has recently been started on this project (by R. Fender et al.), with Australia Telescope Compact Array (ATCA) images taken to establish the location of an appropriate point source to use as a reference for proper motion measurements. Such radio imaging may also provide further information on the jet-like structures seen inside the synchrotron nebula surrounding Cir X-1. Chapter 9 Mass Transfer in an Eccentric Binary 9.1 Overview Mass transfer in X-ray binaries generally occurs through Roche-lobe overflow or via a stellar wind. Significant stellar winds only occur in high-mass donor (O or B) stars. Although the X-ray properties of Cir X-1 most resemble those of a low-mass X-ray binary, the true nature of the donor star and the mechanism for mass transfer remain uncertain (see section 1.3.6). Furthermore, the high eccentricity inferred from the 16.55-d intensity cycle of Cir X-1 will have a significant effect on how mass transfer occurs. In this chapter, I explore the effect of eccentricity on mass transfer through both Roche-lobe overflow and a stellar wind. 9.2 Mass Transfer via Roche-Lobe Overflow Mass transfer in most low-mass X-ray binaries is believed to occur through Roche-lobe overflow, wherein one star fills its critical equipotential surface and matter from the stellar surface streams through the inner Lagrange (L1) point (a saddle point in the net gravitational and centrifugal potential between the stars) toward the other star. As matter flows in, the Coriolis force causes it to orbit the compact object, forming an accretion disk wherein matter loses angular momentum through viscosity and spirals inward. In an eccentric orbit, the situation is much more complicated, with such mass transfer limited to the time near periastron passages. To help understand the behavior of mass transfer in an eccentric binary, I developed a computer code to follow the trajectory of a 137 138 CHAPTER 9. MASS TRANSFER IN AN ECCENTRIC BINARY test particle started from the surface of a donor star and moving in an eccentric binary system. The particle was followed until it reaches the neutron-star accretion disk, returns to the donor, or leaves the system. 9.2.1 Equipotential Surfaces in an Eccentric Binary Orbit The standard Roche-lobe framework describes the potential of a binary pair of stars in circular orbit, co-rotating at the orbital frequency. In an eccentric orbit, not only does the separation between the stars vary, but so does the instantaneous angular frequency of the orbit, making co-rotation impossible. A generalized Roche potential for centrally condensed stars in an eccentric binary orbit has been derived by Avni [3]. A Cartesian coordinate system was used, with the origin at the center of mass of star 1, of mass M1, which rotates with an angular velocity (~ωrot) along the Z axis. The reference frame rotates about the center of mass of star 1 such that star 2 (of mass M2), always lies on the +X axis. The potential at a point (X,Y ) is: ψ = − GM1 d1 − GM2 d2 + GM2 D2 X − 1 2 ω2 rot(X 2 + Y 2) (9.1) where d1 and d2 are the distances from (X,Y ) to the center of mass of each star and D is the instantaneous separation of the stars. To determine the location of the instantaneous inner Lagrange point, this potential is maximized along the line connecting the centers of the two stars. Although this potential is derived for the frame rotating with the orbit (and thus with an angular frequency continuously changing to match the instantaneous Keplerian frequency) the location of the L1 point is specified by a distance from the center of mass of star 1 along the line between the stars, and is thus well-defined in an inertial frame. Likewise, the quantities in equation 9.1 necessary to calculate the location of the L1 point are distances and masses, which can be measured in inertial space, and the angular rotation frequency of star 1, which is also measured in the inertial frame. Thus, this equation can be used to locate the instantaneous L1 point in an inertial frame, which will prove convenient in the calculation below. 9.2.2 Keplerian Orbits Keplerian orbits can be parameterized by a quantity know as the eccentric anomaly, u, which evolves in time according to Kepler's equation: 2πt P = u − e sin u, (9.2) 9.2. MASS TRANSFER VIA ROCHE-LOBE OVERFLOW 139 where t is the time elapsed since periastron, P is the orbital period, and e is the eccentricity. The left-hand side of equation 9.2 is also know as the mean anomaly. The separation vector of the two stars is then given by: x(u) = a(cos u − e) y(u) = a√1 − e2 sin u, (9.3) where a is the semimajor axis of the orbit (half of the maximum separation distance of the stars), and periastron lies along the +x axis. For each time step, equation 9.2 is solved for u via Newton's method, and the positions of the two stars are calculated from equations 9.3. The velocities of the stars can be calculated by substituting du/dt, from the time derivative of equation 9.2, and u into the time derivatives of equations 9.3. 9.2.3 Trajectory of a Particle in the Binary Potential The trajectory of a particle of negligible mass m moving in the potential of a binary system can be calculated by integrating the instantaneous velocity and acceleration of the particle in the inertial (non-rotating) frame of reference as the stars move in their orbits. The instantaneous velocity (~vm) and acceleration are given by: d~rm dt = ~vm, and d~vm dt = ~Fm m = − GM1m (~rm − ~r1) ~rm − ~r13 − GM2m (~rm − ~r2) ~rm − ~r23 , (9.4) (9.5) where the subscripts 1, 2, and m refer to the two stars and the particle respectively, ~Fm is the gravitational force on the particle, M is the mass of a star, and ~r is a position vector. The x and y components of equations 9.4 and 9.5 are integrated by a Runge-Kutta algorithm [57] to obtain the trajectory of the particle. The initial coordinate of the particle is taken to be a point on the surface of the donor (whose radius is defined to be the size of its critical lobe at periastron and held constant around the orbit) on the line between the centers of the two stars. The initial velocity of the particle is the vector sum of the instantaneous orbital velocity of the donor, the velocity of a particle on the surface due to rotation of the donor, and possibly an additional radial velocity away from the surface (see below). 9.2.4 Final Outcome of Particle Trajectories The possible final outcomes of the particle trajectory are as follows: 1. It is assumed that the accretion disk almost fills the instantaneous critical lobe of the neutron 140 CHAPTER 9. MASS TRANSFER IN AN ECCENTRIC BINARY star at periastron (i.e., 80% of the distance between the neutron star and the L1 point at periastron), and that the disk maintains that size at all other orbital phases. The particle is assumed to have entered the disk if it passes closer than this distance from the neutron star at any point in the orbit. 2. The donor is likewise assumed to fill its critical lobe at periastron and maintain that radius throughout the orbit. The particle is considered to be recaptured if it re-enters that radius relative to the donor at any point in the orbit. 3. The particle is assumed to be ejected from the binary (or at least no longer a likely contributor to mass transfer) if its distance from the CM exceeds twice the semimajor axis of the orbit. 4. After two full binary orbits, if none of the above three cases have occurred, the outcome is undetermined. In all cases, the orbital period was taken to be that of Cir X-1, 16.55 days, and the mass of the neutron star was taken to be 1.4 M⊙. For simplicity, the donor star is assumed to rotate uniformly at the average Keplerian frequency (ωrot = 2π/16.55 days). Also, a donor mass of 1 M⊙ has been adopted for the discussion below, where results are shown for two eccentricities: 0.5 and 0.8. Such values might be close to actual values for Cir X-1, but are mainly chosen for illustrative purposes. Figure 9-1 shows the final outcome of particles which were released from the surface of the donor at different orbital phases (evenly spaced in time) of a system with e = 0.5. The top panels (Figures 9-1a, b) are for an initial radial velocity off the surface of 30 km/s, or about 15% of the escape velocity from the star (if it were isolated). In Figure 9-1a, one can see the arrival time as a function of start time, where the different symbols indicate where the particle entered the accretion disk and where it was recaptured by the donor. Only particles that begin transfer during the interval −0.9 to +0.3 days relative to periastron eventually enter the disk, and most of the particles arrive at the disk (Figure 9-1b) closely spaced in time near day 0.3 or shortly after. Figures 9-1c and 9- 1d show that similar behavior occurs for particles with a larger radial velocity from the surface of 60 km/s (∼30% of the escape velocity from the donor). The larger initial velocity allows particles to reach the disk over a somewhat longer portion of the orbit, and also results in most of the particles reaching the disk before periastron. The fact that mass transfer is most effective for particles released before periastron is related to the relative velocity of the stars in their eccentric orbit. Particles released as the stars approach periastron naturally move toward the neutron star, and are more likely to be captured by the disk. Particles released after periastron have to try to "catch up" to a receding neutron star, and after a certain point will simply be recaptured by the donor. The asymmetric profile of particles entering 9.2. MASS TRANSFER VIA ROCHE-LOBE OVERFLOW 141 Figure 9-1: Panels (a) and (c): stop time of particles in a system with e=0.5 (and P=16.55 d) versus time of release from donor (1 M⊙), with the final outcome (recapture or entry into disk) indicated. Panels (b) and (d): normalized rate of particles entering the accretion disk versus their arrival time relative to periastron. The top panels, (a) and (b), are for an initial radial velocity from the surface of 30 km/s, and the bottom panels, (c) and (d), are for 60 km/s. In (a) and (c), the solid diagonal line has a slope of unity and indicates zero elapsed time between release and the final outcome; the horizontal dashed line indicates a final outcome occurring at periastron. 142 CHAPTER 9. MASS TRANSFER IN AN ECCENTRIC BINARY the disk is also related to the relative motion of the stars. Particles released shortly before periastron are carried along the orbit of the donor and their arrival times are thus focussed near the time of periastron. Figures 9-2a and 9-2b show the outcome of a similar transfer scenario as Figure 9-1, but with a larger eccentricity (e = 0.8) and with each particle receiving an initial release velocity of 50 km/s radially away from the surface of the donor (again about 15% of the escape velocity). The higher eccentricity confines mass transfer to a narrower range of orbital phase, mainly a release within 0.15 d prior to periastron passage. Figures 9-2c and 9-2d also show the outcome of transfer in a system with e = 0.8, but with a larger radial velocity of 100 km/s (∼30% of the escape velocity from the donor). In this case the accretion rate shows two sharp peaks after periastron passage, separated by only 0.25 d (Figure 9-2d). The combined high orbital velocity approaching periastron in the eccentric orbit and the high initial velocity from the surface cause some particles to initially "overshoot" the neutron star, eventually to be captured in the disk, but only after a delay that concentrates the arrival time of many particles near day 0.4 Eight illustrative trajectories for some of the test particles shown in Figure 9-1c (e = 0.5), released at various orbital phases with vradial = 60 km/s, are shown in Figure 9-3. These plots show how the motion of the two stars in their eccentric orbits can produce a wide variety of trajectories for a particle transferred from the surface of the donor. 9.2.5 Comparison to the X-ray Intensity Profile The sharp rise time of the number of particles entering the accretion disk in Figures 9-1 and 9-2 resembles the sharp rise in X-ray intensity from Cir X-1 after phase zero (not necessarily periastron) in the ASM light curve (see Chapter 3). These peaks are quite narrow (generally only a fraction of a day) compared to the ∼3-5 days typically observed in the X-ray light curve. However, the actual intensity profile depends on additional factors, such as the duration of time that matter spends in the disk before falling onto the neutron star. These results serve primarily as an example of what can be done once more is known about the parameters of the Cir X-1 system, such as the actual eccentricity and the mass and stellar type of the donor. 9.3 Mass Transfer via a Stellar Wind When the donor star in an X-ray binary does not fill its Roche lobe, a stellar wind or other outflow must be responsible for mass transfer. This is the case in some high-mass X-ray binaries, where 9.3. MASS TRANSFER VIA A STELLAR WIND 143 Figure 9-2: Panels (a) and (c): stop time of particles in a system with e=0.8 (and P=16.55 d) versus time of release from donor (1 M⊙), with the final outcome (recapture or entry into disk) indicated. Panels (b) and (d): normalized rate of particles entering the accretion disk versus their arrival time relative to periastron. The top panels, (a) and (b), are for an initial radial velocity from the surface of 50 km/s, and the bottom panels, (c) and (d), are for 100 km/s. In (a) and (c), the solid diagonal line has a slope of unity and indicates zero elapsed time between release and the final outcome; the horizontal dashed line indicates a final outcome occurring at periastron. 144 CHAPTER 9. MASS TRANSFER IN AN ECCENTRIC BINARY Figure 9-3: Trajectory of a particle moving in an eccentric binary system (Mdon=1 M⊙, e=0.5), where the particle starts with an initial radial velocity of 60 km/s off the surface of the donor at a time t0 relative to periastron. The plot is in a non-rotating frame with the CM at the origin (+). The location of the particle, donor star, and neutron star at t0 are indicated by the diamond, asterisk, and solid dot respectively. The subsequent motion of the particle until enters the accretion disk or is recaptured is shown as a solid curve, and the motion of the donor and neutron star in their orbits during that time is indicated by a dashed or dotted ellipse segment respectively. 9.3. MASS TRANSFER VIA A STELLAR WIND 145 the donor star experiences significant mass loss via a wind from the surface. Since a high-mass companion for Cir X-1 cannot yet be ruled out, I also studied wind transfer in an eccentric binary. I developed a computer code to study the mass transfer rate as a function of orbital phase. The resulting theoretical mass accretion profiles can be compared with the observed X-ray intensity profiles, which at least in some circumstances may be a measure of the accretion rate. 9.3.1 Bondi-Hoyle Accretion Material in the stellar wind passing within a critical radius of the neutron star is sufficiently deflected by the gravitational field of the neutron star to eventually be captured and accrete onto the surface (first presented by Bondi & Hoyle in 1944 [9]). Since the wind material flows around both sides of the neutron star (in the orbital plane; the problem is actually three-dimensional) the net angular momentum of the captured matter is small and accretion occurs without forming an accretion disk. The critical radius, i.e., the maximum accretion radius (racc), is the radius inside which the gravitational potential energy due to the neutron star exceeds the kinetic energy of the particles in the wind due to the relative velocity of the wind and neutron star (~vrel = ~vwind − ~vorb) and the thermal particle motion (related to the sound speed vs): GMnsm racc ≃ 1 2 mv2 rel + 1 2 mv2 s , (9.6) where Mns is the mass of the neutron star and m is the mass of a wind particle (mainly protons). Solving equation 9.6 for the critical accretion radius gives The mass accretion rate is then racc ≃ 2GMns v2 rel + v2 s . Macc = ρ(r)vrelπr2 acc, (9.7) (9.8) where ρ(r) is the density of the wind at a distance r from the center of the donor star, i.e., the separation between the stars. Conservation of mass provides an expression that relates the mass loss rate of the donor due to wind ( Mw), vwind, and ρ(r): Mw = ρ(r)vwind4πr2. (9.9) 146 CHAPTER 9. MASS TRANSFER IN AN ECCENTRIC BINARY Combining equations 9.7, 9.8, and 9.9 gives the fraction of the wind captured by the neutron star: Macc Mw = G2M 2 ns r2 vrel vwind 1 (v2 rel + v2 s )2 . (9.10) 9.3.2 Wind Accretion Profile in an Eccentric Binary For a given set of binary parameters (masses, orbital period, and eccentricity), the positions and velocities of the two stars around their orbits can be calculated according to the expressions given in section 9.2.2. The mass of the neutron star is taken to be 1.4 M⊙ and a period of 16.55 days is used. High-mass stars such as OB supergiants typically have high wind speeds of 1000-2000 km/s. A wind speed of 1000 km/s is used in the case discussed below. The relative velocity between the wind and the neutron star is calculated from the wind speed and the relative orbital velocity of the stars. For a wind temperature T , the sound speed is given by vs =pγkT /m, where k is Boltzmann's constant and γ = 5/3 for a monatomic gas. A typical wind temperature of 10000 K thus corresponds to a sound speed of about 12 km/s. This is much less than the wind speeds in this calculation. So, the wind temperature does not contribute significantly to the final results, but a value of 10000 K has been adopted for concreteness. In this calculation, wind propagation delays are ignored, i.e. the wind velocity vector is calculated using the current position and velocity of the donor star. This simplification is valid when the wind speed is significantly larger than the orbital velocity of the donor. In the limit of very high wind speed, vrel ≈ vw ≫ vs, and so equation 9.10 reduces to G2M 2 ns (9.11) Macc Mw = 1 v4 w . r2 Thus, for a given wind speed, it is primarily the distance between the stars that determines the fraction of the wind captured. However, equation 9.10 is still used in the code, since the orbital motion of the neutron star will produce an asymmetry of the profile relative to phase zero. In an eccentric orbit the separation between the stars changes with orbital phase (θ) according to r(θ) = a 1 − e2 1 + e cos θ =(cid:18) P 2π(cid:19)2/3 G1/3(Mns + Mdon)1/3 1 − e2 1 + e cos θ , (9.12) where the Kepler's third law was used in the second equality. Thus for a fixed orbital period and high wind speed, the wind capture fraction is weakly dependent on the mass of the donor star, strongly dependent on eccentricity, and varies with orbital phase as (1 + e cos θ)2. Thus, only the normalization, and not the shape, of the wind accretion profiles change significantly with different 9.3. MASS TRANSFER VIA A STELLAR WIND 147 donor masses or (high) wind speeds. The profile of mass accretion rate versus orbital phase for a 5 M⊙ donor is shown in Figure 9-4 for three values of eccentricity (0.2, 0.5, 0.8) and three wind speeds (500, 1000, 2000 km/s). As expected, a low eccentricity produces a broad profile, while a high eccentricity shows a much narrower peak. The peak value occurs slightly after phase zero due to the relative velocity of the wind and neutron star near periastron, and this shift is largest when the wind speed is slowest. 9.3.3 Comparison to the X-ray Intensity Profile The X-ray light curve of Cir X-1 as observed by the ASM (see Chapter 3) shows that flaring is usually limited to a quarter to a third of each 16.55-d cycle, suggesting at least a moderate eccentricity. The flaring is complex and varies significantly from cycle to cycle. In contrast to the wind profiles in Figure 9-4, many X-ray cycles show an asymmetric profile, with the intensity declining more slowly than the rise. Some cycles even show a secondary peak well separated from the first. Such behavior cannot be accommodated in the simple wind model presented here. Moreover, the X-ray light curves show a substantial baseline intensity level that does not occur in simulations with a large eccentricity. 148 CHAPTER 9. MASS TRANSFER IN AN ECCENTRIC BINARY Figure 9-4: Wind accretion rate versus orbital phase for three eccentricities and wind speeds for a 5 M⊙ donor. Each curve is normalized by its peak value; the actual wind capture fraction (relative to the mass-loss rate of the donor) decreases significantly with higher velocity. Chapter 10 Eccentric Binary Evolution 10.1 Overview In order to explore the possible evolutionary history of Circinus X-1, I have developed a binary evolution code to model the evolution of a neutron-star and low-mass companion (M < 3 M⊙) in an eccentric orbit. Although a high-mass companion cannot be ruled out based on the observed optical and infrared magnitudes (see section 1.3.6), the X-ray behavior is most consistent with a low-mass X-ray binary, and thus I only consider systems with a low-mass companion in this model. This code was originally developed by C. M. Becker to model accreting white dwarfs in circular binary orbits [4]. In order to keep the model simple, yet include eccentricity, I assumed that mass transfer (via Roche-lobe overflow) occurs only at periastron and that the transfer is conservative (no mass lost from the system). I have also added the effects of tidal dissipation on the orbit and allow a system to evolve even when not transferring mass. I then used this code as the basis for a Monte Carlo population synthesis study of low-mass X-ray binaries that might resemble Cir X-1. I show that the number of systems in the Galaxy with parameters similar to those of Cir X-1 should be of order unity, consistent with the unique status of the source as an eccentric LMXB with a high accretion rate. 10.2 Binary Evolution Theory 10.2.1 Evolution of Binary Parameters The Keplerian elements of the binary system can be specified by the mass of the neutron star (Mns), the mass of the donor star (Mdon), the orbital angular momentum (Jorb), and the eccentricity (e). 149 150 CHAPTER 10. ECCENTRIC BINARY EVOLUTION The orbital period (Porb), angular frequency (Ωorb), and semimajor axis (a) are related to these four quantities by straightforward application of Kepler's third law 2π (cid:19)2 (cid:18) Porb = Ω−2 orb = a3 GMT and the expression for the orbital angular momentum Jorb = G1/2MnsMdonM −1/2 T a1/2(1 − e2)1/2, where MT = Mns + Mdon. (10.1) (10.2) Mass transfer will occur when the donor star fills its critical potential lobe. For simplicity, I assume that the mass transfer is conservative, i.e., all matter lost by the donor is accreted by the neutron star: δMns = −δMdon and δMT = 0. Furthermore, since the critical potential lobe will be smallest at periastron, it is assumed that mass transfer occurs only at periastron. This is likely to be a good approximation for large eccentricities but clearly cannot hold true in the limit of zero eccentricity. For the purpose of studying systems that resemble Cir X-1, systems with very low or zero eccentricity (i.e., e< ∼0.01) are not of interest. Because mass transfer depends on the size of the donor relative to its critical potential lobe, it is necessary to follow the evolution of the donor star radius. There are several means by which the donor radius may change, each of which will be discussed in more detail below. At this point I will only mention what those mechanisms are: (1) The radius of the donor changes over time due to the nuclear evolution of its core. (2) The star responds adiabatically to the removal of mass. (3) The radius, if not at equilibrium, adjusts toward the equilibrium radius on a thermal time scale. The net effect of these contributions on radius evolution can be expressed as: δRdon Rdon =(cid:18) δRdon Rdon (cid:19)nuc +(cid:18) δRdon Rdon (cid:19)th +(cid:18) δRdon Rdon (cid:19)ad . (10.3) Orbital angular momentum is lost through tidal dissipation, magnetic braking, and gravitational radiation. δJ J =(cid:18) δJ J (cid:19)tid +(cid:18) δJ J (cid:19)mb +(cid:18) δJ J (cid:19)gr . (10.4) Likewise, these mechanisms produce a change in eccentricity. Unlike angular momentum (which is assumed to be conserved during mass transfer), the eccentricity changes due to mass transfer. Thus, 10.2. BINARY EVOLUTION THEORY the total change in eccentricity is given by: δe e =(cid:18) δe e (cid:19) m +(cid:18) δe e (cid:19)tid +(cid:18) δe e (cid:19)mb +(cid:18) δe e (cid:19)gr . 151 (10.5) I will discuss each of these mechanisms in detail below. The final two terms in equation 10.5 are expected to occur on much longer time scales than the first two terms and are thus ignored. 10.2.2 Evolution of the Donor Star Radius 10.2.2.1 Nuclear Evolution The equilibrium radius of a star depends on its total mass, as well as the mass of its core produced by nuclear evolution. An expression for this radius is given by Rappaport et al. [60]: Req R⊙ ≃ 0.85(cid:18)Mdon M⊙ (cid:19)0.85 + 4950M 4.5 (1 + 4M 4 c ) c , (10.6) where Mc is the mass of the core expressed in solar units. As the core mass grows through nuclear evolution, the radius of the donor will also increase. Taking the derivative of equation 10.6 and dividing by the donor radius gives the fractional change in the radius due to nuclear evolution: Rdon (cid:19)nuc (cid:18) δRdon = δReq Rdon = Req − 0.85(Mdon/M⊙)0.85 Rdon (cid:20)4.5 − 16M 4 c (1 + 4M 4 c )(cid:21) δMc Mc (10.7) The nuclear evolution of the core is straightforward to compute since in a red giant the luminosity is generated from nuclear burning in a thin shell surrounding a helium core. The change in core mass is determined by the rate at which hydrogen is burned in a shell around the helium core. In the low-mass giants being considered here, hydrogen is fused into helium via the proton-proton chain, which converts 0.7% of the hydrogen mass into energy. Thus, the core gains mass at a rate proportional to the luminosity of the donor (Ldon). Becker gives an expression for this rate [4]: Mc = 1.47×10−11fc(cid:18) Ldon L⊙ (cid:19)(cid:18) X 0.7(cid:19)−1 M⊙ yr−1. (10.8) where X is the hydrogen abundance (mass fraction), taken to be ∼0.75 and fc is an ad hoc constant multiplicative factor of 2 to make this simple evolutionary model more consistent with the results of more sophisticated codes. An approximate analytic formula for the luminosity of a star on the red giant branch is given by 152 CHAPTER 10. ECCENTRIC BINARY EVOLUTION the following expression, adapted from the result of Eggleton [19]: Ldon M⊙ (cid:19)4 L⊙ ≃ 2(cid:18) Mdon + 2×105(Mc/M⊙)6 1 + 2.5(Mc/M⊙)4 + 3(Mc/M⊙)5 . (10.9) 10.2.2.2 Thermal Readjustment Toward Equilibrium If the radius of the donor star differs from its thermal equilibrium radius, it will adjust toward the equilibrium radius on a thermal timescale. For example, if the critical potential lobe of the star restricts its radius to be smaller than it would be if it were in thermal equilibrium, the star will expand on a thermal time scale, driving mass transfer. The following relation is used to describe the readjustment of the stellar radius of the donor on a thermal (Kelvin-Helmholtz) timescale: (cid:18) δRdon Rdon (cid:19)th ≃ Req − Rdon RdonτKH δt , (10.10) where τKH ≃ 3GM 2 7RL ≃ 1.5×107(cid:18) Mdon M⊙ (cid:19)2(cid:18) Rdon R⊙ (cid:19)−1(cid:18) Ldon L⊙ (cid:19)−1 yrs. (10.11) 10.2.2.3 Adiabatic Response to Mass Loss A star will respond adiabatically to mass loss. Rappaport, Joss, and Webbink [59] relate the resulting change of the radius to the change in mass through a factor ξad: Rdon (cid:19)ad (cid:18) δRdon = ξad δMdon Mdon . (10.12) Di Stefano et al. [15] performed detailed evolution studies with a Henyey-type code including mass loss. They found that for stars with a low core mass (Mcor < 0.2M⊙), the radius shrinks when mass is removed, but for higher core mass, the mass loss causes the star to expand (resulting in unstable transfer). They parameterized their results for ξad with respect to core mass: ξad = ξad"1 −(cid:18) Mc Mc(cid:19)2# , (10.13) where ξad = 4.0 and Mc = 0.2 M⊙. 10.2.3 Evolution of Orbital Parameters 10.2. BINARY EVOLUTION THEORY 153 10.2.3.1 Tidal Evolution In a binary system, tidal friction will dissipate energy while conserving the total angular momentum. In particular, in a system where one star is compact (e.g. a neutron star), the sum of the total orbital energy plus the rotational energy of the "donor" star (actually not necessarily donating matter in this discussion) will decrease while angular momentum is exchanged between the orbit and the rotation of the star. In the weak friction model, it is assumed that the star is tidally deformed into the equipotential surface it would have formed a constant time τ ago in the absence of the tidal friction, i.e., a tidal bulge that is offset from the line between the centers of the stars. The rate of change of the orbital and rotational parameters due to this tidal dissipation is given by Hut [29] (using variables as defined above): da dt = −6 GMdonkτ R3 don de dt = −27 GMdonkτ R3 don 1 + q a (cid:19)8 q2 (cid:18) Rdon a (cid:19)8 q2 (cid:18) Rdon 1 + q a (1 − e2)15/2 (cid:20)f1(e2) − (1 − e2)3/2f2(e2) (1 − e2)13/2 (cid:20)f3(e2) − (1 − e2)3/2f4(e2) 11 8 Ωrot Ωorb(cid:21) , Ωorb(cid:21) , Ωrot e (10.14) (10.15) and dΩrot dt = 3 where GMdonkτ 1 R3 don q2r2 a (cid:19)6 g (cid:18) Rdon Ωorb (1 − e2)15/2 (cid:20)f2(e2) − (1 − e2)3/2f5(e2) Ωrot Ωorb(cid:21) , (10.16) 64 e8 16 e6 + 25 8 e4 + 185 2 e2 + 255 f1(e2) = 1 + 31 f2(e2) = 1 + 15 8 e4 + 15 2 e2 + 45 16 e6 4 e2 + 15 8 e4 + 5 f3(e2) = 1 + 15 64 e6 2 e2 + 1 f4(e2) = 1 + 3 8 e4 f5(e2) = 1 + 3e2 + 3 8 e4. Also, q = Mdon/Mns is the mass ratio and Ωrot and Ωorb are the rotational and mean orbital angular velocities. The dimensionless radius of gyration, rg, is related to the moment of inertia of the donor star by I = r2 gMdonR2 donor. A typical value for the radius of gyration is r2 don, and the apsidal motion constant k is also related to the structure of the g = 0.1 for the low-mass relatively unevolved stars considered here. For stars with an outer convective zone (i.e. low-mass stars with M < ∼ 1–2 M⊙), Lecar gives an approximate expression for the product of the apsidal motion constant and the tidal time lag [41]: kτ ≈ 25 ηλvconv(km/s) g/g⊙ s, (10.17) 154 CHAPTER 10. ECCENTRIC BINARY EVOLUTION where g = GM/R2 is the gravitational acceleration. The convective velocity, vconv, is typically ∼1 km/s for a wide range of convective stars. Furthermore, η and λ are the fractional mass and fractional depth respectively of the outer convection zone of the donor star. Lecar gives a table of these two quantities for main-sequence stars with effective temperatures of in the range of T =5200– 7200 K [41]. To facilitate interpolation and extrapolation of values for ηλ, an analytic function was fit to the table entries, resulting in the following relation: log10(ηλ) =  10.66 − 2.7907 × 10−6 T 2 + 7.5778 × 10−10 T 3 − 5.899 × 10−14 T 4 T ≥ 5000K T < 5000K. −1.25375 (10.18) For stars with effective temperatures below 5000 K, the tidal circularization time scale is very short and systems containing such stars will circularize before other mechanisms affect the orbital evolution. In contrast, for stars with effective temperatures above 7200 K, the time scale is very long and tidal dissipation will not play a significant role in the evolution of the orbital parameters. 10.2.3.2 Effect of Mass Transfer on the Orbit The effect of mass transfer in a circular orbit is relatively simple to calculate: if angular momentum is conserved, then the change in the semimajor axis of the orbit depends uniquely on the change in the masses of the stars. Mass transfer onto a star less massive than the donor star will decrease their separation, and transfer onto a more massive star will increase the orbital separation. In an eccentric orbit, the angular momentum depends on both the eccentricity as well as the semimajor axis, and thus conservation of angular momentum is insufficient to determine how both orbital parameters change. So, Eggleton has derived a prescription for the effect of mass transfer on the orbital eccentricity [19]. Mass transfer that occurs at orbital angle θ (position angle of one star relative to the other, where θ = 0 corresponds to periastron) results in a change in eccentricity given by: δe = −2 (1 − q)(cos e + e) δMdon Mdon . (10.19) When mass is transferred at a constant rate around the entire orbit, the average change in eccentricity is zero, since over a complete orbit hcos θi = −e. In the other extreme, when mass transfer occurs only at periastron (as in a highly eccentric orbit), δe = −2 (1 − q)(1 + e) δMdon Mdon . (10.20) Since δMdon < 0 during mass transfer and q = Mdon/Mns, equation 10.20 shows that mass 10.2. BINARY EVOLUTION THEORY 155 transfer will decrease the eccentricity when Mdon > Mns and will increase the eccentricity when the mass ratio is reversed. 10.2.3.3 Magnetic Braking This section on magnetic braking is reproduced from Becker [4]. Most stars have some magnetic field, and the interaction of the field lines with the ejected matter increases the effective "lever arm" and, therefore, the amount of specific angular momentum carried away by the stellar wind. Thus a small amount of mass loss can, in principle, extract a significant amount of angular momentum from the rotation of the star. If tidal interactions between the donor and companion keep the donor star co-rotating with the orbit, then magnetic braking can result in angular momentum losses from the orbit. Skumanich [64] and Smith [66] independently determined a rotation rate/age relation for isolated stars, mostly with masses about 1 M⊙, in clusters of various ages. From this relation an empirical magnetic braking torque was inferred. Verbunt & Zwaan [82] applied this to binary systems as a way to drive angular momentum loss. I use a slightly modified version of this braking law as given by Rappaport et al. [58]: J (cid:19)mb (cid:18) δJ = −1.73×10−27fmb" R4 ⊙(Rdon/R⊙)γM 1/3 T G2/3MwdP 10/3 # δt , (10.21) where γ is the parameterized magnetic braking index (typically γ = 4), fmb is a multiplicative constant of order unity, and "cgs units" are used. In the process of applying magnetic braking to binaries, a number of liberties were taken in applying the results of Skumanich and Smith. The use of the magnetic braking law in binaries extends to shorter periods and different masses for stars than in the original data. The parameter γ is used in place of any real knowledge of how the braking depends on stellar radius since most of the stars used to develop the original relation were G stars with more or less the same radius. 10.2.3.4 Gravitational Radiation This section on gravitational radiation is reproduced from Becker [4]. Gravitational radiation arises from a mass distribution with a time varying quadrupole moment, such as a pair of stars orbiting each other. For such a simple geometry (a circular orbit), the rate of angular momentum loss is given by the Einstein quadrupole radiation formula: J (cid:19)gr (cid:18) δJ 32 5 (4π2)4/3 G5/3 c5 = − MwdMdon M 1/3 T P 8/3 δt , (10.22) 156 CHAPTER 10. ECCENTRIC BINARY EVOLUTION where P is the orbital period [39]. For the systems we will consider, with periods typically larger than 10 hours, gravitational radiation alone drives mass transfer too slowly to have any significant effect. 10.2.4 Stable Mass Transfer Mass transfer can act as a sort of regulator, preventing the donor radius from exceeding the critical potential lobe radius (stable transfer). If mass transfer is unable to prevent the donor from over- filling the critical potential lobe, transfer will be unstable. If mass transfer (or another mechanism) eventually results in the donor underfilling its critical potential lobe, mass transfer will stop. Thus, during stable mass transfer the donor radius can continuously adjust so that it always equals the Roche lobe radius. Based on the assumption of stable transfer, I will now derive the expression that governs the rate of mass transfer. For the case of circular orbits, the Roche-lobe radius is related to the semimajor axis of the orbit through a function f (q) of the mass ratio q = Mdon/Mns: An analytic approximation of this function has been derived by P. Eggleton [18]: RL = f (q) · a. f (q) = 0.49q2/3 0.6q2/3 + ln(1 + q1/3) . (10.23) (10.24) For non-zero eccentricity, a plausible approximation to the effective critical potential lobe size is obtained by replacing the semimajor axis in equation 10.23 with the instantaneous separation d. A more precise measure for size of the critical potential lobe can be obtained by calculating the location of points on the critical surface of an effective potential that accounts for the eccentricity of the orbit, e.g., as given by Avni [3] (see section 9.2.1). However, for simplicity and speed in the evolution code, I will use equation 10.23 with the instantaneous separation. In a Keplerian orbit, the separation as a function of angle (θ) from periastron is given by d = a (1 − e2) 1 + e cos θ . (10.25) Thus at periastron, d = a (1− e). As described above, mass transfer via Roche-lobe overflow requires that the radius of the donor star equal the radius of the critical potential lobe at periastron radius: Rdon ≡ RL = f (q) · a · (1 − e). (10.26) 10.2. BINARY EVOLUTION THEORY 157 Taking the logarithmic derivative of equation 10.26 shows how changes in these parameters are coupled: δRdon Rdon = (1 + q) d ln f d ln q δMdon Mdon + δa a − δe 1 − e . (10.27) The logarithmic derivative of f (q) is calculated from equation 10.24 to be: d ln f d ln q = 2 3  1+q1/3 0.6q2/3 + ln(1 + q1/3)  ln(1 + q1/3) − 0.5 q1/3  . (10.28) Since I have chosen to work with orbital angular momentum rather than the semimajor axis, the logarithmic derivative of the orbital angular momentum (equation 10.2) is needed: δJ J = δMns Mns + δMdon Mdon − 1 2 δMT MT + 1 2 δa a − e δe 1 − e2 . (10.29) The masses in this relation can be re-expressed in terms of only the donor mass and the mass ratio: δJ J = (1 − q) δMdon Mdon + 1 2 δa a − e δe 1 − e2 . (10.30) Combining equations 10.27 and 10.30 to eliminate the semimajor axis gives: δJ J =(cid:20)1 − q − 1 + q 2 d ln f d ln q(cid:21) δMdon Mdon + 1 2 δRdon Rdon + 1 2 e (1 + e) δe e . (10.31) Using equations 10.3, 10.4, and 10.5 to explicitly include each source of evolution into 10.31: 2(cid:18) δJ J (cid:19)mb + 2(cid:18) δJ J (cid:19)gr + 2(cid:18) δJ 2(cid:20)1 − q − J (cid:19)tid − 1 + q 2 e (1 + e)(cid:18) δe d ln q(cid:21) δMdon e (cid:19)tid −(cid:18) δRdon Rdon (cid:19)th Rdon (cid:19)nuc −(cid:18) δRdon e (cid:19) m (1 + e)(cid:18) δe Rdon (cid:19)ad +(cid:18) δRdon Mdon d ln f + e = , (10.32) where all the explicitly time-dependent terms have been moved to the left hand side, and terms directly proportional to mass loss appear on the right hand side. Now using equations 10.12 and 10.19 to replace the last two terms in equation 10.32 with their dependencies on mass transfer, we find 2(cid:18) δJ J (cid:19)mb + 2(cid:18) δJ J (cid:19)gr e J (cid:19)tid − (1 + e)(cid:18) δe + 2(cid:18) δJ (cid:20) ξad + 2(1 − q) − 2hcos θi + e 1 + e e (cid:19)tid −(cid:18) δRdon Rdon (cid:19)nuc −(cid:18) δRdon = Rdon (cid:19)th d ln q(cid:21) δMdon Mdon d ln f (1 − q) − (1 + q) . (10.33) The time-dependent left-hand side of this equation, referred to as the "numerator", governs the 158 CHAPTER 10. ECCENTRIC BINARY EVOLUTION time scale for mass transfer. The right-hand side depends on the mass transferred, rather than time. The term in square brackets on the right-hand side (the "denominator"), determines if mass transfer is stable. As mentioned in section 10.2.3.2, mass transfer limited to periastron results in hcos θi = 1, so that the second and third terms in the denominator cancel. For transfer around the entire orbit, when hcos θi = −e, the third term of the denominator vanishes. All of the terms in the numerator are time dependent and can be calculated from the expressions given in the discussion above. The expressions given for gravitational radiation and magnetic braking are strictly correct only for circular orbits. However, these mechanisms of angular momentum loss are small in a system with a relatively long orbital period such as Cir X-1 (days), and thus the circular-orbit approximation is sufficient to estimate the magnitude of these quantities. 10.3 Binary Evolution Code 10.3.1 Code Algorithm The parameters that need to be specified at the start of the evolution are the inital donor mass and the eccentricity and semimajor axis of the orbit. The initial core mass can be varied, but was generally taken to be zero and evolved as part of the code. The initial neutron star mass was taken to be 1.4 M⊙ in all cases. The initial spin of the donor was assumed to be slow relative to the orbital frequency (Ωrot/Ωorb = 1×10−3 for concreteness) and allowed to spin-up via tidal effects. The donor star was assumed to start in equilibrium (Rdon = Req). It generally underfilled its initial periastron critical potential lobe, but in some cases just filled the critical lobe. Mass transfer started immediately in the latter case. A minimum of 10 time steps are used per decade of time. Smaller time steps are chosen as necessary so that no parameter changes by more than 0.25% during each step, and 107 yr is the largest time step allowed. For each time step: 1. The new core mass is determined from the simplified stellar evolution code. 2. The equilibrium radius of the donor is determined from its total and core mass. 3. The change in donor radius due to nuclear evolution, thermal adjustment, and adiabatic re- sponse to mass loss is calculated. 4. The effects of tidal dissipation, magnetic braking, and gravitational radiation on the orbit are calculated. 10.3. BINARY EVOLUTION CODE 159 5. If the donor star fills its critical potential lobe, determine the amount of mass transfer necessary to keep the donor radius no larger than the radius of critical potential lobe (calculate numerator and denominator from equation 10.33). 6. Update all the system parameters based on the above changes. The evolution is stopped if any of the following conditions occur: 1. The donor star is stripped (Mdon − Mcor < 0.05 M⊙). 2. The mass of the donor star drops below 0.1 M⊙. 3. The denominator of equation 10.33 is negative (mass transfer is unstable). 4. The binary circularizes (e < 0.01), and thus is no longer a candidate to produce a system that resembles Cir X-1. 5. The binary comes close to being unbound. When a becomes extremely large and e approaches unity, the system has become nearly unbound or has unphysical parameters. Thus I stop the code when e > 0.99 and refer to such systems as being "unbound". 6. More than 10 billion years have elapsed (longer than the age of most stars in the Galaxy). 10.3.2 Results As mentioned in section 10.2.3.2, mass transfer tends to decrease the eccentricity while the donor star is more massive than the neutron star. In some cases, this can lead to circularization of the orbit. However, in many cases mass transfer eventually inverts the mass ratio (or Mdon < Mns to begin with) so that the eccentricity and the semimajor axis are increased during continued mass transfer, often nearly unbinding the system. Systems that start with the donor star in contact with its critical potential lobe transfer mass at a high rate and usually unbind (see equation 10.20) in ∼108 years. Figure 10-1 shows the binary evolution calculation results for a 2 M⊙ donor that starts with donor star in contact with its critical potential lobe. In this situation, the thermal expansion of the donor towards the equilibrium radius dominates over other mechanisms, as is evident from the panel in Figure 10-1 showing the various time scales (τth ≡ Rdon/( Rdon)th, τnuc ≡ Rdon/( Rdon)nuc, τcirc ≡ 1/ etid, and τKH is given by equation 10.11). (The thermal expansion time scale is related to the Kelvin-Helmholtz time scale, but is longer by a factor shown in equation 10.10.) Mass transfer initially decreases the eccentricity and semimajor axis (such that periastron sepa- ration remains unchanged since transfer occurs only at periastron). After 1.2×108 yr, 0.3 M⊙ has 160 CHAPTER 10. ECCENTRIC BINARY EVOLUTION been transferred, making the neutron star more massive than the donor. Continued mass transfer then gradually increases the eccentricity and semimajor axis (again such that periastron distance, a(1−e), remains approximately constant) until the system is nearly unbound (after about 3×108 yr). This system maintains a moderately high mass transfer rate during most of its entire lifetime, and its period increases to the level of Cir X-1 (16.55 d) and beyond (the time corresponding to a 16.55-d period is indicated with a dot on several of the curves in Figure 10-1). Thus this system might go through a phase that resembles Cir X-1. At the point where the period is near 16 days, its eccentricity is quite high, about 0.9, and the mass transfer rate is 3.5×109 M⊙/yr. The results of a calculation starting with a 1 M⊙ donor initially underfilling its critical potential lobe are shown in Figure 10-2. This system evolves uneventfully for over a billion years before the nuclear evolution of the core eventually brings the donor into contact with its critical potential lobe. (The time scales for magnetic braking and tides to affect the orbit are much longer than the nuclear time scale for most of the system's lifetime.) Once in contact, the donor begins mass transfer and continues for about 1.3×108 yr until the system is nearly unbound. This system would also go through a phase of moderate mass transfer (∼10−9 M⊙/yr) and high eccentricity with a period near 16 days. The results for mass transfer depend strongly on the size of the critical potential lobe, and thus depend on the separation of the stars at periastron (a(1− e)). Here I illustrate this effect by holding e constant and varying a. Figure 10-3 shows the evolution of eccentricity and mass transfer rate for four systems which differ only in their initial semimajor axes (initially, Mdon = 2 M⊙ and e = 0.9). For the system with the smallest orbit (a = 37 R⊙), the donor star initially just fills its critical potential lobe, similar to the system in Figure 10-1, and thus transfers mass continuously until it unbinds after about 1.8×108 yr. The second system in Figure 10-3 (a = 40 R⊙) initially underfills its critical potential lobe, similar to that in Figure 10-2. Nuclear evolution eventually brings the donor into contact, at which point mass transfer begins. Unlike the system in Figure 10-2, the donor in this binary is initially more massive than the neutron star, so the eccentricity dips before ultimately increasing towards unity. Skipping the third binary for the moment, the final system in Figure 10-3 has a large semimajor axis (a = 128 R⊙) and the donor is far from filling its critical potential lobe. The radius of the donor increases due to nuclear evolution of the core, resulting in a cooler effective temperature. The tidal time scale strongly depends on the temperature through the convection-shell parameters η and λ (see section 10.2.3.1). Thus, the tidal time scale drops rather sharply after ∼2×108 yrs, resulting in rapid circularization of the orbit. The donor star never fills its critical potential lobe; thus, no mass 10.3. BINARY EVOLUTION CODE 161 Figure 10-1: Results of a binary evolution calculation for a system started with the donor star just filling its critical potential lobe, with initial values: Mdon = 2 M⊙, a = 9.34 R⊙, and e = 0.6. The donor remains in contact with its critical potential lobe throughout the evolution; thus, the curve for Rlobe coincides with the curve for Rdon. A solid dot on some of the curves marks the time corresponding to an orbital period of 16.55 days. 162 CHAPTER 10. ECCENTRIC BINARY EVOLUTION Figure 10-2: Results of binary evolution calculation for a system started with the donor underfilling its critical potential lobe, with initial values: Mdon = 1 M⊙, a = 27 R⊙, and e = 0.9. 10.4. MONTE-CARLO POPULATION SYNTHESIS 163 transfer occurs. The third binary in Figure 10-3 (a = 81 R⊙) is intermediate between the second and fourth. It initially underfills its critical potential lobe, but comes into contact though nuclear evolution. The eccentricity initially dips and then rises, as in the second system. However, the mass transfer eventually results in reduced tidal time scale (shorter than in the fourth system). Tidal effects then rapidly reduce the eccentricity and the orbit circularizes, as in the fourth system. Tidal circularization also approximately doubles the periastron separation, pulling the donor out of contact with its critical potential lobe and turning off mass transfer about 4×106 yrs after it began. 10.4 Monte-Carlo Population Synthesis 10.4.1 Systems Resembling Cir X-1 The results of individual binary evolution calculations show that systems resembling Cir X-1 can be produced. However, an entire population of LMXBs must be studied in in order to determine the likelihood of actually observing such a system. Terman, Taam, & Savage [73] have conducted a population synthesis study of LMXBs, following an initial distribution of zero-age binaries (with the more massive star > 8 M⊙) from the main sequence, through a common-envelope phase, to just after the subsequent supernova explosion of the more evolved star. Systems that remain bound after the supernova explosion and do not immediately merge produce a population of low-mass main sequence stars (< ∼3 M⊙) with a neutron-star companion (pre-LMXBs) that are expected to eventually transfer mass through evolution of the orbit and/or the companion star. Using the binary evolution code described in the previous section, I have conducted a Monte- Carlo simulation of the evolution of pre-LMXBs throughout their entire lifetime as potential eccentric LMXBs. The initial population was drawn from the pre-LMXB population derived by Terman, Taam, & Savage [73], which yielded a relatively uniform distribution of systems in a specific range of e and log a. Thus, I chose the eccentricity from a linear distribution in the range e = 0.01 to 0.99, and the semimajor axis from a logarithmic distribution in the range a = 1.25 R⊙ to 400 R⊙. Combinations outside the pre-LMXB distribution of Terman, Taam, & Savage were discarded. The initial neutron-star mass was fixed to be 1.4M⊙ in all cases. Donor star masses were randomly chosen in the range Mdon = 0.8 to 3M⊙, and all donors were assumed to initially have zero core mass. Donor masses that corresponded to an equilibrium radius larger than the periastron critical potential lobe size were discarded as being systems that merge immediately or otherwise fail to contribute to the LMXB population 164 CHAPTER 10. ECCENTRIC BINARY EVOLUTION Figure 10-3: Eccentricity and mass transfer rate versus time for binaries with initial values Mdon = 2 R⊙, e = 0.9, and semimajor axes (a) between 37 and 128 R⊙. 10.4. MONTE-CARLO POPULATION SYNTHESIS 165 For this study, 105 systems were evolved as described in the previous section. The following criteria were established to classify a system as one that resembled Cir X-1: an orbital period within ∼20% of 16.55 days (13 d < P < 20 d), a high mass transfer rate (10−9 M⊙/yr < M < 10−7 M⊙/yr), and at least a moderate eccentricity (e > 0.1). Cir X-1 is 6–10 kpc away, yet still quite bright in X-rays, so any similar such system should also be visible from almost anywhere in the Galaxy. Since there is only one system detected with parameters meeting the above criteria, there is probably only about one such system in the Galaxy at this time. Of the 105 systems, 42% were discarded immediately due to the donor being too large for the orbit. Of the remaining systems, (pre-LMXBs) 15% resembled Cir X-1 at some point in their lifetime. On average, each of the pre-LMXBs eventually resembled Cir X-1 for 4.0×105 yrs. Terman, Taam, & Savage [73] predict a pre-LMXB birth rate of ∼3×10−6–10−5 yr−1 in the Galaxy. Multiplying the predicted birth rate by the average lifetime in a state resembling Cir X-1 results in an expectation of 1–4 systems similar to Cir X-1 in the Galaxy at any given time. Thus, the fact that Cir X-1 exists but is unique is consistent with binary evolution calculations. 10.4.2 Predicted System Parameters for Cir X-1 In addition to determining the lifetime of systems that resemble Cir X-1, the Monte Carlo simulation was also used to study the distribution of system parameters corresponding to 16.55 day orbital periods and 10−9 M⊙/yr < M < 10−7 M⊙/yr. Histograms of the number of binaries versus the initial donor mass and eccentricity for binaries that eventually reach P = 16.55 d and 10−9 M⊙/yr < M < 10−7 M⊙/yr are shown in Figure 10- 4. The initial donor mass distribution shows that most progenitors of systems resembling Cir X-1 have donor stars more massive than the neutron star. As I will show below, these systems come into contact and transfer mass due to the nuclear evolution of the donor. There is also smaller population of low-mass systems (the peak below 1 M⊙) which evolve into contact through angular momentum losses (tides and magnetic braking). The initial eccentricity distribution includes all values above e = 0.33 and is skewed toward higher eccentricity, peaking near e = 0.9. Figure 10-5 shows histograms of the number of systems versus various system parameters for binaries at the time they reach a state resembling Cir X-1: P = 16.55 d and 10−9 M⊙/yr < M < 10−7 M⊙/yr. The donor mass distribution peaks at about 1.25 M⊙ and is almost entirely confined to Mdon < 1.5 M⊙. This is a reflection of the fact that most systems that reach 16.55 d periods during mass transfer are in the stage of increasing orbital period, which can occur only after the donor has become less massive than the neutron star. The core mass distribution shows that most 166 CHAPTER 10. ECCENTRIC BINARY EVOLUTION Figure 10-4: Histograms of number of binaries versus initial Mdon and e which eventually reach P = 16.55 d and 1×10−9 M⊙/yr < M < 1×10−7 M⊙/yr. The y-axis normalization is arbitrary. 10.4. MONTE-CARLO POPULATION SYNTHESIS 167 Figure 10-5: Histograms of the number of binaries versus various system parameters at the time when systems resemble Cir X-1, i.e., P = 16.55 d and 1×10−9 M⊙/yr < M < 1×10−7 M⊙/yr. The y-axis normalization is arbitrary. 168 CHAPTER 10. ECCENTRIC BINARY EVOLUTION systems have developed a moderate core of 0.10–0.16 M⊙, indicating that nuclear evolution is driving transfer in most of the systems. A small fraction of donor stars with a low core mass correspond to the lower-mass systems which evolve mainly through angular momentum losses. The eccentricity distribution is constrained to high values, showing a peak between e = 0.83 − −0.93. A tail extending above e = 0.93 corresponds to the angular-momentum driven systems. The fact that no systems with low eccentricity have a 16.55-day orbit during mass transfer is due to the large periastron separation for a system with a ≈ 30–40 R⊙ (as required for two stars with nearly equal masses, ∼ 1 M⊙, and P = 16.55 d) and small eccentricity. For such systems, the critical lobe exceeds the donor size and no mass transfer will occur. The luminosity of the donor star in most systems is less than about 13 L⊙, as expected for a low-mass star that is not highly evolved. The optical and infrared magnitudes of the donor are discussed below. The mass transfer rate is roughly uniform between 10−8–10−7 M⊙/yr (3×10−8 M⊙/yr is approx- imately the Eddington limit for a 1.4 M⊙ neutron star) and falls off gradually below 10−8 M⊙/yr. There is also a narrow peak in the distribution at about 3×10−9 M⊙/yr, corresponding to the angular-momentum driven systems. The ages of systems when reaching a stage that resembles Cir X-1 are typically 4×107 to 1×109. Systems older than about 1×109 are mainly the angular-momentum driven systems. All of the systems are significantly older than the young age (∼30000–100000 yr) that would be implied for Cir X-1 if it were associated with the nearby supernova remnant. Thus this model implies that association with the supernova remnant is inconsistent with the assumption that Cir X-1 has a low-mass companion. The optical and infrared absolute magnitudes (V, R, I, J, & K) of the donor star the systems that resemble Cir X-1 are shown in Figure 10-6. Most donors have a V magnitude of 2–4. Assuming a distance of 8 kpc for Cir X-1 (see section 1.3.4), the distance modulus is -14.5 magnitudes. Using this and a visual extinction of AV ∼ 4 (see section 1.3.6), the mean absolute visual magnitude of V ∼ 3 in Figure 10-6 corresponds to a apparent magnitude of mV ∼ 21.5, which is very similar to the non-flaring component of the observed visual magnitude [50]. A higher visual extinction would further increase the apparent visual magnitude, making the donor star fainter than the observed magnitude and requiring a contribution from the accretion disk. The K-band absolute magnitude in Figure 10-6 is ∼1–3 for most donors. Applying the same distance modulus correction as above (and ignoring the much smaller extinction in at this longer wavelength), gives an estimated apparent K magnitude of about 16.5, which is more than 4 mag- nitudes fainter than the observed non-flaring value [24]. Thus, the K-band flux of a low-mass 10.4. MONTE-CARLO POPULATION SYNTHESIS 169 Figure 10-6: Optical and infrared absolute magnitudes (V,R,I,J,K) of the donor star in systems that resemble Cir X-1, i.e., those from Figure 10-5. The y-axis normalization of the histograms is arbitrary. 170 CHAPTER 10. ECCENTRIC BINARY EVOLUTION companion is much too faint to account for the observed flux, implying a significant contribution from the accretion disk (X-ray heating of the companion would be a variable contribution due to the different viewing angles around the orbit). Thus, the optical and infrared magnitudes of donor stars in the modeled systems are not inconsistent with (brighter than) the observed fluxes. 10.5 Conclusions The eccentric binary evolution code developed for this chapter has shown that a system such as Cir X-1 can be naturally produced with a low-mass companion. A population synthesis study based on this code has shown that the number of systems expected to resemble Cir X-1 is consistent with its unique status in the Galaxy. These results also strongly favor a high eccentricity and donor less massive than the neutron star as current system parameters. Currently the orbital period is the only well-established orbital parameter of the system, making detailed comparison of the evolution results to actual values difficult. For example, in most cases, systems reached 16.55-d periods during mass transfer while unbinding (P increasing). However, if the current ephemeris equation (equation 1.1) is to be taken literally, the period may be decreasing rapidly. If such a change becomes well established, then this would be a significant constraint on models for the system. Furthermore, if Cir X-1 truly is a young LMXB (105 yr) associated with a nearby supernova remnant (see Chapter 1), then it must have evolved through a different scenario (i.e., Mdon > 3 M⊙) than the LMXB model developed here, which generally requires ages of at least 4×107 yr. This evolution code (or a slightly modified version that allows circular orbits) can also be applied to other systems and can be used as the basis for a population synthesis study of LMXB properties in general. Some of this spin-off work is already in progress. Chapter 11 Conclusions The All-Sky Monitor light curve of Cir X-1 has shown consistent flaring above a steady, bright baseline, as well as significant variety among intensity profiles of each cycle. I have developed computer codes to calculate theoretical accretion-rate profiles for an eccentric binary in which mass transfer occurs via Roche-lobe overflow or via a stellar wind. These profiles generally show much less structure than the X-ray intensity profile. The theoretical profiles show that mass transfer is strongly peaked near periastron (in the case of a stellar wind and moderate eccentricity) or entirely confined to phases near periastron (for Roche-lobe overflow). The fact that the intensity of Cir X-1 maintains a high baseline level at all orbital phases might be related to the gradual emptying of the accretion disk (not included in the models). More difficult to understand is the strong mid-phase flares that sometimes occur in the X-ray light curve, when the neutron star and donor star are believed to be relatively far apart (near apastron). I developed a binary evolution code for an eccentric low-mass X-ray binary. About 15% of systems drawn from a post-supernova population of pre-LMXBs [73] were found to possess parameters that might be similar to those of Cir X-1 during some stage of their evolution. On average, each LMXB system resembled Cir X-1 for 4×105 yrs. This result implies that the number of systems in the Galaxy that should currently resemble Cir X-1 should be of order a unity, consistent with its unique status as an eccentric LMXB with a high accretion rate. Our multi-frequency campaigns showed that the current X-ray, radio, and infrared flaring all occur within a day of each other, shortly after phase zero. Unfortunately recent radio flares have been very weak, allowing only marginal detection of the flares. Clearly, a return of bright radio flares would be a welcomed opportunity to study correlated behavior in more detail. Infrared pho- tometry on timescale shorter that hours might also prove useful in future campaigns. Optical and 171 172 CHAPTER 11. CONCLUSIONS IR spectroscopy carried out during our second multi-frequency campaign showed evidence for emis- sion lines with high-velocity Doppler shifts (∼300–400 km/s). Compared with similar measurements made with the Hubble Space Telescope [47], these observations strongly suggest the need for a spec- troscopy study around an entire 16.55-d cycle in order to search for evidence of the binary orbit. Another promising future project is the attempt to detect the angular velocity (proper motion) of the system away from the supernova remnant with which it might be associated [67], thus establish- ing that the system is young. This project will require precise radio astrometry using VLBI (very long baseline interferometry). Preliminary work for the proper motion study was recently carried out at ATCA by R. Fender. In this project I have used extensive RXTE observations of Circinus X-1 to characterize much of the complex behavior exhibited by the source. Based on several timing and spectral studies, I classify the variability we have observed from Cir X-1 into three categories: (1) absorption dips, (2) motion along a Z-source hardness-intensity track, and (3) shifts of the Z-source hardness-intensity track. I will now summarize the main characteristics associated with each type of variability. Absorption dips generally occur within the 0.5 d before to 1.5 d after phase zero and last sec- onds to hours in duration. During dips, the intensity drops below the baseline intensity level while hardness ratios initially increase, but then return to a low value during strong dips. A faint com- ponent (∼10% of the total flux) remains unaffected by the heavy absorption, and is probably due to scattering by material distributed over a larger region than the X-ray source itself. This con- clusion is similar to that presented by Brandt et al. [12] based on ASCA observations of a single low-to-high transition. Here I have shown that multiple dips occur near phase zero in a complex and cycle-dependent manner, as might be expected for obscuration due to inhomogeneities in the mass transfer stream or intermittent "bumps" on the outer edge of the accretion disk. I have shown that the evolution of the energy spectrum throughout transitions into dips is consistent with a variably absorbed bright component and faint unobscured component, and that an iron emission line present both inside and outside dips appears to be associated with the faint (scattered) component. I have also identified the curved tracks that absorption dips produce in color-color and hardness intensity diagrams. The tracks are similar to those due to dips in the black-hole candidates GRO J1655-40 and 4U 1630-47 [36, 74]. Having recognized spectral tracks due to absorption, I was able to focus other studies on other behavior that may be more intrinsic to the source. I have identified much of the current behavior of Cir X-1 as that of a Z-source low-mass X-ray bi- nary. Hardness-intensity diagrams show that the track for Cir X-1 (e.g., Figure 6-4) differs somewhat from the canonical "Z" track (see Figure 1-1), but the timing characteristics on the branches strongly support the Z-source interpretation. In hardness-intensity diagrams, the "horizontal branch" (HB) 173 in Cir X-1 becomes almost vertical on the left end. Power density spectra of Cir X-1 from times cor- responding to the vertical portion of the HB show horizontal-branch QPOs that evolve in frequency between 1.3 Hz and 30 Hz (increasing in frequency moving down the branch). On the horizontal portion of the HB, this QPO occurs at ∼30–35 Hz and fades in strength, becoming only a "knee" in the power spectrum. A broad peak begins to arise near 4 Hz in power spectra from the horizontal portion of the HB. This peak becomes a prominent normal-branch QPO on the NB, and remains centered near 4 Hz. The flaring branch for Cir X-1 generally turns above rather than below the normal branch (although it's orientation depends on the energy bands used in the hardness ratio). No QPOs are present in power spectra from the flaring branch. The energy spectrum on the horizontal branch of Cir X-1 is fit well with a two-component model consisting of a soft disk blackbody and a higher-temperature blackbody (presumably from closer to the surface). In this model, motion along the HB is mainly associated with an increasing inner radius of the disk, implying that as the luminosity increases across the HB, the inner edge of the disk moves further away from the surface (contrary to expectations that the higher luminosity, and thus probably higher mass accretion rate, should result in a decrease in the radius of the inner edge of the disk). Energy spectra on the normal branch indicate that the hard blackbody fades away, leaving only the disk blackbody. On the lower NB, a feature in the spectrum develops above 10 keV. This feature becomes more prominent on the flaring branch. When the intensity of Cir X-1 is at the "quiescent" (1.0 Crab) level, it is generally on the vertical portion of the HB. Much of the "flaring" activity above the "quiescent" level corresponds to motion across the HB. From the HB/NB apex, the intensity can decrease either down the normal branch (with hardness correlated with intensity) or back across the HB. The lower end of the normal branch is often at a similar intensity level to that of the vertical HB; however, from the low-intensity end of the normal branch, small bursts or "mini-flares" occur, resulting in motion up the flaring branch and producing very low frequency noise (VLFN) in the power spectrum. Bildsten [7, 8] has proposed that VLFN in low mass X-ray binaries is produced by small patches on the neutron-star surface burning intermittently due to slow non-convective combustion, rather than the rapid ignition of the entire surface which occurs in type I bursts. The "flares" or "bursts" seen on the flaring branch in Z sources, and Cir X-1 in particular, may be part of a continuous range from weak VLFN bursts, to small flares, to type I bursts. I have shown that some of the variability in Cir X-1 is due to shifts of the "Z" track in hardness- intensity diagrams. These shifts are at least in part related to the cycling hardness ratios we observe with the All-Sky Monitor, i.e., gradual hardening of the spectrum during much of the 16.55-d cycle and more rapid hardening of the spectrum at the onset of the flaring state. There is a tendency for 174 CHAPTER 11. CONCLUSIONS horizontal-branch behavior to occur at the low-intensity, hard portion of the cycle, and normal or flaring-branch behavior to occur during high-intensity, soft portions of the cycle. Thus, evolution of the timing properties versus orbital phase is related to both the shifting Z-track and motion around the Z track. Clearly this can produce behavior more complicated than we observed in our first set of observations, presented in Chapter 4. Although there is still much to be understood about the physical mechanisms responsible for the correlated spectral and timing behavior in Cir X-1, this project has demonstrated that these unanswered questions for Cir X-1 are related to similar issues in an entire class of systems (Z sources). In fact, some of the unique features of the behavior of Cir X-1 may prove useful in future studies of the Z and atoll classes of low-mass X-ray binaries. If the baseline intensity level of Cir X-1 decreases from its current bright level, it will provide an opportunity to test the hypothesis that Cir X-1 shows atoll-source behavior at lower mass- accretion rates. Establishing the existence of a source showing the behavior of both classes under different conditions would be helpful in determining the physical distinctions between the classes (e.g., different magnetic fields would be ruled out). Aspects of the behavior of Cir X-1 that differ from that of typical Z sources may also provide important constraints on models for Z-source behavior. For example, models of horizontal-branch oscillations (typically observed at 13–60 Hz) will have to address what physical property (e.g., mass, spin, or magnetic field of the compact object) of Cir X-1 results in significantly lower-frequency HBOs (1.3–35 Hz). Also, the large shifts of the Z-track in hardness-intensity diagrams for Cir X-1, and the fact that such shifts are associated with the 16.55-d orbital cycle of Cir X-1, are not explained by the hypothesis that mass accretion rate increases along the Z track. Kuulkers et al. [33, 35, 34] have suggested that secular shifts in several Z sources are due to inclination effects. However, the shifts in those sources were not observed to occur periodically as in Cir X-1, nor were the shifts as large (see section 5.4.4). Another possibility is, if accretion occurs both through a disk and through a spherical inflow (or other means), then the accretion rate of each component may be responsible for one aspect of the observed behavior. In the past, the unique status of Cir X-1 among X-ray binaries has often resulted in its being relegated to the discussion of "other sources". The results presented here show that Cir X-1 has much in common with other low-mass X-ray binaries and that its unusual features may provide the opportunity to study a system with somewhat atypical physical parameters. Appendix A Fourier Timing Techniques A.1 Introduction Temporal analysis in X-ray astronomy is generally accomplished through examination of the Fourier power density spectrum (PDS) and complex cross spectrum. The following is a brief review of some of the standard techniques that are directly applicable to RXTE data (see van der Klis [76], e.g., for a more complete discussion of standard techniques). In addition, specific formulae are derived to show how the analysis software handles such complications as data gaps and variable number of detectors. The analysis software we developed at M.I.T. operates on time series data stored in "DS (DataS- tream)" format, which dynamically changes storage format to reduce file sizes, and allows data to be piped between programs to chain together several tasks. These features result in smaller files and faster processing than the standard FTOOLS package provided by NASA. Useful information can be stored as keywords in header blocks of DS files. Since the timing programs in particular rely heavily on using keyword parameters, keyword names will be referenced in the throughout this appendix, which is intended to provide documentation to our software. 175 176 APPENDIX A. FOURIER TIMING TECHNIQUES A.2 Power Density Spectra A.2.1 Leahy Normalization The discrete Fourier transform of a time series of N bins (keyword NUMPOINT) containing counts xk (k = 0, .., N − 1) is aj = N −1 Xk=0 xke2πijk/N , j = −N/2, ..., N/2 − 1. (A.1) Frequency bin j corresponds to a frequency of ν = j/T = j/N tb, where T is the total duration of the time series and tb is the duration of a single time bin. The normalization used by Leahy et al. [40] defines the power in frequency bin j to be Pj = 2 Nphaj2, j = 0, ..., N/2 (A.2) where Nph is the total number of photons in the transform. In this normalization, the distribution of powers for Poisson counting noise has a mean value of 2 and obeys the χ2 distribution with 2 degrees of freedom. This has the advantage of straightforward estimation of the significance of high points in the PDS. A.2.2 Estimate of the Poisson Noise Level Real time series data always contains counting noise (x = xsource + xnoise). This adds power P = Psource + Pnoise. In the absence of detector effects, Poisson counting noise produces a constant power equal to 2 in the Leahy normalization. However, detector dead-time effects result in a Poisson-level that varies with Fourier frequency. Zhang et al. [88, 89] (see also [51]) have calculated this effect to be < Pnoise(ν) >= 2[1 − 2r0τd(1 − τd 2tb )] − 2 N − 1 N r0τd( τd tb ) cos(2πtbν) + 2rvler0τ 2 vle[ sin(πτvleν) πτvleν )]2, (A.3) where r0 is the count rate per PCU detector, τd is the dead time per event (10 µs for the PCA), tb is the time bin size, and N is the number of time bins in the transformed segment. The last term accounts for dead time due to very large events (VLEs), occurring at a rate rvle per PCU with a dead-time window (τvle) of 55 µs or 155 µs for each VLE [89]. A.2. POWER DENSITY SPECTRA 177 A.2.3 Fractional RMS Variability Dividing a Leahy-normalized PDS by the mean count rate from the source results in a normalization giving the square of the fractional root-mean-squared variability per Hz. In this normalization, the square root of the integrated power under a PDS feature (such as a QPO peak) gives the fractional rms variability of the intensity due to that feature. The average total count rate per second (keyword RATE) is R = 1 tbPN k=1 xk N = ¯x tb . However, to obtain the mean rate of the source R ′ , the background b must be subtracted: ′ R = 1 tbPN k=1(xk − bk) N = 1 tb (¯x − ¯b) = R − B, (A.4) (A.5) where B is the average background rate per second (keyword BKGD). Generally, the average back- ground rate is estimated from slew data or other means modeled with software provided with the FTOOLS package. A.2.4 Estimates of Power Variance The variance of powers in the Leahy normalization is 4 (so σ = 2 = µ). The variance of the mean power from M consecutive frequency bins or from W PDSs is: σµ = 2 √M W . (A.6) However, intrinsic noise from the source often causes the actual variance of the averaged powers to exceed the theoretical variance. Thus, a useful error estimate is derived from the sample variance resulting from averaging M W powers Pi, i = 1, .., M W (dropping the frequency index for clarity): s2 = 1 M W − 1 MW Xi=1 (P 2 i − < P >2) The sample variance of the mean power is s2 µ = s2 M W = 1 M W (M W − 1) MW Xi=1 (P 2 i − < P >2) (A.7) (A.8) 178 APPENDIX A. FOURIER TIMING TECHNIQUES A.2.5 Data Gaps Ideally, each PDS should be derived from a continuous time series with no gaps. However, gaps are common in real data, making it necessary to work with partially filled (but not sparse) time series'. The simplest way to handle data gaps is to fill the empty bins with the mean number of counts per bin. (In practice the mean is subtracted from the data to suppress the DC term in the PDS, so unoccupied bins are filled with zeros.) To compensate for an incomplete time series, averages should be weighted by the fraction exposure f , the fraction of occupied bins in the time series (keyword FILL). I.e., in averaging M frequency bins from each of W power spectra, each with fractional exposure fw, Likewise, the sample variance is < P >= PM m=1PW m=1PW PM w=1 fwPmw w=1 fw s2 = which can be re-expressed as 1 w=1 fw − 1 PM m=1PW M W Xm=1 (P 2 mw− < P >2) Xw=1 (A.9) (A.10) (A.11) (A.12) s2 = PM m=1PW PM m=1PW w=1 fw w=1 fw − 1 (< P 2 > − < P >2) . Finally, the sample variance of the mean power is where < P 2 >= PM m=1PW m=1PW PM w=1 fw P 2 mw fw w=1 s2 µ = s2 w=1 fw PM m=1PW = < P 2 > − < P >2 PM m=1PW w=1 fw − 1 Both < P > and < P 2 > are saved in the PDS files and used in error estimates (COL keywords with PDS and PDS2 labels). A.2.6 Variable Number of Detectors For instrument safety, one or more of the five proportional counter units (PCUs) of the PCA may be shut off during an observation. Transforms are not performed for time segments which include a discontinuous detector turn-on/turn-off. However, separate PDSs with different numbers of PCUs turned on, ¯pw (keyword NUMPCU), can still be averaged if care is taken to handle the various parameters properly. When averaging W PDSs the combined average of the number of PCUs on, ¯pw (keyword NUMPCU), is weighted by the total number of occupied bins nw = WwNwfw (from A.2. POWER DENSITY SPECTRA keywords NUMFFT × NUMPOINT × FILL): 179 (A.13) = PW w=1 nw ¯pw ntot w=1 nw ¯pw w=1 nw ¯p = PW PW where ntot is the total number of occupied bins. Alternately, the PCU average can be expressed as p=1 npp/ntot, where np is the number of bins with p PCUs on. ¯p =P5 For the rms normalization, the relevant rates are those actually observed, regardless of the number of detectors involved. Thus, the combined average rate and background rate from W PDSs, each with average rate ¯xw and background rate ¯bw, are simply weighted by the number of occupied bins: ¯x = PW w=1 nw ¯xw ntot , ¯b = PW w=1 nw¯bw ntot . (A.14) Deadtime effects in each PCU are independent of the other PCUs, so the average rate per PCU r0 used in the Poisson-level estimate (equation A.3) should be weighted by the number of PCUs on: r0 = p=1 npp ¯xp p p=1 npp 1 tbP5 P5 = 1 tbP5 p=1 np ¯xp ntot ¯p (A.15) where ¯xp is the average number of counts in all time bins having p PCUs on. The numerator of the last term above is simply the total number of counts, which is equal to ntot ¯x. Thus, the average rate per PCU reduces to r0 = 1 tb ¯x ¯p = R ¯p . (A.16) The Poisson-level estimate also requires the average product per PCU of the VLE rate and the good count rate. For a constant number of PCUs p: < rvler0 >p= 1 t2 k=1 p vk p k=1 p b Pntot Pntot xk p = 1 t2 b Pntot k=1 vkxk ntotp2 , (A.17) where vk is the VLE rate per bin. The average over different numbers of PCUs must be based on the number of occupied bins for each p: < rvler0 >= P5 p=1 npp P5 p=1 npp < rvler0 >p p=1 npp < rvler0 >p . (A.18) = P5 ntot ¯p (The keyword VLEP saves this VLE product as ¯p2 < rvler0 > to be consistent with the RATE and BKGD keywords, which are not per PCU.) 180 APPENDIX A. FOURIER TIMING TECHNIQUES A.3 Complex Cross Spectra The Fourier transforms of a time series in two energy channels can be used to produce a complex cross spectrum in addition to the two power density spectra. Using the same notation as for the PDS (see above), the cross spectrum derived from two Fourier transforms aj and bj is defined as Cj = aj ∗bj, j = 0, ..., N/2. (A.19) A.3.1 Time Lags The argument of the complex-valued cross spectrum is the phase delay (in radians) between intensity fluctuations in the two channels (once again dropping the frequency subscript): δφ = arctan(cid:18) Im(C) Re(C)(cid:19) = arctan(cid:18) Re(a)Im(b) − Im(a)Re(b) Re(a)Re(b) + Im(a)Im(b)(cid:19) . (A.20) This is easily converted to a time delay via δt = δφ/2πν. A positive value of δt indicates that intensity fluctuations in the second channel lag those in the first. A.3.2 Coherence Function The magnitude of the cross spectrum can be used to derive the coherence function between two channels. The coherence function is a measure of the degree of linear correlation between the two time series at each Fourier frequency [81]. For noiseless signals, the coherence function is defined as γ2 = < a∗b > 2 < a∗a >< b∗b > , = < C > 2 < P1 >< P2 > (A.21) where P1 and P2 are the power spectra derived from a and b respectively and the angle brackets denote an average over m independent measurements (ideally m → ∞). For signals with Poison counting noise, the quantities used in equation A.21 should be for the source only, thus γ2 = < (a − anoise)∗(b − bnoise) > 2 < P1 − P1,noise >< P2 − P2,noise > γ2 = < a∗b > − < a∗bnoise > − < anoise ∗b > + < anoise < P1 − P1,noise >< P2 − P2,noise > ∗bnoise > 2 (A.22) (A.23) γ2 = < C > − < a∗bnoise > − < anoise ∗b > + < Cnoise > 2 < P1 − P1,noise >< P2 − P2,noise > . (A.24) According to Vaughan and Nowak [81], in the Gaussian limit the numerator in equation A.24 A.3. COMPLEX CROSS SPECTRA can be expressed as < Csource > 2 = < C > 2− < n2 >, 181 (A.25) where n2 ≡ (P1,sourceP2,noise + P1,noiseP2,source + P1,noiseP2,noise)/m. Thus, the intrinsic source coherence function can be expressed in terms of the observed powers and cross spectra and the estimated Poisson noise powers: γ2 = < C > 2 − (P1,sourceP2,noise + P1,noiseP2,source + P1,noiseP2,noise)/m < P1 − P1,noise >< P2 − P2,noise > . (A.26) 182 APPENDIX A. FOURIER TIMING TECHNIQUES Appendix B Standard PCA Light Curves This appendix contains standard 20-kilosecond light curves of all our 1996–1997 PCA observations of Cir X-1. By displaying all data with a uniform scale, the wide range of variability and intensity levels is apparent. Each 20-ks panel is labelled along the right side with the letter of the study with which it was associated (A–H, see Table 3.1) and several time labels referring to time zero of that panel: the truncated Julian date (JD-2450000.5), the calendar date (YR-MO-DAY), the mission elapsed time (MET = seconds since 1994 Jan. 1.0), and the orbital phase (see equation 1.1). The counting rates are derived from 16 s time bins over the full PCA energy band and are the average rates from PCUs 0, 1, and 2, which were operating during all observations (PCUs 3 and 4 were often switched off). A counting rate of 2.6 kcts/s corresponds to an intensity of ∼1.0 Crab (1060 µJy at 5.2 keV). Background is not subtracted but is negligible. The data have not been corrected for detector deadtime; such corrections would increase the baseline 1 Crab level by about 5% and the peak 2.7 Crab level by about 13%. 183 184 APPENDIX B. STANDARD PCA LIGHT CURVES 185 186 APPENDIX B. STANDARD PCA LIGHT CURVES 187 188 APPENDIX B. STANDARD PCA LIGHT CURVES 189 190 APPENDIX B. STANDARD PCA LIGHT CURVES 191 192 APPENDIX B. STANDARD PCA LIGHT CURVES 193 194 APPENDIX B. STANDARD PCA LIGHT CURVES 195 196 APPENDIX B. STANDARD PCA LIGHT CURVES 197 198 APPENDIX B. STANDARD PCA LIGHT CURVES Bibliography [1] M. Ali Alpar and Jacob Shaham. Is GX 5-1 a millisecond pulsar? Nature, 316:239–241, 1985. [2] T. Augusteijn et al. Coordinated X-ray and optical observations of Scorpius X-1. The Astro- physical Journal, 265:177–182, 1992. [3] Yoram Avni. The eclipse duration of the X-ray pulsar 3U 0900-40. The Astrophysical Journal, 209:574–577, 1976. [4] Christopher M. Becker. Studies of Supersoft X-ray Sources in the ROSAT Database of Pointed Observations. PhD thesis, Massachusetts Institute of Technology, 1997. [5] T. Belloni and G. Hasinger. An atlas of aperiodic variability in HMXB. Astronomy and Astrophysics, 230:103–119, 1990. [6] T. Belloni and G. Hasinger. Variability in the noise properties of Cygnus X-1. Astronomy and Astrophysics, 227:L33–L36, 1990. [7] Lars Bildsten. Rings of fire: Nuclear burning as the origin of sub-Hertz noise and weak X-ray bursts in accreting neutron stars. The Astrophysical Journal, 418:L21–L24, 1993. [8] Lars Bildsten. Propagation of nuclear burning fronts on accreting neutron stars: X-ray bursts and sub-Hertz noise. The Astrophysical Journal, 438:852–875, 1995. [9] H. Bondi and F. Hoyle. Monthly Notices of the Royal Astronomical Society, 104:273, 1944. [10] H. V. Bradt, R. E. Rothschild, and J. H. Swank. X-ray Timing Explorer mission. Astronomy and Astrophysics Supplement Series, 97:355–360, 1993. [11] Hale Bradt, Robert Shirey, and Alan Levine. Observations of Circinus X-1 with RXTE. In L. Scarsi, H. Bradt, P. Giommi, and F. Fiore, editors, Nuclear Physics B Proceedings Supple- ments, The Active X-ray Sky: Results from Beppo-SAX and Rossi-XTE. Elsevier Science B. V., 1998. 199 200 BIBLIOGRAPHY [12] W. N. Brandt, A. Fabian, T. Dotani, F. Nagase, H. Inoue, and T. Kotani. ASCA observations of the iron K complex of Circinus X-1 near zero phase: spectral evidence for partial covering. Monthly Notices of the Royal Astronomical Society, 283:1071–1082, 1996. [13] Deepto Chakrabarty and Edward H. Morgan. The 2 hour orbit of the bursting millisecond X-ray pulsar SAX J1808-3658. Submitted to Nature, 1998. [14] L. Chiapetti and S. J. Bell Burnell. X-ray observations of Circinus X-1 in its low state. Monthly Notices of the Royal Astronomical Society, 200:1025–1031, 1982. [15] R. Di Stefano, L. A. Nelson, W. Lee, T. Wood, and S. Rappaport. Luminous supersoft X-ray sources as type Ia progenitors. In R. Canal and P. Lapuente-Ruiz, editors, Thermonuclear Supernovae, page 147, 1996. [16] Richard G. Dower, Hale V. Bradt, and Edward H. Morgan. Circinus X-1: X-ray observations with SAS 3. The Astrophysical Journal, 261:228–250, 1982. [17] A. R. Duncan, R. T. Stewart, and R. F. Haynes. Hα position determination of the binary Circinus X-1. Monthly Notices of the Royal Astronomical Society, 265:157–160, 1993. [18] P. Eggleton. Approximations to the radii of Roche lobes. The Astrophysical Journal, 268:368– 369, 1983. [19] P. P. Eggleton. Evolutionary Processes in Binary and Multiple Stars. Cambridge Press, Cam- bridge, 1997. [20] B. Fortner, F. K. Lamb, and G. S. Miller. Origin of 'normal-branch' quasiperiodic oscillations in low-mass X-ray binary systems. Nature, 342:775–777, 1989. [21] J. Garc´ia-S´anchez and J. M. Paredes. Particle injection in the Circinus X-1 radio outbursts. Astronomy and Astrophysics, 323:867–880, 1997. [22] R. Giaconni. Scorpius x-1. Monthly Notices of the Royal Astronomical Society, 9:L439, 1962. [23] I. S. Glass. Variations of Circinus X-1 in the infrared. Monthly Notices of the Royal Astronomical Society, 183:335–340, 1978. [24] I. S. Glass. Long-term infrared behavior of Cir X-1. Monthly Notices of the Royal Astronomical Society, 268:742–748, 1994. [25] W. M. Goss and U. Mebold. The distance of Cir X-1. Monthly Notices of the Royal Astronomical Society, 181:255–258, 1977. BIBLIOGRAPHY 201 [26] Amos Harpaz, Saul Rappaport, and Noam Soker. The rings around the Egg Nebula. The Astrophysical Journal, 487:809–817, 1997. [27] G. Hasinger and M. van der Klis. Two patterns of correlated X-ray timing and spectral be- haviour in low-mass X-ray binaries. Astronomy and Astrophysics, 225:79–96, 1989. [28] G. Hasinger, M. van der Klis, K. Ebisawa, T. Dotani, and K. Mitsuda. Multifrequency observa- tions of Cygnus X-2: X-ray observations with Ginga. Astronomy and Astrophysics, 235:131–146, 1990. [29] P. Hut. Tidal evolution in close binary systems. Astronomy and Astrophysics, 99:126–140, 1981. [30] P. C. Joss. Helium-burning flashes on an accreting neutron star - A model for X-ray burst sources. The Astrophysical Journal, 225:L123–L127, 1978. [31] Paul C. Joss and Saul A. Rappaport. Neutron stars in interacting binary systems. Annual Review of Astronomy and Astrophysics, 22:537–592, 1984. [32] L. J. Kaluzienski, S. S. Holt, E. A. Boldt, and P. J. Serlemitsos. Evidence for a 16.6 day period from Circinus X-1. The Astrophysical Journal, 208:L71–L75, 1976. [33] E. Kuulkers and M. van der Klis. GX 340+0 with EXOSAT: Its correlated X-ray spectral and timing behavior. Astronomy and Astrophysics, 314:567–575, 1996. [34] E. Kuulkers, M. van der Klis, T. Oosterbroek, K. Asai, T. Dotani, J. van Paradijs, and W. G. H. Lewin. Spectral and correlated timing behavior of GX 5-1. Astronomy and Astrophysics, 289:795–821, 1994. [35] E. Kuulkers, M. van der Klis, and B. A. Vaughan. Secular variations in the Z source Cygnus X-2. Astronomy and Astrophysics, 311:197–210, 1996. [36] E. Kuulkers, R. Wijnands, T. Belloni, Mario Mendez, M. van der Klis, and J. van Paradijs. Absorption dips in the light curves of GRO J1655-40 and 4U 1630-47 during ourburst. The Astrophysical Journal, 494:753–758, 1998. [37] F. K. Lamb, N. Shibazaki, M. A. Alpar, and J. Shaham. Quasi-periodic oscillations in the bright galactic-bulge X-ray sources. Nature, 317:681–687, 1985. [38] P. Lamb and P. W. Sanford. Compton scattering effects observed in Sco X-1 and similar sources. Monthly Notices of the Royal Astronomical Society, 188:555–563, 1979. [39] L.D. Landau and E.M. Lifshitz. The Classical Theory of Fields. Pergamon, Oxford, 1962. 202 BIBLIOGRAPHY [40] D. A. Leahy, W. Darbro, R. F. Elsner, M. C. Weisskopf, P. G. Sutherland, S. Kahn, and J. E. Grindlay. On searches for pulsed emission with application to four globular cluster X-ray sources: NGC 1851, 6441, 6624, and 6712. The Astrophysical Journal, 266:160–170, 1983. [41] Myron Lecar, J. Craig Wheeler, and Christopher F. McKee. Tidal circularization of the binary X-ray sources Hercules X-1 and Centaurus X-1. The Astrophysical Journal, 205:556–562, 1976. [42] Alan M. Levine, Hale Bradt, Wei Cui, J. G. Jernigan, Edward H. Morgan, Ronald Remillard, Robert E. Shirey, and Donald A. Smith. First results from the All-Sky Monitor on the Rossi X-ray Timing Explorer. The Astrophysical Journal, 469:L33–L36, 1996. [43] W. G. H. Lewin, J. van Paradijs, and E. P. J. van den Heuvel, editors. X-ray Binaries. Num- ber 26 in Cambridge Astrophysics Series. Cambridge University Press, Cambridge, 1995. [44] Walter H. G. Lewin, Lori M. Lubin, Jianmin Tan, Michiel van der Klis, Jan van Paradijs, Wim Penninx, Tadayasu Dotani, and Kazuhisa Mitsuda. Quasi-periodic oscillations in the Z source GX 5-1. Monthly Notices of the Royal Astronomical Society, 256:545–570, 1992. [45] Yukihiro Makino. Circinus X-1 Observed with Ginga. PhD thesis, Osaka University, Japan, 1993. [46] K. Makishima, Y. Maejima, K. Mitsuda, H. V. Bradt, R. A. Remillard, I. R. Tuohy, and R. Hoshi. Simultaneous X-ray and optical observations of GX 339-4 in an X-ray high state. The Astrophysical Journal, 308:635–643, 1986. [47] R. Mignani, P. A. Caravoe, and G. F. Bignami. Assymmetric Hα disk emission dominates the HST/FOS spectrum of Circinus X-1. Astronomy and Astrophysics, 323:797–800, 1997. [48] M. Coleman Miller, Frederik K. Lamb, and Dimitrios Psaltis. Sonic-point model of kilohertz quasi-periodic brightness oscillations in low-mass X-ray binaries. The Astrophysical Journal, 1998. in press. [49] K. Mitsuda et al. Energy spectra of low-mass binary X-ray sources observed from Temna. Publications of the Astronomical Society of Japan, 36:741–759, 1984. [50] A. Moneti. Optical and infrared observations of Circinus X-1. Astronomy and Astrophysics, 260:L7–L10, 1992. [51] E. H. Morgan, R. A. Remillard, and J. Greiner. RXTE observations of QPOs in the black hole candidate GRS 1915+105. The Astrophysical Journal, 482:993–1010, 1997. BIBLIOGRAPHY 203 [52] P. Murdin, D. L. Jauncey, R. F. Haynes, I. Lerche, G. D. Nicolson, S. S. Holt, and L. J. Kluzienski. Binary model of Circinus X-1. Astronomy and Astrophysics, 87:292–298, 1980. [53] G. D. Nicolson, M. W. Feast, and I. S. Glass. Recent changes in the optical, infrared and radio emission from Circinus X-1. Monthly Notices of the Royal Astronomical Society, 191:293–299, 1980. [54] T. Oosterbroek, M. van der Klis, E. Kuulkers, J. van Paradijs, and W. G. H. Lewin. Circinus X- 1 revisited: Fast-timing properties in relation to spectral state. Astronomy and Astrophysics, 297:141–158, 1995. [55] W. Penninx, W. G. H. Lewin, J. Tan, K. Mitsuda, M. van der Klis, and J. van Paradijs. Quasi- periodic oscillations in the Z source GX 340+0. Monthly Notices of the Royal Astronomical Society, 249:113–121, 1991. [56] P. Predehl and J. H. M. M. Schmitt. X-raying the interstellar medium: ROSAT observations of dust scattering halos. Astronomy and Astrophysics, 293:889–905, 1995. [57] William H. Press, Brian P. Flannery, Saul A. Teukolsky, and William T. Vetterling. Numerical Recipies in C. Cambridge University Press, Cambridge, 1988. [58] S. Rappaport, P. C. Joss, and F. Verbunt. A new technique for calculations of binary stellar evolution, with application to magnetic braking. The Astrophysical Journal, 275:713–731, 1983. [59] S. Rappaport, P. C. Joss, and R. F. Webbink. The evolution of highly compact binary stellar systems. The Astrophysical Journal, 254:616–640, 1982. [60] S. Rappaport, Ph. Podsiadlowski, P. C. Joss, R. Di Stefano, and Z. Han. The relation between white dwarf mass and orbital period in wide binary radio pulsars. Monthly Notices of the Royal Astronomical Society, 273:731–741, 1995. [61] Eric M. Schlegel (editor). The RXTE Technical Appendix. NASA / Goddard Space Flight Center, Greenbelt, MD, 1995. Currently available via the world wide web or anonymous ftp at ftp://legacy.gsfc.nasa.gov/xte/nra/appendix f/. [62] Robert E. Shirey, Hale V. Bradt, Alan M. Levine, and Edward H. Morgan. X-ray timing and spectral evolution of Circinus X-1 versus orbital phase with the Rossi X-ray Timing Explorer. The Astrophysical Journal, 469:L21–L24, 1996. 204 BIBLIOGRAPHY [63] Robert E. Shirey, Hale V. Bradt, Alan M. Levine, and Edward H. Morgan. Quasi-periodic oscillations associated with spectral branches in RXTE observations of Circinus X-1. The Astrophysical Journal, 506, 1998. [64] A. Skumanich. The Astrophysical Journal, 171:565, 1972. [65] Alan P. Smale. A type i burst with radius expansion observed from Cygnus X-2 with the Rossi X-ray Timing Explorer. The Astrophysical Journal, 498:L141–L145, 1998. [66] M. A. Smith. Rotational studies of lower main-sequence stars. Publications of the Astronomical Society of the Pacific, 91:737–745, 1979. [67] R. T. Stewart, J. L.. Caswell, R. F. Haynes, and G. J. Nelson. Circinus X-1: A runaway binary with curved radio jets. Monthly Notices of the Royal Astronomical Society, 261:593–598, 1993. [68] R. T. Stewart, G. J. Nelson, W. Penninx, S. Kitamoto, S. Miyamoto, and G. D. Nicolson. Long- term infrared behavior of Circinus X-1. Monthly Notices of the Royal Astronomical Society, 253:212–216, 1991. [69] A. F. Tennant. High-frequency oscillations from Cir X-1. Monthly Notices of the Royal Astro- nomical Society, 226:971–978, 1987. [70] A. F. Tennant. Observations of 1.4-Hz quasi-periodic oscillations in Cir X-1. Monthly Notices of the Royal Astronomical Society, 230:403–414, 1988. [71] A. F. Tennant, A. C. Fabian, and R. A. Shafer. The discovery of X-ray bursts from Cir X-1. Monthly Notices of the Royal Astronomical Society, 219:871–881, 1986. [72] A. F. Tennant, A. C. Fabian, and R. A. Shafer. Observations of type I X-ray bursts from Cir X-1. Monthly Notices of the Royal Astronomical Society, 221:27P–31P, 1986. [73] James L. Terman, Ronald E. Taam, and Craig O. Savage. A population synthesis study of low- mass X-ray binary systems. Monthly Notices of the Royal Astronomical Society, 281:552–564, 1996. [74] J. A. Tomsick, I. Lapshov, and P. Kaaret. An X-ray dip in the X-ray transient 4U 1630-47. The Astrophysical Journal, 494:747–752, 1998. [75] Hiroshi Tsunemi, Shunji Kitamoto, Makoto Manabe, Shigengori Miyamoto, and Koujun Ya- mashita. All Sky Monitor on board the Ginga satellite and its performance. Publications of the Astronomical Society of Japan, 41:391–403, 1989. BIBLIOGRAPHY 205 [76] M. van der Klis. Fourier techniques in X-ray timing. In H. Ogelman and E. P. J. van den Heuvel, editors, Timing Neutron Stars, pages 27–69. Kluwer Academic Publishers, 1989. [77] M. van der Klis. Similarities in neutron star and black hole accretion. The Astrophysical Journal Supplement Series, 92:511–519, 1994. [78] M. van der Klis. Rapid aperiodic variability in X-ray binaries. In W. G. H. Lewin, J. van Paradijs, and E. P. J. van den Heuvel, editors, X-ray Binaries, pages 252–307. Cambridge University Press, 1995. [79] M. van der Klis. Kilohertz quasiperiodic oscillations in low-mass X-ray binaries. In R. Buccheri and J. van Paradijs, editors, The Many Faces of Neutron Stars, page (in press). Dordrecht: Kluwer, 1998. [80] B. A. Vaughan et al. Searches for millisecond pulsations in low-mass X-ray binaries II. The Astrophysical Journal, 435:362–371, 1994. [81] Brian A. Vaughan and Michael A. Nowak. X-ray variability coherence: How to compute it, what it means, and how it constrains models of GX 339-4 and Cygnus X-1. The Astrophysical Journal, 474:L43–L46, 1997. [82] F. Verbunt and C. Zwaan. Magnetic braking in low-mass X-ray binaries. Astronomy and Astrophysics, 100:L7–L9, 1981. [83] J. A. J. Whelan et al. The optical and radio counterpart of Circinus X-1 (3U 1516-56). Monthly Notices of the Royal Astronomical Society, 181:259–271, 1977. [84] Rudy Wijnands and Michiel van der Klis. KiloHertz quasi-periodic oscillations in the Z sources GX 340+0, Cygnus X-2, GX 17+2, GX 5-1, and Scorpius X-1. In S. S. Holt and T. Kallman, editors, Proceedings of the 8th Annual Astrophysics Conference in Maryland, 1997. [85] Rudy Wijnands and Michiel van der Klis. Discovery of the first accretion-powered millisecond X-ray pulsar. Submitted to Nature, 1998. [86] K. S. Wood et al. Searches for millisecond pulsations in low-mass X-ray binaries. The Astro- physical Journal, 379:295–309, 1991. [87] K. Yoshida, K. Mitsuda, K. Ebisawa, Y. Ueda, R. Fujimoto, and T. Yaqoob. Low state prop- erties of the low-mass X-ray binary X1608-522 observed with Ginga. Publications of the Astro- nomical Society of Japan, 45:605–616, 1993. 206 BIBLIOGRAPHY [88] W. Zhang, K. Jahoda, J. H. Swank, E. H. Morgan, and A. B. Giles. Dead-time modifications to fast Fourier transform power spectra. The Astrophysical Journal, 449:930–935, 1995. [89] W. Zhang, E. H. Morgan, K. Jahoda, J. H. Swank, T. E. Strohmayer, G. Jernigan, and R. I. Klein. Quasi-periodic X-ray brightness oscillations of GRO J1744-28. The Astrophysical Journal, 469:L29–L32, 1996. [90] W. Zhang, T. E. Strohmayer, and J. H. Swank. Discovery of two simultaneous kilohertz quasi- periodic oscillations in the persistent flux of GX 349+2. The Astrophysical Journal, 500:L167– L169, 1998.
astro-ph/0703610
2
0703
2007-03-26T11:01:26
Color Intensity Projections: A simple way to display changes in astronomical images
[ "astro-ph" ]
To detect changes in repeated astronomical images of the same field of view (FOV), a common practice is to stroboscopically switch between the images. Using this method, objects that are changing in location or intensity between images are easier to see because they are constantly changing. A novel display method, called arrival time color intensity projections (CIPs), is presented that combines any number of grayscale images into a single color image on a pixel by pixel basis. Any values that are unchanged over the grayscale images look the same in the color image. However, pixels that change over the grayscale image have a color saturation that increases with the amount of change and a hue that corresponds to the timing of the changes. Thus objects moving in the grayscale images change from red to green to blue as they move across the color image. Consequently, moving objects are easier to detect and assess on the color image than on the grayscale images. A sequence of images of a comet plunging into the sun taken by the SOHO satellite (NASA/ESA) and Hubble Space Telescope images of a trans-Neptunian object (TNO) are used to demonstrate the method.
astro-ph
astro-ph
Page 1 of 9 Color Intensity Projections  A simple way to display changes in astronomical images Keith S Cover1, Frank J. Lagerwaard1, Suresh Senan1 VU University Medical Center, Amsterdam1 Corresponding author: Keith S Cover, PhD VU University Medical Center Post Box 7057 1007 MB Amsterdam The Netherlands Email:  [email protected] Accepted for publication by  Publications of the Astronomical Society of the Pacific Abstract To detect changes in repeated astronomical images of the same field of view (FOV), a  common practice is to stroboscopically switch between the images. Using this method,  objects that are changing in location or intensity between images are easier to see because  they are constantly changing. A novel display method, called arrival time color intensity  projections (CIPs), is presented that combines any number of grayscale images into a  single color image on a pixel by pixel basis. Any values that are unchanged over the  grayscale images look the same in the color image. However, pixels that change over the  grayscale image have a color saturation that increases with the amount of change and a  hue that corresponds to the timing of the changes. Thus objects moving in the grayscale  images change from red to green to blue as they move across the color image.  Consequently, moving objects are easier to detect and assess on the color image than on  the grayscale images. A sequence of images of a comet plunging into the sun taken by the  SOHO satellite (NASA/ESA) and Hubble Space Telescope images of a trans­Neptunian  object (TNO) are used to demonstrate the method.   Subject headings: methods: data analysis, techniques: image processing, comets: general   Page 2 of 9 1 INTRODUCTION Detection of changes in a sequence of images of the same region of the sky is widely  used in astronomy to detect comets, asteroids, supernova, variable stars and other objects.  While there are several fully automated methods to detect these changes (Jewitt et al.  1998, Gladman 1998, Allen et al. 2001, Petit et al 2004), in many cases visual inspection  for these changes is still used both for detection and for reviewing the results of changes  detected automatically.  A commonly used method of visual inspection is to stroboscopically switch between  images. Alternately, if the comparison of only 2 or 3 images is required, the images are  sometimes displayed in the red, green and blue image planes (Buie et al. 2003). In the  resulting color image, constant objects appear grayscale while changing objects appear in  color. Buie et al. report the detection of changes by visual inspection to be much faster  for the three color method than the stroboscopic method. They also report the three color  method also allows the detection of some objects that have gone undetected by fully  automated methods. However, a major limitation of the three color method is that it is  limited to only 3 images because there are only 3 primary colors. Also, a major limitation  of any method that depends on visual inspection is that it requires human intervention  with all it complications including human error.      The method of color intensity projections (CIPs) was originally developed by the authors  for applications in medical imaging but it was apparent from the start that the method  may find useful applications in other fields including astronomy. The first publication of  CIPs demonstrated its use in presenting the motion of lung tumors and normal organs  during respiration (Cover et al. 2006). In the first version of CIPs, called percentage time  CIPs, the hue of the color is used to encode the percentage of time there is tissue in a  pixel. The second publication used the arrival time version of CIPs to visualize the flow  of blood into and out of the brain using an X­ray contrast agent (Cover et al. 2007). The  arrival time version of CIPs is calculated by adding a preprocessing step to the  percentage time method. All the details of the arrival time CIPs method are described in  the references but are also included in this paper for convenience.       While the primary mission of the SOHO satellite is to constantly monitor the sun, it has a  long history of observing comets both grazing the sun and plunging into it (Sekanina  2002; Biesecker 2002). In particular, the LASCO C3 coronagraph of SOHO (Morrill et  al.  2006) has observed many of these comets. With its field of view extending from 3.8  Page 3 of 9 to 30 solar radii and images acquired dozens of times a day, the C3 coronagraph provides  ideal data for applying CIPs.  The detection of trans­Neptunian objects (TNO) rely on the changes from one  astronomical image to the next. Hubble Space Telescope (HST) (NASA/ESA) images of  the binary TNO 1998 WW31 are also used to demonstrated the CIPs display method  (Veillet et al. 2002) 2 METHOD A total of 18 grayscale images of a comet plunging into the sun that were acquired by the  SOHO LASCO C3 coronagraph during November 2nd and 3rd, 2006 are used to  demonstrate the arrival time CIPs display technique. The images were cropped to 1014  by 690 pixels for processing and display purposes. In addition, six grayscale images of  the binary TNO 1998 WW31 acquired by the Hubble Space Telescope between July  2001 and February 2002 where also used for demonstration. The HST images were  cropped to 50 by 50 pixels for processing.  Grayscale images are characterized by a single value per pixel. While they are often  displayed in grayscale they can also be presented in false color. Color images are defined  by 3 values per pixel. The values may correspond to the intensity of the red, green and  blue components or some other numerical representation of color.    The first step in calculating an arrival time CIP image is to calculate a new image  sequence from the original time ordered sequence. The new sequence, which will be  referred to as the cumulative maximum intensity projection (cumulative MaxIP), is  simple to calculate. The first image of the cumulative MaxIP sequence is set equal to the  first image of the original image sequence. The second image of the cumulative MaxIP  sequence is set equal to the pixel by pixel maximum of the first two images of the  original image sequence. The third image of the cumulative MaxIP sequence is set equal  to the pixel by pixel maximum of the first three images of the original image sequence.  Following the same pattern, all the images of the cumulative MaxIP sequence can be  calculated.  The arrival time CIPs is then calculated by applying the percentage time CIPs method to  the cumulative MaxIP sequence. The first step is to scale the pixel values of the  cumulative MaxIP sequence to values between 0 and 1. This new image sequence, the  Page 4 of 9 normalized cumulative MaxIP sequence, is calculated by dividing each pixel in each  cumulative MaxIP image by the maximum value of all pixels in all images. The second  step is to calculate, on a pixel­by­pixel basis, the maximum, mean and minimum of the  normalized cumulative MaxIP sequence. The maximum (PMAX), mean (PMEAN) and  minimum images (PMIN) are then combined, on a pixel by pixel basis, using the following  equations to get the hue, saturation and brightness (HSB) of the composite color image:     Brightness = PMAX     Saturation = (PMAX­PMIN)/PMAX     Hue = 2/3*(PMEAN­PMIN)/(PMAX­PMIN) (1) The HSB for each pixel is then converted to the standard red, green and blue (RGB)  presentation of color, using standard subroutines (Arnold et al. 2005). The hue is  multiplied by 2/3 so there is a clear difference between the hue corresponding to the  earliest and latest arrival times.  Calculating a CIP for the 18 SOHO images with 1014 by 690 pixels took less than a  second to calculate on an AMD Duron running at 1GHz. If additional processing speed is  needed for a very large data set, the algorithm can be easily parallelized because each  pixel is calculated independently.  3 RESULTS AND DISCUSSION The upper 4 panels of Fig. 1 (a­d) show 4 of 18 images of a comet plunging into the sun  taken by the C3 coronagraph. Structures common to all the images are 1) the occulting  disk in the center of each image that blocks out the sun and 2) jets of gas emerging  radially out from the sun in several directions.  The comet is clearly visible on panels (b) and (c) to the lower right of the sun.  Extrapolating the motion of the comet makes the comet easier to find on panels (a) and  (d). Closer examination of the images also shows the apparent motion of the stars from  the left to the right on the images. The apparent motion of the stars is actually due to the  motion of the SOHO satellite, along with the earth, in orbit around the sun. As CIPs images present pixels that are unchanged in all the grayscale images as a shade  of gray in the color image, it is easy to pick out in the CIP image, panel (e), the structures  that are unchanged over the grayscale images. These are the occulting disk and the jets of  gas.  Page 5 of 9 The moving objects in the CIP image, such as stars, planets and the comet, change from  red to green to blue as they move across the FOV, as indicated by the color scale in the  upper left corner of the CIP. Careful examination of the CIP will show the very first  occurrence of an object is actually a shade of gray. This is another characteristic of  arrival time CIPs.  The comet plunging into the sun can be clearly seen to the lower left of the sun in the CIP  image. The uneven spacing between the various hues of the comet are due to the uneven  timing of the C3 images. The movement of the background stars from the left to the right  of the CIP standout because their tracks are all the same length.  Two planets can also be seen in the CIPs image. An object brighter than most of the stars  and with a track less than half the length of the stars can be seen to the right of the sun. It  is the planet Mars. The very bright object just to the upper left of the sun is also a planet.  The planet, which is Venus, can be easily seen to be moving in the opposite direction to  the stars.  Some objects only appear in a single hue, such as the yellow object near the base of the  rod supporting the occulting disk. The single hue indicates a one time event that only  appears on a single image, and the hue itself is indicative for the timing of this one time  event. Artifacts such as cosmic rays can be discerned by this characteristic.      Through the example of SOHO images of a comet plunging into the sun, CIPs has been  demonstrated as a simple and fast method for the detection of changes in astronomical  images. While CIPs has been demonstrated for medical imaging and astronomical  applications, there are likely a wide range of other applications where it may be applied.  The relative motion of the binary TNO in Fig. 2 is clearly shown by the CIP.  One of the  binary objects appears stationary as the grayscale images were shifted to align it on all 6  grayscale images. The orbit and direction of motion of the second object relative to the  first is clearly apparent from the CIP. As the grayscale images were not evenly spaced in  time, the hue gives the order rather than the precise timing of the object in the grayscale  images.  One of the strengths of CIPs is that many different types of information can be encoded  in the hue. As mentioned in the Methods section, the hue can be used to represent the  Page 6 of 9 percentage time an object spends at a particular location rather than the time it arrives at  the location. Another possible modification is to make the first occurrence of an object  have a red hue rather than the grayscale of arrival time CIPs. This can be accomplished  by modifying the percentage time CIPs so that the hue is determined by the time of the  peak signal in each pixel.  Another variation of CIPs that may be useful in the detection and visualization of  variable stars is introducing artificial motion to the images. This can be accomplished by  shifting each grayscale image by a pixel or two relative to the previous image before  calculating the CIPs. Thus, fixed stars will appear to move in a manner similar to SOHO  images as shown in Fig. 1. The artificial motion will make the variation of the intensity of  a star over time more apparent. If a star field is crowded, circular or spiral patterns of  artificial motion might be worth considering to reduce overlaps.    Whichever version of CIPs are being considered for an application, it would be best to  apply the various versions of CIPs to a variety of data sets to find which variation  performs the best for each application.   ACKNOWLEDGEMENTS This work was funded by the VU University Medical Center, Amsterdam. The VU  University Medical Center is pursuing a patent on color intensity projections on behalf of  the authors.  REFERENCES Allen R.L., Bernstein G.M. & Malhotra R.. 2001, ApJ 549, L241.  Arnold K., Gosling J. & Holmes D.  2005, Java Programming Language, The 4th Edition.  Prentice Hall PTR. Biesecker D. A., Lamy P., St. Cyr O. C., Liebaria A. & Howard R. A. 2002, ICARUS,  157, 323. Page 7 of 9 Buie M.W., Millis R.L., Wasserman L.H., Elliot J.L., Kern S.D., Clancy K.B., Chiang  E.I., Jordan A.B., Meech K.J., Wagner R.M. & Trilling D.E. 2003, Earth, Moon and  Planets, 92, 113. Cover K. S., Lagerwaard F. J. & Senan S. 2006, Int. J. Radiat. Oncol. Biol. Phys., 64,  954.  Cover K. S., Lagerwaard F. J., van den Berg R., Buis D. R. & Slotman B. J.  2007,  Neurosurgery, 60, 511. Gladman B., Kavelaars J.J., Nicholson P.D., Loredo T.J. & Burns J.A.  1998, AJ, 116,  2042 Jewitt D., Luu J. & Trujillo C. 1998. AJ, 115, 2125. Morrill J.S., Korendyke C.M., Brueckner G.E., Giovane F., Howard R.A., Koomen M.,  Moses D., Plunkett S.P., Vourlidas A., Esfandiari E., Rich N., Wang D., Thernisien A.F.,  Lamy P., Llebaria A., Biesecker D., Michels D., Gong Q. & Andrews M. 2006, Solar  Physics, 233, 331. Petit J.M., Holman M., Scholl H., Kavelaars J. & Gladman B. 2004, MNRAS, 347, 471. Sekanina Z. 2002, ApJ, 566, 577.   Veillet C, Parker JW, Griffin I, Marsden B, Doressoundiram A, Buie M, Tholen DJ,  Connelley M, Holman MJ. 2002, Nature 416, 711. Page 8 of 9 Figure 1. Images from the SOHO LASCO C3 coronagraph showing a comet plunging  into the sun. Panels (a) ­ (d) show the standard presentation for 4 of the 18 grayscale  images acquired. Panel (e) shows the arrival time CIP of the 18 component images. The  color bar in the upper right corner of the CIP gives the timing of the different hues.    Page 9 of 9 Figure 2. Images of  binary TNO 1998 WW31 taken by the HST presented as a CIP. The  images have been aligned so the brightest object (a) appears stationary. The second object  was imaged in July (b), August (c), September (d) and December (e) of 2001 and again in  January (f) and February (g) of 2002 to delineate its orbit. 
0802.1769
1
0802
2008-02-13T07:05:57
Transient horizontal magnetic fields in solar plage regions
[ "astro-ph" ]
We report the discovery of isolated, small-scale emerging magnetic fields in a plage region with the Solar Optical Telescope aboard Hinode. Spectro-polarimetric observations were carried out with a cadence of 34 seconds for the plage region located near disc center. The vector magnetic fields are inferred by Milne-Eddington inversion. The observations reveal widespread occurrence of transient, spatially isolated horizontal magnetic fields. The lateral extent of the horizontal magnetic fields is comparable to the size of photospheric granules. These horizontal magnetic fields seem to be tossed about by upflows and downflows of the granular convection. We also report an event that appears to be driven by the magnetic buoyancy instability. We refer to buoyancy-driven emergence as type1 and convection-driven emergence as type2. Although both events have magnetic field strengths of about 600 G, the filling factor of type1 is a factor of two larger than that of type2. Our finding suggests that the granular convection in the plage regions is characterized by a high rate of occurrence of granular-sized transient horizontal fields.
astro-ph
astro-ph
Astronomy & Astrophysics manuscript no. 9022 October 25, 2018 c(cid:13) ESO 2018 Letter to the Editor Transient horizontal magnetic fields in solar plage regions R. Ishikawa1,2, S. Tsuneta2, K. Ichimoto2, H. Isobe3, Y. Katsukawa2, B. W. Lites4, S. Nagata5, T. Shimizu6, R. A. Shine7, Y. Suematsu2, T. D. Tarbell7, and A. M. Title7 1 Department of Astronomy, University of Tokyo, Hongo, Bunkyo-ku, Tokyo 113-0033, Japan 2 National Astronomical Observatory of Japan, 2-21-1 Osawa, Mitaka, Tokyo 181-8588, Japan 3 Department of Earth and Planetary Science, University of Tokyo, Hongo, Bunkyo-ku, Tokyo, 113-0033, Japan 4 High Altitude Observatory, National Center for Atmospheric Research, P.O. Box 3000, Boulder CO 80307-3000, USA 5 Kwasan and Hida Observatories, Kyoto University, Yamashina, Kyoto, 607-8471, Japan 6 ISAS/JAXA, Sagamihara, Kanagawa, 229-8510, Japan 7 Lockheed Martin Solar and Astrophysics Laboratory, B/252, 3251 Hanover St., Palo Alto, CA 94304, USA Received/Accepted ABSTRACT Aims. We report the discovery of isolated, small-scale emerging magnetic fields in a plage region with the Solar Optical Telescope aboard Hinode. Methods. Spectro-polarimetric observations were carried out with a cadence of 34 seconds for the plage region located near disc center. The vector magnetic fields are inferred by Milne-Eddington inversion. Results. The observations reveal widespread occurrence of transient, spatially isolated horizontal magnetic fields. The lateral extent of the horizontal magnetic fields is comparable to the size of photospheric granules. These horizontal magnetic fields seem to be tossed about by upflows and downflows of the granular convection. We also report an event that appears to be driven by the magnetic buoyancy instability. We refer to buoyancy-driven emergence as type1 and convection-driven emergence as type2. Although both events have magnetic field strengths of about 600 G, the filling factor of type1 is a factor of two larger than that of type2. Conclusions. Our finding suggests that the granular convection in the plage regions is characterized by a high rate of occurrence of granular-sized transient horizontal fields. Key words. Sun: faculae, plages, Sun: granulation, Sun: magnetic fields, Sun: photosphere 1. Introduction Sunspots and active regions are believed to form from mag- netic flux globally emerging from the solar interior into the solar atmosphere. The emergence of magnetic flux is due to the buoyancy of large-scale, coherent toroidal flux tubes in the convection zone (Parker 1955). Emerging horizon- tal fields become more or less vertical to the photosphere, and the vertical fields provide natural connectivity with ob- served X-ray loops in the corona. This established paradigm is widely supported by observations and theory, but a con- cern has been raised on theoretical grounds that flux rising buoyantly through the convection zone will fail to maintain a coherent magnetic structure, and therefore fail to emerge (Magara 2001; Moreno-Insertis et al. 1995). Here we report on observations at granular scales of widespread emergence of isolated magnetic fields that may be related to the con- cern. Our observations were made with the Solar Optical Telescope (SOT, Tsuneta et al. 2007) aboard Hinode (Kosugi et al. 2007). This telescope performs high- resolution (∼0.′′3) polarimetric observations with a very sta- ble point spread function. Its Spectro-Polarimeter (SP) ob- tains the complete polarization state (Stokes I, Q, U , and V ) of the absorption profiles of the Zeeman-sensitive pho- tospheric lines FeI 630.15 nm and FeI 630.25 nm. We ob- served a plage region located near the center of the solar disk. A narrow area of 2.′′5 (8 slit positions, East-West) × 164′′(North-South) was rapidly scanned with a coarse pixel size of 0.′′32 by the SP, resulting in the moderate spatial res- olution of ∼0.′′6, and very high time cadence of 34 seconds over 40 minutes. 2. Ubiquitous transient fields in a plage region We obtain the continuum intensity, Ic, linear polarization, LP = R pQ2 + U 2dλ/ R Icdλ, and circular polarization, CP = R V dλ/ R Icdλ. The integration for LP is performed between -21.6 pm and +21.6 pm from the average center of the Stokes I profiles (FeI 630.25 nm) for pixels without magnetic signals, and the integration for CP is between +4.32 pm and +21.6 pm from the center of each Stokes I profile. In the weak field limit, LP is proportional to squares of the transverse field strength, and CP to the line-of-sight field strength. We discovered the frequent appearance and disappear- ance of nearly horizontal magnetic structures in the plage region: we find 52 occurrences of transient horizontal mag- netic fields in this short time series. These events appear to be randomly located spatially and temporally in the 2.′′5 × 164′′region observed for a duration of 40 min. These events are not associated with existing long-lived magnetic fields, and appear in a region of insignificant vertical fields. An 2 R. Ishikawa et al.: Transient horizontal magnetic fields in solar plage regions Fig. 1. The plage region located near disk center (72′′W, 76′′S) was observed from 16:00 to 16:40 UT on February 10, 2007. The Stokes profiles at two slit positions with integration time of 1.6 s each and scan step of 0.′′16 each are summed, and 2 pixels along the slit with a pixel size of 0.′′16 are also summed to obtain polarization accuracy better than 0.1%: The images consist of 8 slit positions (0.′′32 width) with a total scan time of 34 s. The pixel size along the slit is 0.′′32. The evolution of physical quantities for the plage region are shown: (a) CP (vertical magnetic field), (b) LP (horizontal magnetic field), (c) Ic. The region where LP is larger than 0.3% is enclosed by red lines. The emergence of the horizontal magnetic flux starts at ∆t = 0 s. Solar north is up and east to the left in all images of this report. episode of such transient horizontal magnetic fields is shown in Figure 1. The maximum size of the transient horizontal magnetic field is ∼1.′′4 ×1′′, and its duration is about 6 min. A remarkable feature is that the horizontal magnetic struc- ture first appears inside the granule, subsequently moves to the inter-granular lane, and finally disappears in the inter- granular lane. The horizontal magnetic structure continues to be smaller than or at most as large as the size of the gran- ule where it appears. Horizontal magnetic fields with nega- tive vertical magnetic fields at one end appear at ∆t=34 s (black patch in Fig.1), where there are no pre-existing ver- tical magnetic fields@at ∆t=0 s. Vertical magnetic fields with positive polarity start to be seen clearly at the op- posite end at ∆t=102 s. After ∆t=238 s, it is difficult to identify the bipolar configuration around the shrinking hor- izontal magnetic fields due to the weak vertical magnetic signals. The transient horizontal fields discovered in the plage regions have lifetimes ranging from a minute to about ten minutes, comparable to the lifetime of granules. Among 52 events, 43 horizontal magnetic structures appear inside the granules, and four appear in inter-granular lanes with the remaining five events ambiguous in position. Since 52 events are detected in the 2.′′5 × 164′′observing area during the 40 minutes, a new horizontal field appears every 46 seconds in the same observing region. The turnover time of the gran- ules is approximately 1000 sec with a velocity of 2 km s−1 and with a depth comparable to the horizontal scale. There are approximately 182 granules in the observing area, as- suming that the size of the granules is 1.′′5 × 1.′′5. 84% of these granules are not associated with stable strong ver- tical magnetic fields, and we use this smaller sample for estimating the frequency of events. Were every granule to have one embedded horizontal magnetic field structure, the horizontal field would have appeared at the surface every 6.6 seconds (∼1000 seconds/152 granules) in the observ- ing area. This shows that more than approximately 10% of the granules have embedded horizontal fields, suggesting a relatively common occurrence of horizontal magnetic fields underneath the plage region. We examined the Stokes profiles along the strong hor- izontal magnetic structure at ∆t=136 s (Figure 2). The Stokes Q and U profiles representing the horizontal mag- netic fields are blue-shifted by 0-1 km s−1 with respect to the line center of each Stokes I profiles. The strong hori- zontal magnetic field rises along with the upward convective motion of the granule in the early phase of the event; these magnetic fields would be subject to the convective flows. The polarity of the very weak vertical field component de- tected by the Stokes V profile reverses along the horizontal magnetic structure, indicating either convex- or concave- upward magnetic fields, depending on the resolution of the 180-degree ambiguity in the direction of the horizontal field. Because we follow the emergence of this event from its in- ception, the only logical configuration is convex-upward; i.e., an emerging omega-shaped loop. This scenario is con- sistent with the higher upward flow near the center of the granular cell. The transient horizontal fields reported here differ from the typical emerging flux driven by the Parker instabil- ity (Parker 1955; Shibata et al. 1989; Strous et al. 1996; Lites et al. 1998; Kubo et al. 2003, and references therein). This difference can be seen in an observation made with SOT which we describe next. 3. Comparison with buoyancy-driven emerging flux We also present an example of an emerging flux event in a remnant active region located around the disk cen- ter. The properties of the event are considerably different from the previous example. We find only one such event in the observing area of 3.′′8 × 82′′ during 50 minutes' ob- servation. The episode starts with the appearance of hori- zontal magnetic fields, that appear to cause a considerable R. Ishikawa et al.: Transient horizontal magnetic fields in solar plage regions 3 deformation of the surrounding granules (Figure 3). The maximum size of the horizontal structure is ∼1.′′6 along its major axis. At ∆t=-128 s, prior to the emergence of magnetic flux, the granules become elongated with their alignment parallel to the orientation of the emerging fields. At ∆t=0 s, strong horizontal magnetic field structure ap- pears in the region between the elongated granules and existing vertical fields, and vertical magnetic components are detected at both ends of the horizontal magnetic struc- ture. Note that we employ unsigned circular polarization, nCP = R V dλ/ R Icdλ, instead of CP because highly de- formed Stokes V profiles, some of which have three lobes, are seen in the region. The integration is carried out with the same wavelength range as LP . The horizontal flux emerges with separating vertical components at both ends. At ∆t=128 s, the horizontal magnetic structure becomes longer and stronger, and the vertical magnetic fields also strengthen. The bright point appears where the vertical field is strong. In the next frame (∆t= 256 s), the horizontal magnetic structure disappears (most likely rising above the line formation height), and we are left with vertical bipolar components. The series of observations indicate that the emerging magnetic field structure is stiff against the gran- ular motion since it affects the surrounding granules even before its emergence into the photosphere. The Stokes profiles along the horizontal magnetic struc- ture at ∆t=128 s are shown in Figure 4. The Stokes Q and U profiles show that the direction of the magnetic vec- tor is roughly parallel to the major axis of the flux tube. The Stokes U profiles representing the horizontal magnetic structure are blue-shifted by 1-2 km s−1 at the top of the loop with respect to the line center of the Stokes I profiles, while the Stokes V profile is strongly red-shifted at one end of the horizontal magnetic structure with a velocity up to 5 km s−1, which is below but close to the local sound speed (6-7 km s−1). The strong downward flow may be related to the convective collapse (Bellot Rubio et al. 2001). The expansion velocity is estimated to be about 2 km s−1 from the change of the length of the horizontal magnetic struc- ture. This velocity is comparable to the upward velocity at the top of the loop. The high-velocity downflows at one end of the flux tube continuing over 4 minutes show that plasma descends in the flux tube in association with the ris- ing flux tube. At the other footpoint, a few km s−1 upflow is detected mainly at ∆t=128 s. This may be a reverse flow due to the local enhancement of gas pressure triggered by a downflow near the footpoint. Another possibility is siphon flow driven by the pressure gradient of the two foot points (Cargill & Priest 1980). Note that even with the upflow at one end of the footpoints, the higher downflow dominates, and reduces the net mass in the flux tube. The magnetic buoyancy will be sustained as a result. This property is closer to that of emerging flux by the Parker instability than the previous one. Given the significant difference in the two episodes pre- sented here, we hereafter call this episode type1, and the previous episode type2. We do not find any events of type1 in the data set where we find many type2 events. The oc- currence rate of type1 is much lower than that of type2. The three components of the magnetic field are estimated by a Milne-Eddington inversion for the Stokes spectra shown in the Figures 2 and 4. (The black lines indicate the observed Stokes spectra, while the red lines the fitted spectra.) The average horizontal magnetic field strengths are 560 Gauss (type1) and 580 Gauss (type2). The field lines are roughly aligned along the major axis of the strong horizontal mag- netic field region. For type2, we have chosen the event hav- ing the strongest LP , so that we were able to perform a reliable Stokes inversion. Other type2 events generally have smaller LP . A remarkable difference between the types 1 and 2 is in the average filling factors, which are 0.44 and 0.17, respec- tively. The filling factor is an areal fraction of the magnetic atmosphere within a resolution element. We estimate the maximum diameter of the flux tube under the assumption of a single horizontal flux tube partially occupying a pixel. The width is equal to the size of the telescope's point spread function (0.′′32) multiplied by the filling factor. The size of the type1 horizontal flux tube thus derived is less than ∼100 km, while that for the type2 is less than ∼40 km. They are much smaller than the scale height of the line forming layer, indicating that actual diameter of the flux tube is not as small as this. Nevertheless, it is true that the type2 tube is much thinner than the type1 tube. 4. Discussion The occurrence of isolated small-scale, transient horizon- tal magnetic fields in the quiet Sun was discovered us- ing spectro-polarimetric data from the Advanced Stokes Polarimeter at a spatial resolution of 1′′(Lites et al. 1996). These horizontal fields are interpreted to be convection- driven or buoyancy-driven emerging bipolar fields. Very re- cently, the high time variability and ubiquity of horizontal fields in the quiet Sun were clearly shown in data from the SOLIS and GONG instruments (Harvey et al. 2007). Subsequently, Hinode data has permitted us to directly see the ubiquitous presence of horizontal magnetic fields in the quiet Sun at unprecedented small scales (Lites et al. 2008; Centeno et al. 2007; Orozco Su´arez et al. 2007). It is as yet unclear if the granular-size emergence phenomenon in the plage region reported herein pertains to flux elements hav- ing a stronger degree of polarization than much of that reported for the quiet Sun. It is important to understand whether the plage and quiet Sun horizontal fields have a similar origin. e 1 The equi-partition field strength Be is the field strength B 2 where magnetic energy is equal to kinetic energy: 8π = 2 ρυ2. The typical equipartition field strength Be is about 400 Gauss for granules with a velocity of υ = 2×105 cm s−1, and the density ρ is 3×10−7 g cm−2 at the solar surface. The magnetic field strengths of both type1 and type2 are appar- ently stronger than the equi-partition field. The magnetic field strength of type1 higher than the equi-partition field strength is consistent with the observed robustness against granular motion (Cheung et al. 2007). The type1 event is interpreted as essentially the instability-driven emergence of flux tubes. As strong a field for type2 cannot be readily explained. It appears that the horizontal flux tubes of the type 2 event are carried by the upflow of convection to the photosphere, and then the horizontal flow of the convection transports them to the inter-granular lane. The type2 event is apparently subject to granular motion. The buoyant force is proportional to (ρe − ρi)πga2, where ρe and ρi are the plasma density inside and out- side the horizontal flux tubes, and a the radius of the flux tube. The drag force is given by 0.5Cdρeu2a, where Cd is an 4 R. Ishikawa et al.: Transient horizontal magnetic fields in solar plage regions Fig. 2. The frame at ∆t = 136 sec in Fig.1 is enlarged on the left. Stokes profiles (I, Q, U , and V ) along the major axis (blue line) are shown to the right with black lines. The fitted profiles are shown with red lines. The profiles marked by A and B correspond to the NE and SW ends of the horizontal field structure, respectively. The vertical line in the Stokes profiles shows the average center of the Stokes I absorption profiles (FeI 630.25 nm) without magnetic fields, and provides a fiducial for the zero point in velocity. The tick mark indicates 2.16 pm. aerodynamic coefficient of order unity, and u the turbulent velocity. The magnetic tension force is πa2B 2/4πL, where L is the curvature radius of horizontal fields. We substitute a granular size of 103 km for L to obtain the upper limit of magnetic tension. The three forces become comparable for a of 50 km and B of 600 G. The type2 tube diameter of 40 km means that the effect of buoyancy is relatively smaller than the drag force. This implies that a thinner flux tube is more strongly influenced by turbulent motion. Convective collapse (Parker 1978) would not work due to its horizontal nature throughout the evolution for the type 2 event. Indeed, Abbett (2007) shows that a convective surface dynamo can generate horizontally directed mag- netic structures. The transient horizontal fields revealed by Hinode in our study suggest the possibility of such a local dynamo process (Cattaneo 1999; Vogler & Schussler 2007). An alternative physical picture is presented by the extended fields caused by an exploding magnetic structure (Moreno-Insertis et al. 1995) and/or by horizontal mag- netic fields that failed to emerge (Magara 2001). Acknowledgements. The authors thank B. C. Low for his valuable comments and encouragement. Hinode is a Japanese mission devel- oped and launched by ISAS/JAXA, with NAOJ as a domestic part- ner and NASA and STFC (UK) as international partners. It is oper- ated by these agencies in co-operation with ESA and NSC (Norway). The National Center for Atmospheric Research is sponsored by the National Science Foundation. References Abbett, W. P. 2007, ApJ, 665, 1469 Bellot Rubio, L. R., Rodr´ıguez Hidalgo, I., Collados, M., Khomenko, E., & Ruiz Cobo, B. 2001, ApJ, 560, 1010 Cargill, P. J. & Priest, E. R. 1980, Sol. Phys., 65, 251 Cattaneo, F. 1999, ApJ, 515, L39 Centeno, R., Socas-Navarro, H., Lites, B., et al. 2007, ApJ, 666, L137 Cheung, M. C. M., Schussler, M., & Moreno-Insertis, F. 2007, A&A, 467, 703 Fig. 3. NOAA AR 0931 located not far from disk center (289′′E, 83′′S) was observed from 17:35 to 18:25 UT on December 28, 2006. The images consist of 24 consecutive slit positions (0.′′16 width) with a total scan time of 128 s. The pixel size along the slit is 0.′′16. (a) nCP (vertical magnetic field), (b) LP (horizontal magnetic field), (c) Ic. The region where LP is larger than 0.5% is enclosed by red lines. Fig. 4. The frame at ∆t = 128 sec in Fig.3 enlarged on the left. The caption is the same as Fig.2. Harvey, J. W., Branston, D., Henney, C. J., & Keller, C. U. 2007, ApJ, 659, L177 Kosugi, T., Matsuzaki, K., Sakao, T., et al. 2007, Sol. Phys., 243, 3 Kubo, M., Shimizu, T., & Lites, B. W. 2003, ApJ, 595, 465 Lites, B. W., Kubo, M., Socas-Navarro, H., et al. 2008, ApJ, 672, 1237 Lites, B. W., Leka, K. D., Skumanich, A., Martinez Pillet, V., & Shimizu, T. 1996, ApJ, 460, 1019 Lites, B. W., Skumanich, A., & Martinez Pillet, V. 1998, A&A, 333, 1053 Magara, T. 2001, ApJ, 549, 608 Moreno-Insertis, F., Caligari, P., & Schuessler, M. 1995, ApJ, 452, 894 Orozco Su´arez, D., Bellot Rubio, L. R., del Toro Iniesta, J. C., et al. 2007, ApJ, 670, L61 Parker, E. N. 1955, ApJ, 121, 491 Parker, E. N. 1978, ApJ, 221, 368 Shibata, K., Tajima, T., Steinolfson, R. S., & Matsumoto, R. 1989, ApJ, 345, 584 Strous, L. H., Scharmer, G., Tarbell, T. D., Title, A. M., & Zwaan, C. 1996, A&A, 306, 947 Tsuneta, S., Suematsu, Y., Ichimoto, K., et al. 2007, Sol. Phys., sub- mitted Vogler, A. & Schussler, M. 2007, A&A, 465, L43
astro-ph/0609800
1
0609
2006-09-28T22:09:54
The Formaldehyde Masers in Sgr B2: Very Long Baseline Array and Very Large Array Observations
[ "astro-ph" ]
Observations of two of the formaldehyde (H2CO) masers (A and D) in Sgr B2 using the VLBA+Y27 (resolution ~0.01") and the VLA (resolution ~9") are presented. The VLBA observations show compact sources (<10 milliarcseconds, <80 AU) with brightness temperatures >10^8 K. The maser sources are partially resolved in the VLBA observations. The flux densities in the VLBA observations are about 1/2 those of the VLA; and, the linewidths are about 2/3 of the VLA values. The applicability of a core-halo model for the emission distribution is demonstrated. Comparison with earlier H2CO absorption observations and with ammonia (NH3) observations suggests that H2CO masers form in shocked gas. Comparison of the integrated flux densities in current VLA observations with those in previous observations indicates that (1) most of the masers have varied in the past 20 years, and (2) intensity variations are typically less than a factor of two compared to the 20-year mean. No significant linear or circular polarization is detected with either instrument.
astro-ph
astro-ph
The Formaldehyde Masers in Sgr B2: Very Long Baseline Array 23 Sept 2006 version and Very Large Array Observations Ian M. Hoffman St. Paul's School, 325 Pleasant Street, Concord, NH 03301 [email protected] W. M. Goss National Radio Astronomy Observatory, Socorro, NM 87801 and Patrick Palmer Department of Astronomy and Astrophysics, University of Chicago, 5640 S. Ellis Ave., Chicago, IL 60637 ABSTRACT Observations of two of the formaldehyde (H2CO) masers (A and D) in Sgr B2 using the VLBA+Y27 (resolution ≈ 0.′′01) and the VLA (resolution ≈ 9′′) are presented. The VLBA observations show compact sources (. 10 milliarcseconds, . 80 AU) with brightness temperatures > 108 K. The maser sources are partially resolved in the VLBA observations. The flux densities in the VLBA observations are about 1/2 those of the VLA; and, the linewidths are about 2/3 of the VLA values. The applicability of a core-halo model for the emission distribution is demonstrated. Comparison with earlier H2CO absorption observations and with ammonia (NH3) observations suggests that H2CO masers form in shocked gas. Comparison of the integrated flux densities in current VLA observations with those in previous observations indicates that (1) most of the masers have varied in the past 20 years, and (2) intensity variations are typically less than a factor of two compared to the 20-year mean. No significant linear or circular polarization is detected with either instrument. Subject headings: ISM: individual(Sgr B2) -- ISM: molecules -- masers -- radio lines: ISM -- 2 -- 1. Introduction The nature of Galactic formaldehyde (H2CO) masers is a growing mystery. While hundreds of Galactic OH, H2O, and CH3OH masers are known, only five Galactic star- forming regions have associated H2CO maser emission. To date, this emission is seen only in the 110 → 111 transition at 6 cm wavelength. Shortly after the discovery of the first H2CO maser in NGC 7538 (Downes & Wilson 1974; Forster et al. 1980), a radiative pumping model was proposed (Boland and de Jong 1981). The H2CO masers discovered subsequently did not meet the conditions required for this mechanism (Gardner et al. 1986; Mehringer, Goss, & Palmer 1994, hereafter MGP94; Hoffman et al. 2003; hereafter H03). Thus, twenty-five years after the discovery of the first H2CO maser, these sources remain rare and the excitation mechanism remains unknown. Sgr B2, the northernmost component of the extended Sgr B radio source, is located within a few hundred pc of the Galactic center (Reid et al. 1988). (The distance to the Galactic center is assumed to be 8.5 kpc in this paper.) Sgr B2 is comprised of three main star-forming complexes designated north (N), middle or main (M), and south (S), and many smaller Hii regions. The H2CO masers occur throughout Sgr B2, shown in Figure 1. The heating mechanisms and complex chemistry of the region are subjects of ongoing study (e.g., Gaume & Claussen 1990; Goicoechea et al. 2004). Sgr B2 contains nine individual H2CO maser regions, several of which have multiple velocity components. All of the masers are unresolved at 1′′ angular resolution, except for maser C which MGP94 suggest consists of several masers blended within the beam. These regions are near Hii regions distributed over the ∼3.6 arcmin2 complex (MGP94). Whiteoak and Gardner (1983) and MGP94 designated the maser regions with letters (Fig. 1). The H2CO masers are observed over the velocity range +40 km s−1. vLSR . +80 km s−1, while other species, such as H2O masers, are observed over a larger range −30 km s−1. vLSR . +120 km s−1 (Kobayashi et al. 1989; McGrath, De Pree, and Goss 2004). Of the nine maser regions observed in Sgr B2 by MGP94, the G maser was shown to be time variable, at least quadrupling in intensity over 10 yr. (Similarly, H03 found one NGC 7538 feature to triple in intensity over ≈ 10 yr.) As initially noted by Whiteoak and Gardner (1983), all of the Sgr B2 H2CO masers lie close to OH, H2O, CH3OH, and NH3 masers. For most of the masers in MGP94, the separation to an OH maser was less than 0.05 pc. Recent successes in search techniques for new masers (Araya et al. 2004, 2005, 2006a) and in high-resolution observational techniques (H03) promise to provide empirical constraints for the development of a realistic model for the Galactic H2CO maser emission. The necessary steps in compiling an empirical picture of the H2CO emission in Sgr B2 are (1) detailed imaging of the masers in order to quantify the intrinsic properties of the emission (e.g., -- 3 -- brightness temperature), (2) assessment of intensity variability in the masers, and (3) precise astrometry for elucidating spatial relationships between the H2CO masers and more common masers (OH, H2O, CH3OH). In this paper, we present new observations of the H2CO masers in Sgr B2 using the the Very Long Baseline Array (VLBA) and Very Large Array (VLA) of the NRAO1. 2. Observations and Results 2.1. VLBA+Y27 We observed the A and D H2CO masers in Sgr B2 using the ten antennas of the VLBA and the 27 antennas of the VLA as an 11-station VLBI array. Parameters of the observations are summarized in Table 1. The total observing time was approximately 8.0 hours alternating between the A and D pointing positions which are separated by approximately 1′. At each pointing position there is a useful correlated field of view of approximately 1′′ (e.g., Bridle & Schwab 1999). To observe each of the nine H2CO masers in Sgr B2 optimally would have required observations at nine pointing centers. Because of signal-to-noise considerations, we observed only the two masers measured to be most intense by MGP94. The baseline lengths of the VLBA+Y27 array range from 52 km to 8611 km; the array is not sensitive to angular scales larger than 0.′′25. The antennas have right- (R) and left- (L) circularly polarized feeds from which RR, LL, RL, & LR cross-correlations were formed. The visibilities were integrated for 8.4 s. The amplitude scale is set using online system temperature monitoring and a priori antenna gain measurements. The station delays were determined from observations of J1733- 130. The maser observations were phase referenced to J1745-283 (called W56 in Bower et al. 2001). Because the properties of this source limit our observations, we discuss it in some detail. The absolute position uncertainty of this source is 12 milliarcseconds (hereafter: mas) (Reid et al. 1999). Bower et al. find that J1745-283 is probably the core of an extragalactic jet source and that its apparent size at 5 GHz (30 mas) is determined by scatter-broadening. We also observed J1745-283 on 22 November 2002 using only the ten VLBA stations (a snapshot observation with a bandwidth of 4 MHz in both polarizations). The 2002 observation yields a deconvolved angular size of 21 × 18 mas and an integrated flux density of 28 mJy; the 2003 observation yields a size of 28 × 16 mas and a flux density of 31 mJy. (The flux density 1The VLA and VLBA are components of the National Radio Astronomy Observatory (NRAO), a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc. -- 4 -- determined by Y27 [the VLA alone] during the 2003 observations was 161 mJy.) Bower et al. report an angular size 42 × 25 mas and peak intensity 89 mJy beam−1 from VLBA+Y1 observations. Therefore, the flux density must have decreased significantly between 1999 and 2002. Such variability would not be unusual for an inverted spectrum source like J1745-283 (e.g., Urry & Padovani 1995). Our imaging of J1745-283 with the VLBA makes use of only the inner-most stations (VLA, Pie Town, Los Alamos, Fort Davis, Kitt Peak, and Owens Valley) because the source is resolved on longer baselines. Some additional resolution for the maser observations was gained by self-calibration; but, because only the shorter baselines in the VLBA data were absolutely calibrated in phase, the position registration accuracy of the resulting VLBA images is approximately 15 mas. In summary, we could not make use of the full potential resolution of the VLBA when using this phase referencing calibrator, and no other suitable nearby source is known. We detected masers toward the A and D regions. The F maser is also within the correlated field of view near the A maser, but was not detected with the current sensitivity. The image and spectrum of the A maser from the VLBI data is shown in Figure 2. Because the radio continuum of Sgr B2 is fully resolved by the VLBA, no continuum subtraction was necessary. The positions, center velocities (vLSR), linewidths (∆vFWHM), peak flux densities (S0), and deconvolved major and minor axes and position angles are summarized in Table 2. In parentheses following each entry are the 1-σ errors. The deconvolved sizes of the VLBI images of the A and D masers are ≈ 10 mas. At the distance of Sgr B2, this size corresponds to a linear diameter of approximately 80 AU, comparable to the sizes (30 to 130 AU) observed for the H2CO masers in NGC 7538 and G29.96-0.02 using the VLBA (H03). The brightness temperatures of the Sgr B2 A and D masers in the current VLBI data are 5.9 × 108 K and 1.2 × 108 K, respectively, similar to the 107−8 K observed for other H2CO masers (H03). No significant linear or circular polarization is detected in either maser (≤ 20% for the strongest maser [A]).2 2All data reduction and analysis was performed with the the The Groningen Image Processing System (GIPSY) software package (http://www.ast ro.rug.nl/∼gipsy/) and the Astronomical Image Processing System (AIPS) software package (http://www.nrao.edu/aips/). -- 5 -- 2.2. VLA Because of the known variability of H2CO masers (Forster et al. 1985; H03), the Sgr B2 masers were observed in 2005 with the VLA in order to access any possible time-variability. Parameters of the observations are summarized in Table 1. The array configuration available was DnC for an observation period of approximately 7 hours. Two pairs of RR and LL bands were recorded centered at the expected velocities of the A and the D masers. Of the nine maser regions described by MGP94, eight of the sources were detected. Six velocity components in the regions observed by MGP94 lie outside the velocity range of the current observations. No new H2CO maser regions were discovered. The 2005 data have inferior angular resolution (10′′ versus 1′′) but improved spectral resolution (0.19 km s−1 versus 1.5 km s−1) compared with the 1993 VLA observations of MGP94. Comparison of the current data to the MGP94 data is uncertain due to (1) the severe blending of the strong H2CO absorption with the nearby masers and (2) the insufficient velocity resolution of the MGP94 data, which did not spectrally resolve the lines. Variability of the flux density of the masers is apparent and is discussed in §3.2. The 2005 results are summarized in Table 3 and a spectrum from these data (region E) is shown in Figure 3. The C maser region is not discussed in this paper due to confusion of the maser spectra with nearby absorption and continuum emission. With the current 10′′ angular resolution of the VLA data, as with the 1′′ resolution of MGP94, none of the masers are spatially resolved. In §3.2, we also compare the current data with the 1983 VLA observations of Gardner et al. (1986). The velocity resolution of the 1983 data is 0.76 km s−1 and the angular resolution is 1.′′25. Although we resolve many of the maser line profiles in velocity for the first time with the 2005 VLA data, most of the linewidths presented in Table 3 agree with the values measured by Gardner et al. (1986). No circular polarization is detected in any of the H2CO masers. This corresponds to a 3-σ upper limit of approximately 4% circular polarization for the strongest maser (A). 3. Discussion 3.1. Angular Scatter Broadening Images of radio sources are angularly broadened by scattering by the ionized component of the interstellar medium. In the direction of the Galactic center, this problem becomes severe (e.g., Lazio & Cordes 1998). As discussed in §2.1, J1745-283, is an extragalactic -- 6 -- point-like source whose image is broadened to ∼30 mas at 6 cm wavelength by scattering (Bower et al. 2001). In this section we discuss the extent to which this scattering medium affects the current VLBI observations of the H2CO masers in Sgr B2. The observed size for J1745-283, 28 × 16 mas, is larger than the observed deconvolved size for the A maser, 13 × 5 mas. This result may be expected even if both sources are intrinsically point-like because the proximity of the maser to the scattering medium results in reduced angular broadening (e.g., Rickett 1990). Nevertheless, as discussed below, we expect significant angular broadening in the images of the masers. Therefore, the deconvolved angular sizes in Table 2 are upper limits to the intrinsic sizes and the brightness temperatures are lower limits. Gwinn et al. (1988) quantified the scattered sizes of H2O and OH masers in Sgr B2, finding a λ2.2 dependence for the broadening. From their determinations of minimum angular sizes of 0.3 mas for 1.35-cm H2O masers and 100 mas for 18-cm OH masers, we expect a min- imum apparent size of approximately 9 mas for the 6.2-cm H2CO masers. The deconvolved angular sizes of the masers in Table 2 are in agreement with this expectation. 3.2. Variability of the H2CO masers in Sgr B2 In comparing the current VLA observations with earlier VLA data, both the differing angular resolution and spectral resolution must be considered. As discussed in §2.2, both the differences in array configuration and in correlator setup were significant. To compensate for the difference in velocity resolution, in Table 4 we tabulate the velocity-integrated flux density for observations made in 1983 (Gardner et al. 1986), 1993 (MGP94), and 2003 and 2005 (this paper). However, it is impossible to compensate for emission outside of the velocity range covered in the current observations and for the confusion of features caused by the lower angular resolution. (The notes in the final column of Table 3 summarize limitations of the current data from these causes.) The error values in Table 4 are dominated by systematic uncertainties rather than by thermal noise in most cases. We find that the A maser has approximately doubled in velocity-integrated flux density since 1993, while B, G, and H increased between 1983 and 1993, and subsequently decreased. The velocity-integrated flux densities from regions D, E, F, and I have not changed signifi- cantly since 1993. -- 7 -- 3.3. Core/Halo Morphology For the A and D H2CO masers in Sgr B2, a comparison of the current VLBA+Y27 data with past and current VLA data shows two major differences: (1) only a fraction of the flux density seen with the VLA is observed in the VLBI data (∼70% for the A maser; ∼40% for the D maser); and (2) the velocity widths in the VLBI data are more narrow than in the VLA data (∼80% for the A maser; ∼50% for the D maser). In this section we discuss a schematic model for the structure of the emission which addresses both the observed angular distribution and velocity widths. The flux density detected by the VLA but resolved by VLBI baselines must lie at angular scales between ≈ 100 mas (the resolution of the shortest VLBA+Y27 interferometer spacings) and ≈ 500 mas (the masers are unresolved in the MGP94 observations with 1′′ resolution), while the flux density observed with the VLBI is emitted by a source ∼10 mas. These limits suggest a core-halo source morphology. A similar morphology was proposed by H03 for the H2CO masers in NGC 7538 and G29.96-0.02. A model consisting of two coin- cident circular Gaussian components with different angular sizes can reproduce the observed results. For the A maser, the flux density observed with the VLA is about 2000 mJy. To determine the decrease in flux density with increasing projected baseline length, we use the highest signal-to-noise baselines only: those between Y27 and other antennas. The average flux densities observed on baselines with Y27 are: 1400 mJy at 0.71 Mλ (to Pie Town, av- erage projected spacing ∼44 km)3; 1300 mJy at 3 Mλ; and 1200 mJy at 5 Mλ. These data are fit by a two component Gaussian model with a 600 mJy in a 300 mas halo and 1400 mJy in a 10 mas core. The brightness temperature of the halo is TB ≈ 4 × 105 K; for the 10 mas core component, TB ≈ 8 × 108 K (Table 2, Fig. 2). A more precise model is justified only after higher signal to noise measurements. It is reasonable to assume that the background radiation is relatively uniform on 300 mas angular scales because the continuum emission is well resolved on the VLBA baselines. Therefore, the difference in brightness temperatures between the core and halo components may be attributed to differences in maser gain. We estimate that the halo has a gain of approximately 10 (in the exponential amplification regime) and the core has a gain of approximately 2000 (in the saturated regime). Because the observed angular sizes are upper limits for the intrinsic sizes, the derived brightness temperatures are lower limits for the intrinsic brightness temperatures; therefore, the actual gains may exceed the above values. The gain of the maser medium is dependent upon four factors: the pathlength, the density 3Mλ denotes million wavelengths -- 8 -- of H2CO, the level inversion, and the velocity coherence (see §4.2 of H03). Therefore, one or more of these factors must be significantly different between the core and halo components. We suggest that the narrower line widths arising from the core component indicate that the velocity coherence is enhanced in the core region compared to the halo, yielding a higher gain. Similar arguments may be applied to the D maser, but the resulting ranges of size and TB are not well constrained because of the higher noise level on the VLA-Pie Town baseline. Additional maser species in other star-forming regions -- excited OH (Palmer, Goss, & Devine 2003), OH in supernova remnants (Hoffman et al. 2005), as well as other H2CO masers (H03) -- exhibit a similar narrowing of line widths between VLA and VLBA angular scales. 3.4. Associations with Other Molecular Lines H2CO absorption toward Sgr B2 has been studied extensively with 10′′ × 20′′ resolution both by Mart´ın-Pintado et al. (1990) and Mehringer et al. (1995). Mart´ın-Pintado et al. (1990) noted that the absorption is dominated by three velocity components (vLSR= 55, 64, and 80 km s−1), each with τ > 1. Linewidths of these absorption features ranged from 9 -- 26 km s−1, i.e. much greater than that of the maser features (< 1 km s−1; see Table 3). Mehringer et al. (1995) provide optical depth profiles toward selected positions. All but maser E lies within 30′′ of one of these positions displayed. (Maser E lies at a position with no radio continuum.) It is striking that the maser velocities, except for maser I (which is more than a beamwidth away from the position of a displayed profile), do not occur in velocity ranges with large H2CO optical depths. Masers A and F occur in a velocity range with τ =1, but this velocity range is a rather sharp minimum in the optical depth profile at this position. Therefore, we conclude that the gas containing the H2CO masers is distinct from the bulk of the H2CO containing gas observed in absorption. Insight into why the maser containing gas may be distinct is provided by (3,3) and (4,4) NH3 observations made with the VLA by Mart´ın-Pintado et al. (1999). With their 3′′ resolution, 80-90% of the NH3 emission is resolved out, and only small angular scale features remain. Among these features are a number of rings and arcs, most naturally interpreted as complete or partial shells with linear sizes ∼2 pc. The shells are hot (Tk ∼ 50 -- 70 K), and the H2 densities derived for them are typically a factor of 10 greater than those derived from the H2CO absorption studies of Mart´ın-Pintado et al. (1990). H2CO masers C, D, E, G, and H fall within the field imaged by Mart´ın-Pintado et al. (1999). As these authors note, all of the H2CO masers occur at positions on hot NH3 shells. The higher temperature (Tk > 100 K) region of apparent interaction between shells A and B contains the closely spaced D, G, -- 9 -- and H masers. Mart´ın-Pintado et al. (1999) propose that the pumping mechanism for the H2CO masers as well as that for the NH3 and class II CH3OH masers must depend on the physical conditions in the hot shells. In the search for additional constraints on the H2CO emission environment, we examine the possible association of H2CO masers with other molecular emission for which physical conditions are better understood. In Table 5 we present a summary of possible associations with H2O masers observed by McGrath, Goss, and De Pree (2004), CH3OH masers observed by Houghton and Whiteoak (1995), and OH masers observed by Argon, Reid, and Menten (2000). The excitation conditions for the different species are mutually exclusive, but if the masers exist in a shocked region as proposed by Mart´ın-Pintado et al. (1999) for Sgr B2 (and for NGC 7538 by H03), a rapid change in densities, temperatures, and velocity fields over a small linear distance is to be expected. Of the NH3 masers in Sgr B2 with interferometically determined positions, only masers M1 and M6 lie similarly near an H2CO maser (E), and both differ in velocity by >10 km s−1 from the H2CO maser. The existence of H2O and CH3OH masers near the H2CO F maser is significant because previously the F maser had no known maser or continuum associations (MGP94). The H2CO maser in NGC 7538 (H03) and many of those reported in Section 3.2 varied in intensity with timescales from years to decades. H03 noted a common variability of some of the H2CO and H2O masers in NGC 7538. This was further confirmed by long-term observations by Lekht et al. (2003, 2004b). Similarly in Sgr B2, the long-term monitoring of the H2O masers presented by Lekht et al. (2004a, 2004c) may indicate common variability of H2O masers with the H2CO A maser. However, variablity of an H2CO maser in IRAS 18566+0408 was recently seen to occur on much more rapid timescale (Araya et al. 2006b). 4. Conclusions We present VLBA+Y27 images of the Sgr B2 A and D H2CO masers. The measured sizes (. 80 AU) and brightness temperatures (> 108 K) are comparable to those found in other VLBI studies of H2CO masers. However, about half of the flux density from these regions is resolved out with the VLBA data. A comparison between VLA and VLBA obser- vations shows that the missing flux density exhibits a broader linewidth than the emission from the compact VLBI source. We demonstrate quantitatively the applicability of a core- halo model for these masers. We also present new VLA observations of the H2CO masers in Sgr B2 with improved velocity resolution. We have detected variability in several of the masers. -- 10 -- H2O and CH3OH masers discovered near the H2CO masers may indicate associations among the species, suggestive of a related origin. The association of H2CO masers in Sgr B2 with hot NH3 shells proposed by Mart´ın-Pintado et al. (1999), together with the arguments for shock excitation of the maser region in NGC 7538 in H03, provide an encouraging stepping stone toward a solution of the problem of the excitation of these poorly understood masers. We are indebted to an anonymous referee for a very careful reading of this paper and for directing our attention to two very important references. Facilities: VLA (), VLBA (). REFERENCES Araya, E., Hofner, P., Linz, H., Sewilo, M., Watson, C., Churchwell, E., Olmi, L., & Kurtz, S. 2004, ApJS, 154, 579 Araya, E., Hofner, P., Kurtz, S., Linz, H., Olmi, L., Sewilo, M., Watson, C., Churchwell, E. 2005, ApJ, 618, 339 Araya, E., Hofner, P., Goss, W. M., Kurtz, S., Linz, H., & Olmi, L. 2006a, ApJ, 643, L33 Araya, E., Hofner, P., Sewilo, M., Linz, H., Kurtz, S., Olmi, L., Watson, C., & Churchwell, E. 2006b, ApJ (in press) Argon, A. L., Reid, M. J., & Menten, K. M. 2000, ApJS, 129, 159 Boland, W. & de Jong, T. 1981, A&A, 98, 149 Bower, G. C., Backer, D. C., Sramek, R. A. 2001, ApJ, 558, 127 Bridle, A. H. & Schwab, F. R. 1999, in Synthesis Imaging in Radio Astronomy II. Edited by G. B. Taylor, C. L. Carilli, and R. A. Perley. ASP Conference Series, 180, 371 Downes, D. & Wilson, T. L. 1974, ApJ, 191, L77 Forster, J. R., Goss, W. M., Wilson, T. L., Downes, D., & Dickel, H. R. 1980, A&A, 84, L1 Forster, J. R., Goss, W. M., Gardner, F. F., & Stewart, R. T. 1985, MNRAS, 216, 35P Gardner, F. F., Whiteoak, J. B., Forster, J. R., & Pankonin, V., 1986, MNRAS, 218, 385 Gaume, R. A. & Claussen, M. J. 1990, ApJ, 351, 538 -- 11 -- Goicoechea, J. R., Rodr´ıguez-Fern´andez, N. J., Cernicharo, J. 2004, ApJ, 600, 214 Gwinn, C. R., Moran, J. M., Reid, M. J., Schneps, M. H. 1988, ApJ, 330, 817 Hoffman, I. M., Goss, W. M., Palmer, P., & Richards, A. M. S. 2003, ApJ, 598, 1061 (H03) Hoffman, I. M., Goss, W. M., Brogan, C. L., Claussen, M. J. 2005, ApJ, 620, 257 Houghton, S. & Whiteoak, J. B. 1995, MNRAS, 273, 1033 Kobayashi, H., Ishiguro, M., Chikada, Y., Ukita, N., Morita, K.-I., Okumura, S. K., Kasuga, T., & Kawabe, R. 1989, PASJ, 41, 141 Lazio, T. J. W. & Cordes, J. M. 1998, ApJ, 505, 715 Lekht, E. E., Munitsyn, V. A., & Tolmachev, A. M. 2003, Astronomy Reports, 47, 838 Lekht, E. E., Ram´ırez, O., Tolmachev, A. M., & Berulis, I. I. 2004, Astronomy Reports, 48, 171 Lekht, E. E., Munitsyn, V. A., & Tolmachev, A. M. 2004, Astronomy Reports, 48, 200 Lekht, E. E., Ram´ırez, O., & Tolmachev, A. M. 2004, Astronomy Reports, 48, 965 Mart´ın-Pintado, J., de Vicente, P., Wilson, T. L., & Johnston, K. J. 1990, A&A, 236, 193 Mart´ın-Pintado, J., Gaume, R. A., Rodr´ıguez-Fern´andez, N., de Vicente, P., & Wilson, T. L. 1999, ApJ, 519, 667 McGrath, E. J., Goss, W. M., & De Pree, C. G. 2004, ApJS, 155, 577 Mehringer, D. M., Goss, W. M., & Palmer, P. 1994, ApJ, 434, 237 (MGP94) Mehringer, D. M., Palmer, P., & Goss, W. M. 1995, ApJS, 97, 497 Palmer, P., Goss, W. M., & Devine, K. E. 2003, ApJ, 599, 324 Reid, M. J., Schnepps, M. H., Moran, J. M., Gwinn, C. R., Genzel, R., Downes, D., & Ronnang, B. 1988, ApJ, 330, 809 Reid, M. J., Readhead, A. C. S., Vermeulen, R. C., & Treuhaft, R. N. 1999, ApJ, 524, 816 Rickett, B. J. 1990, ARA&A, 28, 561 Urry, C. M., Padovani, P. 1995, PASP, 107, 803 -- 12 -- Whiteoak, J. B. & Gardner, F. F. 1983, MNRAS, 205, 27P This preprint was prepared with the AAS LATEX macros v5.2. -- 13 -- Table 1. Observational Parameters Instrument VLBA+Y27 VLA Observing date(s) Position (J2000.0) Synthesized beam Flux density calibrator Phase calibrator Bandpass calibrator Rest frequency Number channels Channel spacing Velocity Resolutiona Center velocity Total velocity range Typical noise per channel aafter Hanning smoothing 2003 May 17 & 22 2005 October 11, 13, & 14 17 47 20.0463, -28 23 46.587 17 47 19.8562, -28 22 12.900 (A) 20 × 15 mas, P.A.= 1◦ (D) 35 × 21 mas, P.A.= 1◦ 17 47 19.96, -28 22 59.8 10.′′7 × 9.′′0, P.A.= 88◦ -- J1745-283 J1733-130 3C286 J1751-253 J1733-130 4829.6569 MHz 4829.6590 MHz 128 1.953 kHz 0.24 km s−1 63 3.052 kHz 0.19 km s−1 50.6 km s−1 & 75.6 km s−1 51.4 km s−1 & 74.6 km s−1 ±7.5 km s−1 35 mJy beam−1 ±5.5 km s−1 5 mJy beam−1 Table 2. VLBA+Y27 H2CO Results from 2003 May Source α (J2000) δ (J2000) vLSR (km s−1) ∆vFWHM (km s−1) S0 θb (mJy beam−1) mas mas θa P.A. deg A D 17 47 19.856(1) −28 22 12.99(2) 23 46.59(2) 20.047(2) 75.33(1) 50.07(2) 0.59(2) 0.36(5) 645(15) 160(20) 13(3) 10(4) 5(2) 8(4) 110(20) 5(40) -- 1 4 -- -- 15 -- Table 3. VLA H2CO Results from 2005 October Source α (J2000) δ (J2000) vLSR A B D E F G H I 17 47 19.94 −28 22 13.0 22 40.6 23 47.2 20.04 20.01 18.64 24 24.5 19.61 19.57 20.43 22 13.5 23 49.9 23 46.7 24.72 21 43.0 (km s−1) 75.31(7) 51.0(2) 50.1(1) 53.9(1) 49.1(1) 51.5(1) 76.3(2) 48.7(3) 70.6(3) 73.1(1) 74.6(1) 70.9(1) ∆vFWHM (km s−1) S0 Notes (mJy beam−1) 0.71(2) 0.4(2) 0.75(5) 0.88(9) 0.94(4) 0.60(6) 0.3(1) 0.8(2) 0.3(2) 0.9(5) 0.8(3) 0.92(6) 1900 80 510 320 230 120 40 120 50 50 70 60 U U F F U U U U U,N Note. -- F -- full velocity range of features observed earlier not covered N -- possible new velocity feature U -- position, velocity, and/or intensity uncertain due to blending (see §2.2) -- 16 -- Table 4. Integrated Flux Density Comparisons 1983 -- 2005 Source Integrated Flux Density (mJy km s−1) 2005c 1983a (VLA) (VLA) 2003c (VLBA) 1993b (VLA) A B D E F G H I 720±15 80±15 700±25 230±80 < 60 < 30 140±60 < 60 850±15 150±10 1400±30 350±20 70±15 320±15 350±15 100±15 640±50 1900±400 -- 55±20d -- -- -- -- -- 20±10 1200±300 330±30 40+40 −20 75+50 −25 35±20 75±20 afrom Whiteoak and Gardner 1983 bfrom MGP94 cthis paper dIncludes only one of the velocity components de- tected in the VLA observations -- 17 -- Table 5. Nearby Masers Source H2O A F H CH3OH B D E F I OHd A B C D H Displacement from H2CO Maser ∆θa (′′) ∆vLSR (km s−1) ∆ℓb (AU) 0.6 0.1 0.2 0.7 0.7 0.7 0.8 0.5 4.5 0.5 0.6 0.6 0.15 0.05 0.06 0.18 0.07 0.10 0.21 0.24 5000 850 1700 5500 5500 6000 7400 4200 38000 4200 5000 5000 130 40 50 150 60 850 180 200 0.1 18.8 9.9 6.6 11.9 c c c c c 2.5 1.2 0.2 0.2 0.2 0.2 0.9 1.1 0.1 2.0 aangular separation bprojected linear separation cnumerous velocity components dZeeman shifting makes velocity comparison uncertain References. -- McGrath, Goss, & De Pree (2004); Houghton & Whiteoak (1995); Argon, Reid, & Menten (2000) -- 18 -- 50 100 150 (N) F A B C (M) H D (S) G E 18 16 22 20 RIGHT ASCENSION (J2000) ) 0 0 0 2 J ( I I N O T A N L C E D I -28 22 00 30 23 00 30 24 00 30 17 47 26 24 Fig. 1. -- Image of the continuum radiation at 4.8 GHz from the Sgr B2 region from the 1993 VLA-BnA observations (MGP94). The contours are -6, 6, 24, 64, 128, 256, and 384 times the image RMS noise level of 0.4 mJy beam−1 and the greyscale at the top is in units mJy beam−1. The beam is 1.′′12 × 0.′′88 at a position angle of 52◦. The image has not been corrected for attenuation by the primary beam of the VLA antennas (FWHM approximately 10′). The labelled crosses indicate the locations of the H2CO maser regions (after MGP94) and the italicized letters in parentheses denote the common designations of the major star- forming complexes. Masers A and D were detected using the VLBA+Y27. -- 19 -- 200 300 400 500 600 -28 22 12.84 12.86 12.88 12.90 12.92 12.94 12.96 12.98 13.00 13.02 13.04 ) 0 0 0 2 J ( I N O T A N L C E D I 17 47 19.868 19.866 19.864 19.862 19.860 19.858 19.856 19.854 19.852 RIGHT ASCENSION (J2000) Fig. 2. -- VLBA+Y27 image of the H2CO A maser at vLSR = 75.4 km s−1. The contour levels are −4, 4, 8, 12 and 16 times the image RMS noise level of 40 mJy beam−1 (no negative contours appear). The beam (lower left corner) is 26 × 9 mas at a position angle of 4◦. The greyscale at the top is in units of mJy beam−1. (inset) Spectrum at the A maser image peak. The observed properties are summarized in Table 2. The brightness temperature of the emission is 5.9 × 108 K. -- 20 -- Fig. 3. -- VLA spectrum of the E maser from the 2005 observations. The beam is 10.′′7 × 9.′′0 at position angle 88◦ and the velocity resolution is 0.19 km s−1.
astro-ph/9706176
1
9706
1997-06-17T18:38:52
Supersonic motions in dark clouds are not Alfven waves
[ "astro-ph" ]
Supersonic random motions are observed in dark clouds and are traditionally interpreted as Alfv\'{e}n waves, but the possibility that these motions are super-Alfvenic has not been ruled out. In this work we report the results of numerical experiments in two opposite regimes: beta ~1 and beta << 1, where beta is the ratio of gas pressure and magnetic pressure: beta=P_g/P_m. Our results, combined with observational tests, show that the model with beta ~ 1 is consistent with the observed properties of molecular clouds, while the model with beta << 1 has several properties that are in conflict with the observations. We also find that both the density and the magnetic field in molecular clouds may be very intermittent. The statistical distributions of magnetic field and gas density are related by a power law, with an index that decreases with time. Magnetically dominated cores form early in the evolution, while later on the intermittency in the density field wins out.
astro-ph
astro-ph
Supersonic motions in dark clouds are not Alfv´en waves Theoretical Astrophysics Center, Juliane Maries Vej 30, DK-2100 Copenhagen, Denmark Paolo Padoan Astronomical Observatory and Theoretical Astrophysics Center, Juliane Maries Vej 30, DK-2100 Ake Nordlund Copenhagen, Denmark ABSTRACT Supersonic random motions are observed in dark clouds and are traditionally interpreted as Alfv´en waves, but the possibility that these motions are super-Alfv´enic has not been ruled out. In this work we report the results of numerical experiments in two opposite regimes; β ≈ 1 and β ≪ 1, where β is the ratio of gas pressure and magnetic pressure: β = Pg/Pm. Our results, combined with observational tests, show that the model with β ≈ 1 is consistent with the observed properties of molecular clouds, while the model with β ≪ 1 has several properties that are in conflict with the observations. We also find that both the density and the magnetic field in molecular clouds may be very intermittent. The statistical distributions of magnetic field and gas density are related by a power law, with an index that decreases with time. Magnetically dominated cores form early in the evolution, while later on the intermittency in the density field wins out. Subject headings: turbulence - ISM: kinematics and dynamics- magnetic field 1. Introduction The observation of supersonic motions in molecular clouds (eg Zuckerman & Palmer 1974) raised the question of how these motions could be supported (Norman & Silk 1980; Fleck 1981; Scalo & Pumphrey 1982). Supersonic motions are expected to quickly dissipate their energy in highly radiative shocks, because of the very short cooling time of molecular gas or metal rich atomic gas (Mestel & Spitzer 1956; Spitzer 1968; Goldreich & Kwan 1974). Strictly related was the issue of the support of molecular clouds (MCs) against gravitational collapse, since it was soon realized that the observed motions could not be understood as a gravitational collapse (Zuckerman & Evans 1974; Morris et al. 1974), although MCs contain many Jeans' masses. Theoreticians therefore formulated the hypothesis that MCs were primarily magnetically supported (Mestel 1965, Strittmatter 1966; Parker 1973; Mouschovias 1976a,b; McKee & Zweibel 1995) and interpreted the observed motions as long-wavelength hydro-magnetic waves (Arons & Max 1975; Zweibel & Josafatsson 1983; Elmegreen 1985, Falgarone & Puget 1986). It was also shown that the properties of the observed 'turbulence' (Larson 1981; Leung, Kutner & Mead 1982; Myers 1983; Quiroga 1983; Sanders, Scoville & Solomon 1985; Goldsmith & Arquilla 1985; Dame et al. 1986; Falgarone & P´erault 1987) could be understood if the motions were sub-Alfv´enic (Myers & Goodman 1988; Mouschovias & Psaltis 1995, Xie 1996). -- 2 -- Nevertheless, recent attempts to detect the Zeeman effect, in lines of molecules such as OH and CN, that probe regions of dense gas (Crutcher et al. 1993; Crutcher et al. 1996), resulted in a number of non-detections of the effect, and therefore in rather stringent upper limits for the magnetic field strength, despite the fact that regions expected to favor detections were targeted. It is therefore possible, that the field strength in dense molecular gas is weaker than assumed in theoretical studies. In the present work, we describe the dynamics of MCs, with the numerical solution of the equations of 3-D compressible magneto-hydrodynamics (MHD), in a regime of highly supersonic random motions. We avoid any sort of idealization of the physical system, apart from describing it as a fluid, and simply rely on the physical assumption (supported by the observations) that the motions in MCs are random. The main limitation of our numerical models, compared with MCs, is the absence of gravity. We have excluded gravity on purpose (our code is capable of handling self-gravity), because one of the aim of our work is to show that (magneto-)hydrodynamic processes alone are able to explain most of the observed properties of MCs, without the need of gravity. Although gravity is certainly responsible for the fragmentation into stars of high density regions, supersonic random motions shape the structure of molecular clouds with a minor contribution from gravity (Padoan 1995; Padoan, Jones & Nordlund 1997a; Padoan, Nordlund & Jones 1997b), apart from the possibility that gravity is the energy source of the motions on the large scale. The importance of supersonic motions in fragmenting the gas is certainly apparent on very small scales, where young and probably transient clumps are not originated by gravitational instability (Falgarone, Puget & P´erault 1992; Langer et al. 1995). Previous numerical studies have shown that compressible turbulence can qualitatively explain several observational properties of MCs, even if the effect of gravity and magnetic fields are not included. The first two-dimensional (2-D) simulations of turbulence, with rms Mach number larger than one, were performed by Passot & Pouquet (1987). These were the first simulations where shocks are shown to develop inside a turbulent flow. The importance of shock formation inside mildly supersonic flows was later confirmed in 3-D simulations by Lee, Lele & Moin (1991). It was immediately recognized that shocks might have been responsible for the fragmentation of the density field inside MCs, and especially for the origin of their filamentary structure (Passot, Pouquet & Woodward 1988). Kimura & Tosa (1993) simulated the passage of a strong shock through a turbulent molecular cloud, and found that this process can generate dense clumps, with a power law mass spectrum. V´azquez-Semadeni (1994) made use of 2-D numerical simulations to show that supersonic turbulence generates a very intermittent density field, reminiscent of the clumpy nature of MCs. The density field was also found to be self-similar, which could be the reason for the hierarchical structure of MCs (Scalo 1985, Houlahan & Scalo 1992). Falgarone et al. (1994), analyzing the numerical simulation by Porter, Pouquet & Woodward (1994), argued that the properties of the profiles of molecular emission spectra from MCs can be interpreted as arising from turbulent motions. Other numerical works included gravity in the turbulent flows, yet without describing the magnetic field. Turbulence was shown to be able to prevent gravitational collapse (cf Chandrasekhar 1958; Arny 1971; Bonazzola et al. 1987, 1992) in the 2-D numerical simulations by L´eorat, Passot & Pouquet (1990). V´azquez-Semadeni, Passot & Pouquet (1995) modeled the galactic disc, on the scale of 1 Kpc, as a turbulent self-gravitating flow. They simulate a 2-D turbulent flow that is forced by the energy released by star formation (expansions of HII regions), and found that the main mechanism of cloud formation is the turbulent ram pressure, rather than gravity. They were not able to form self gravitating clouds, due to limitations in the thermal modeling, and the consequent low density contrast. A 3-D description of a magnetized self-gravitating cloud was given by Carlberg & Pudritz (1990), and was used to simulate molecular emission spectra by Stenholm & Pudritz (1993). Carlberg & Pudritz -- 3 -- found that the magnetic field and hydro-magnetic waves can support the cloud against gravity. The clouds contract, because of ambipolar diffusion, on a timescale of approximately four free-fall times. These simulations do not solve the MHD equations, but instead make use of a 'sticky particles' code. Energy is injected in the form of a spectrum of Alfv´en waves, and the outcome of the computation is dependent on the spectral index, that is a free parameter. This way of forcing the particles is rather unphysical, because an arbitrary spectrum of Alfv´en waves is imposed, instead of being obtained as a result of the simulated magneto-hydrodynamics. Passot, V´azquez-Semadeni & Pouquet (1995) introduced the magnetic field in their previous 2-D model for the galactic disc (V´azquez-Semadeni, Passot & Pouquet 1995), and obtained a flow with rough equipartition of kinetic and magnetic energy, probably in rough equipartition also with the mean thermal energy. The same simulation, and others with larger density contrast and resolution, have been studied by V´azquez-Semadeni, Ballesteros-Paredes & Rodr´iguez (1997), who were able to reproduce the observed relation between line-width and size (Larson 1981). Gammie and Ostriker (1996) solved the MHD equations in a slab geometry, including self-gravity. By forcing the flow with a nonlinear spectrum of MHD waves, they were able to prevent the gravitational collapse of their 1-D cloud model. Apart from the intentional exclusion of gravity, we have improved significantly on all previous calculations of turbulent flows. First of all we have solved the MHD equations in three-dimensions, while all previous solutions are in two or one dimensions. In MHD the dimensionality of the flow has a fundamental importance, because it determines the topological freedom of the magnetic field. Another improvement of our work is the high rms Mach number of the flows (∼ 5), while previous models are only mildly supersonic (Mach number ∼ 1). While previous models of magnetized clouds focused on the role of the magnetic field as opposed to gravity, and therefore have assumed a magnetic pressure much larger than the gas pressure, the main purpose of the present work is to show that MCs are well described as flows with much lower magnetic pressure than previously assumed, and probably in rough equipartition with the gas pressure. We report on the results of two numerical models. In model A, β ≈ 1, in model B, β ≪ 1, where β is the ratio of gas pressure and magnetic pressure: β = Pg/Pm. The observations of magnetic field strengths are in agreement with a scenario where the mean magnetic pressure, hB2ai, is dynamically low (model A). Magneto-hydrodynamic (MHD) flows develop a very intermittent spatial distribution of the magnetic energy, and therefore, when the field is detected at a favorable position, its strength is far stronger than the mean field strength. Dense cores with sub-Alfv´enic velocity dispersion can still be generated, in agreement with the observations. 2. The experiments The study of the dynamics of MCs belongs to the field of random MHD flows, in a highly supersonic regime. The Reynolds number and the magnetic Reynolds number in MC flows are very large. Their random nature is therefore a basic feature of the dynamics of these flows, and requires an appropriate description. For this reason, a realistic description of the dynamics of molecular clouds had to be based on the numerical solution of the compressible MHD equations in three dimensions, in a random and highly supersonic regime. We did it with relatively high resolution (1283), with a code designed for turbulence and MHD turbulence experiments (Nordlund, Galsgaard & Stein 1994; Stein, Galsgaard & Nordlund 1994; Galsgaard & Nordlund 1996; Nordlund, Stein & Galsgaard 1996; Nordlund & Galsgaard 1997), specifically adapted to be able to deal with very strong shocks and very large density contrasts. -- 4 -- One of our purposes with these particular experiments is to explicitly address the classical arguments about dissipation time scales for super- and sub-Alfv´enic motions, and we therefore perform the experiments in the spirit of "freely decaying turbulence"; studying the decay of an initial, solenoidal velocity field, without applying any external forcing. We include external forcing in other experiments that we have made (e.g., Padoan et al. 1997a, 1997b). Although we have already developed a version of the code with the inclusion of the gravitational force, all the experiments were run without gravity, for the following reasons: • We are mainly interested in studying the magneto-hydrodynamics, rather than the gravitational instability. • The observed motions have velocities comparable with the virial velocity, or larger, on a range of scales, and the clouds are not free-falling. • If the results of our experiments are discussed only up to a time shorter than or comparable to the dynamical (or free-fall) time, all our conclusions remain basically unchanged. This time is about a few million years on a scale of 10 pc, and clouds are not supposed to live much longer than that, before star formation takes place and becomes energetically important. We solve the compressible MHD equations: 2.1. The equations ∂ ln ρ ∂t + v · ∇ ln ρ = −∇ · v, ∂v ∂t + v · ∇v = − 1 ρ ∇P + 1 ρ j × B + f , ∂e ∂t + v · ∇e = − P ρ ∇ · v + QJoule + Qviscous + Qradiative, ∂B ∂t = ∇ × v × B, j = ∇ × B, (1) (2) (3) (4) (5) plus numerical diffusion terms, and with periodic boundary conditions. v is the velocity, B the magnetic field, f an external force (= 0 in these particular experiments), and p = ρT is the pressure at T ≈ const. Conditions in the cold molecular clouds that we are modeling are such that an isothermal approximation is adequate; the radiative heat exchange is so efficient that the temperature remains low in most places. Even if the temperature momentarily increases in shocks, the subsequent cooling is rapid, and the result is shock structures that are qualitatively and quantitatively similar to isothermal shocks. -- 5 -- We have thus used isothermal conditions in most of our runs and have verified that this is appropriate, by rerunning segments of some experiments using the full energy equation. No significant change of the statistics was found and, since using the full energy equation increases the cost of the experiments considerably (the strong cooling required to maintain a low temperature forces a much smaller time step), we performed most of the experiments at constant temperature. The absence of an explicit resistivity η in the induction equation corresponds to an assumption of flux freezing on well resolved scales. The code uses shock and current sheet capturing techniques to ensure that magnetic and viscous dissipation at the smallest resolved scales provide the necessary dissipation paths for magnetic and kinetic energy. As shown by Galsgaard & Nordlund (1976, 1977), dissipation of magnetic energy in highly turbulent MHD plasmas occurs at a rate that is independent of the details of the small scale dissipation. In ordinary hydrodynamic turbulence the corresponding property is one of the cornerstones of Kolmogorov (1941) scaling. We have not included ambipolar diffusion in the present experiments, because it occurs on a time-scale significantly longer than the dynamical time, as recently shown by Myers & Khersonsky (1995) and is thus expected to be of secondary importance. However, our code has the ability to handle ambipolar diffusion, and we plan to study its effects in a subsequent paper. 2.2. The code The code solves the compressible MHD equations on a 3D staggered mesh, with volume centered mass density and thermal energy, face centered velocity and magnetic field components, and edge centered electric currents and electric fields (Nordlund, Stein & Galsgaard 1996). The original code works with "per-unit-volume" variables; mass density, momenta, and thermal energy per unit volume. In the super-sonic regime relevant in the present application, we found it advantageous to P rewrite the code in terms of "per-unit-mass" variables; lnρ, u, and E = 3 ρ . With these variables, the time evolution of all variables is governed by equations of the type 2 Df Dt = ∂f ∂t + v · f = ...; (6) i.e., equations that specify the time rate of change following the motion. These are better conditioned than the divergence type equations that result from using per-unit-volume variables (the large -- order M 2 -- density variation in isothermal shocks cause the per-unit-volume fluxes to vary over several orders of magnitude). We use spatial derivatives accurate to 6th order, interpolation accurate to 5th order, and Hyman's 3rd order time stepping method (Hyman 1979). In order to minimize the viscous and resistive influence on well resolved scales, we use monotonic 3rd order hyper-diffusive fluxes instead of normal diffusive fluxes, and in order to capture hydrodynamic and magneto-hydrodynamic shocks we add diffusivities proportional to the negative part of the velocity divergence, and resistivity proportional to the negative part of the cross-field (two-dimensional) velocity divergence. Further details of the numerical methods are given by Nordlund, Galsgaard & Stein (1996) and Nordlund & Galsgaard (1997). -- 6 -- 2.3. Weak and strong magnetic field For the purpose of this work we have run two experiments: one with βi = 4 (model A), and the other with βi = 0.04 (model B), where βi = (Pg/Pm)i is the initial ratio of gas and magnetic pressure. In both experiments the initial density is uniform, and the initial velocity is random. We generate the velocity field in Fourier space, and we give power, with a normal distribution, only to the Fourier components in the shell of wave-numbers 1 < k < 2. We perform a Helmholtz decomposition, and use only the solenoidal component of the initial velocity. The initial magnetic field is uniform, and is oriented parallel to the z axis: B = B0z. In experiment B, the initial velocity field is perpendicular to the magnetic field (zero z component), in order to excite the Alfv´en waves, while in experiment A the initial velocity field has equal amplitude in all three components. Both experiments are decaying flows, because no external forcing is applied. We define the Alfv´enic Mach number, MA, as the ratio of the flow rms velocity and the Alfv´en velocity. In model A, MA ≈ 10, and in model B, MA ≈ 1, initially. Under isothermal conditions, in our units, the ordinary Mach number M = MA/β1/2, so the ordinary Mach number is initially M ≈ 5 in both cases. Fig. 1 shows the time evolution of the rms Mach number and rms density in the two experiments. In both cases, the rms density grows to the value σn ≈ 0.5M, where M is the rms Mach number of the flow, in about one dynamical time, defined as the ratio of half the linear size of the box over the initial rms velocity, tdyn = 0.5Lbox/σv,0. The initial Mach number in experiment B is a bit smaller than in experiment A, because the amplitude of the initial z component of the velocity field is zero. Nevertheless, energy is immediately transferred to the z component of the velocity, and the rms z component is subsequently about half the value of the other two components. Fig. 2 illustrates the time evolution of the kinetic and magnetic energies, expressed in units of the mean thermal energy. The time evolution of the magnetic energy is totally different in the two experiments. In model B the field lines are only weakly perturbed by the flow, and exchange their energy with the flow periodically (oscillations in magnetic and kinetic energy, Em and Ek); in model A, instead, the magnetic field is almost passively advected, and its intensity and direction are strongly influenced by compressions and stretching of field lines. Em grows until it is in approximate equipartition with the thermal energy of the gas, but remains well below equipartition with the kinetic energy. In this section we compare our numerical results with observed properties of molecular clouds. 3. Observational tests 3.1. Dissipation of supersonic motions The time evolution of the kinetic energy is plotted in Fig. 2. After two dynamical times, t = 2.0tdyn, the flow in experiment A still contains about 30% of its initial kinetic energy or, in other words, the rms velocity is still about 55% of its initial value. In experiment B, the mean kinetic energy of the oscillations after two dynamical times is about 40% of its initial value, i.e. the rms velocity is about 63% of the initial -- 7 -- value. Thus, contrary to the beliefs that partly motivated developing the Alfv´en wave model of MC turbulence, the dissipation in model A is not particularly rapid in absolute terms, and the dissipation in model B is not particularly small. Thus, the advantage of the Alfv´en wave model, as far as the energy dissipation time-scale is concerned, is little: tdiss,B = 1.3tdiss,A. Although this was already appreciated in Zweibel & Josafatsson (1983) and Elmegreen (1985), it should be noticed that here we do not include the dissipation mechanism of ion-neutral friction. Alfv´en waves dissipate mainly because they transfer their energy to motions along the magnetic field, that are eventually dissipated in shocks. This was found also by Gammie and Ostriker (1996). It is clear therefore that, from the point of view of the dissipation mechanism, model B is not significantly superior to model A. The dissipation of the supersonic motions is not a problem for any of the models, however, because: • If motions are present there must be a source for their energy; such a source may be active for more than one dynamical time. • After one dynamical time 50% of the initial energy in model A is still in the flow. At the scale of 10 pc the dynamical time is about 3 × 106 years: clouds do not need to exist much longer than that before star formation significantly affects their energetics (Blitz & Shu 1980). 3.2. Magnetic energy probability distribution We illustrate the probability distribution of the magnetic energy in units of the mean thermal pressure: 1/β = B2/ < Pgi. We recall that initially β is uniform in both models, and its value is βi = 4 in model A, and βi = 0.04 in model B. The distribution of the magnetic energy is shown in Fig. 3. Model A develops a very intermittent distribution, with an exponential tail. Taking into account the relation between field strength and density (see next section), one finds that approximately 0.5% of the total mass of the system contains a field 10 times stronger than the mean value. For example, in a cloud of 104M⊙, one can find a couple of clumps of about 30M⊙ each, with a field strength B = 40µG, while the mean field of the cloud is only B = 4µG. Model B is much less intermittent than model A. 3.3. The relation B-n In regions of maser emission, at densities of about n = 107cm−3, a field strength of the order of B = 103 − 104µG is observed, while in regions of molecular emission, with approximately n = 102 − 103cm−3, the field is found to be of the order of B = 10µG (eg Myers & Goodman 1988). A relation of the type B ∝ n0.3−0.6 may be deduced from the observations (eg Troland, Crutcher & Kaz`es 1986; Heiles 1987; Dudorov 1991), but it is quite uncertain, especially in the light of the above discussion about the intermittency of the distribution of the magnetic energy. The fact remains, however, that observationally the field strength certainly grows with the gas density, -- 8 -- and this is found over a range of 6 orders of magnitudes in density, and 3 orders of magnitude in field strength. Fig. 4 shows the relation B − n in models A and B. The dispersion in the relation B − n is shown in the form of 1 − σ 'error' bars in the plot of Fig. 4. In model A the slope of the relation decreases with time. We find B ∝ n0.8 at t = 0.2tdyn, B ∝ n0.7 at t = 1.0tdyn, B ∝ n0.5 at t = 2.0tdyn, and B ∝ n0.3 at t = 3.0tdyn. We interpret this time evolution of the relation B − n in the next section. In model B there is no significant correlation between B and n. The lack of correlation between B and n was to be expected: most of the density enhancement occurs by convergent flows along the field lines, in the z direction, and these motions do not affect the magnetic field. A nice illustration of this is provided by the snapshots in Fig. 5 (lower row), where the density field is stratified in planes that are predominantly perpendicular to the z direction (vertical direction in the images). The densest regions develop at nodes in the large-scale waves, as found by Carlberg & Pudritz (1990). One could possibly argue that the lack of correlation of B and n is consistent with the majority of observations, because no field is detected, and hence nothing can be said about the correlation. However, as further discussed in Section 4 below, the upper limits of the non-detections speak against this interpretation. Also, the cases where a field is detected would then have to be explained with ad hoc arguments, rather than as a natural part of a statistical distribution. An interesting consequence of the model A results is that the observed relation could in principle be used to set an independent estimate of the age of molecular clouds: the comparison between model A and the observations would indicate that molecular clouds, on the average, have survived, since they were almost uniform in density and field distribution, for 2 or 3 dynamical times, that is about 107 years, on a scale of 20 pc. This estimate has little to do with the estimate of the time-scale for massive stars to destroy the parent molecular cloud (Blitz & Shu 1980), since that calculation has nothing to say about the lifetime of the cloud after it cooled from diffuse warm gas. To conclude, note that Fig. 4 illustrates that, even if in model A the mean magnetic field is such that the hβi ≈ 0.7, in regions only 10 times denser than the mean, the field can be as intense as in model B. 3.4. Cloud and flow topology An understanding of the spatial structure of the density field in dark clouds is very important for a correct interpretation of observational data, and for the formulation of a number of physical models. The snapshots in Fig. 5 give an idea of the dimensionality of the structures in the density field. It is apparent that experiment B (lower row of panels) develops a stratified density field, with sheet-like structures, the sheets being approximately perpendicular to the magnetic field (the z direction corresponds to the vertical direction in the figures). This is consistent with the fact that only the motions in the z direction can compress the gas to very high density. The topology of the density field in experiment A (upper row of panels in Fig. 5) has a clear evolution in time. In the very beginning, until t ≈ 0.4tdyn, the density grows predominantly in sheets. These are the fronts of blobs of coherent motion, advancing at supersonic velocity. Later, these fronts start to intersect each other, and the density increases especially in filaments (at the intersections of fronts). The evolution continues with the intersection of filaments into knots, at t ≈ 1.5tdyn. The fully developed topology, at -- 9 -- t = 2.0tdyn, is characterized by both filaments and knots (cores). The evolution of the magnitudes of the mass density and the magnetic field may be discussed with reference to Lagrangian version of the continuity equation, and the scalar induction equation D ln ρ Dt = −∇ · v, D ln B Dt = −∇⊥ · v, (7) (8) where ∇⊥· stands for the divergence perpendicular to the magnetic field, following both the motion and the change of orientation of the field lines. Although −∇ · v vanishes for the solenoidal initial condition, the supersonic motions rapidly lead to the formation of shock fronts, where the local value of −∇ · v is large and positive because of the discontinuity in the velocity perpendicular to the shock. The initially homogeneous magnetic field is carried along by the perpendicular components of the velocity field, and is hence also collected into sheets, except at those rare locations where the initial field happens to be strictly parallel to the velocity field. This explains why sheets initially form in both mass density and magnetic flux density, and why the B − n relation initially has an exponent close to unity. Note that the usual argument for a slope of 2/3, that applies to isotropic and non-shocking compressible motion does not apply here, because of the development of discontinuities. In term of Eqs. 7 and 8, the 2/3 follows if −∇⊥ · v typically picks up two of the three directional contributions to −∇ · v. However, at a shock, the divergence is dominated by the derivative in one particular direction; the one perpendicular to the shock front. The magnetic field that is swept into the discontinuity quickly becomes almost parallel to the shock front, because the component in the plane of the shock grows exponentially with time. Thus, as long as the topology is dominated by sheets, the mass density and the magnetic flux grow more or less in unison in the sheets, corresponding to an exponent in the B − n relation close to unity. In the subsequent evolution, there are effects that tend to reduce the exponent in the B − n relation. First, the non-linear evolution of the initially solenoidal velocity field also leads to the development of regions of space with a positive divergence, in which both the mass density and the magnetic flux density decline. In these regions, there is no particular dominance of the cross-field divergence, and thus the three-dimensional divergence picks up un additional contribution relative to the two-dimensional divergence, consistent with the classical 2/3 argument outlined above. In experiment B the motions are mainly Alfv´en waves, and therefore the velocity is predominantly perpendicular to the direction of the magnetic field. This is illustrated in Fig. 6, where we have plotted the histogram of cos(α), where α is the angle between v and B. Note that, even though motions across the field lines are the most common ones, it is the less frequent motions along the field lines that dominate the dissipation, as mentioned in connection with the discussion of Fig. 5 above. The motions across field lines are subject to magnetic restoring forces, and do not lead to density enhancements. It is the motions along the field lines that lead to the sheet like density enhancements visible in Fig. 5. In experiment A, the magnetic field is advected by the flow, and the stretching of field lines instead produces some alignment between v and B, already before one dynamical time has passed, as illustrated in Fig. 6. Alignment between B and v may be caused by two, complementary effects: 1) Dynamical alignment -- 10 -- is expected when the magnetic energy approaches and exceeds the kinetic energy; the Lorentz force then forces the flow to be predominantly along the magnetic field lines. 2) Kinematic alignment occurs when a spatially non-uniform velocity field causes stretching of magnetic field lines, and hence a correlation of B and v. Pure shear, for example, tends to align an embedded magnetic field with the direction of the flow. Motions that are aligned with the magnetic field affect the mass density without affecting the magnetic flux density. In particular, the non-linear concentration into first sheets and then filaments due to the interaction of shock fronts continues into the formation of knots in the density field, by the convergence of matter flowing along filaments. There is no corresponding process available to a divergence free vector field such as the magnetic field; once the field has concentrated into filaments, it cannot concentrate further; the magnetic field in a filament is insensitive to flow along the filament. In the same way that converging flows along the magnetic field may lead to extreme concentrations of mass, those regions where the flow is diverging along magnetic field lines may lead to extreme rarefactions of mass, without affecting the magnetic flux density. In B − n scatter plots, this corresponds to the development of more extreme excursions of the mass, relative to those of the magnetic flux density, and hence a flattening of the B − n relation with time. In model A, dynamical alignment is at most significant in the few cores that develop a strong (sub-Alfv´enic) magnetic field in the early evolution of the experiment. In scatter plots of B against n most contributions come from regions where dynamical alignment is unimportant. We thus conclude that the evolution of the B − n relation in model A, towards a smaller exponent with time, is caused by the kinematic alignment of B and v. 3.5. Stellar extinction Padoan, Jones, & Nordlund (1997) have shown that near-infrared stellar extinction determinations can be used to infer the 3-D probability distribution of the gas density in dark clouds. They have shown that there is qualitative and quantitative agreement between the inferred 3-D density distribution in dark clouds, and the one produced by their experiment of random supersonic flows. The method is based on the plot of the dispersion of the extinction measurements in cells, versus the mean extinction in the same cells (Lada et al. 1994). In Padoan, Jones & Nordlund (1997), the theoretical plots are produced by generating random density fields of given statistics and power spectra. Here we show one example of the same plot, but produced directly from the density field of experiments A and B. The plot is shown in Fig. 7, that can be compared directly with the plots shown in Padoan, Jones, & Nordlund. The plot of model A looks like the observational one. It has a smaller slope because the rms Mach number in Experiment A is smaller than in the observed cloud. The plot for Experiment B seems instead to differ from the observational one. The absence of locations with very low extinction in model B is due to the fact that the density field in experiment B is mainly organized in sheets, perpendicular to the magnetic field direction. -- 11 -- 4. Discussion It is difficult to get an objective view on the magnetic field strength in dark clouds from the literature. The reasons are the following: • Negative results from observational programs, in which detections have been reported, often remain unpublished. • The positions searched for Zeeman splitting never represent a statistically meaningful sample. Favorable regions are always selected, because the observations are very time consuming. • The total number of regions in dark clouds, for which OH Zeeman observations are published, is still small. The best study of OH Zeeman splitting in dark clouds we are aware of is by Crutcher et al. (1993). They selected 4 cores in the Taurus dark cloud complex, 2 in the Libra complex, 2 in ρ Oph, 1 in the Orion molecular ridge, 1 position in L889, and the core of B1 (Barnard 1), in the Perseus region. The only certain detection is in the cloud B1. For the other regions, the weighted average value of the field is +2.7 ± 1.5µG in Taurus, −2.1 ± 2.8µG in Libra, +6.8 ± 2.5µG in ρOph, −0.6 ± 2.1µG in L889, and −4.7 ± 3.5µG in L1647. Two well known regions of very intense magnetic field are Orion A and Orion B. OH Zeeman splitting in Orion A revealed a magnetic field strength of B = −125µG (Troland, Crutcher, & Kaz`es 1986), and B = +38µG toward Orion B (Crutcher, & Kaz`es 1983). Crutcher et al. (1996) have reported the first attempt to measure CN Zeeman splitting. They observed towards the cores OMC-N4 and S106-CN. They did not detect the magnetic field, and their upper limits are several times smaller than the value expected if the observed motions were sub-Alfv´enic. An average field of +9µG was found in Cas A by Heiles and Stevens (1986). The OH Zeeman splitting should probe regions with n = 103 − 104 cm−3, while the CN Zeeman splitting, should probe regions of n ≈ 106 cm−3. Note that all works are biased toward regions of strong field, because only regions considered to be favorable for magnetic field detections are selected for Zeeman splitting observations. Despite that, in many cases the field is not detected, and upper limits are quite stringent. In our model A we find that, if the mean field strength is hB2i1/2 = 5µG, and the mean density hni = 103cm−3, at the time t = 2.0tdyn, then 20% of the total mass is in dense clumps, with density ten times larger than the mean, n ≈ 104cm−3, and field strength five times, or more, larger than the mean, B ≥ 25µG. Therefore, even if field strengths of about 30µG are detected sometimes in dense cores, the Ai1/2 ≈ CS. mean Alfv´en velocity in the molecular cloud may be just comparable to the sound speed, hv2 On the basis of the observational results by Crutcher et al. (1993, 1996), it may be concluded that model A is in consistent with the observational estimates of magnetic field strength in dark clouds. The particular values of the magnetic field used in the comparison here should not be taken too literally; the small scale field strengths could be larger than estimated by the Zeeman effect if the magnetic field is tangled. It is difficult to envisage how the measurements could be consistent with model B, however, since the magnetic field in a sub-Alfv´enic model by definition has an energy that exceeds the kinetic energy. -- 12 -- As demonstrated by Galsgaard & Nordlund (1996), a magnetically dominated plasma is able to quickly dissipate structural complexity, independent of the value of the resistivity. Thus, a sub-Alfv´enic field could not remain strongly tangled, and hence could not avoid detection. Using the same argument, we expect those cores where a strong (sub-Alfv´enic) field has indeed been observed to have a relatively simple magnetic field structure. Model A can be used for the description of a typical molecular cloud with linear size of about 5 pc and hni = 500 cm−3. At the time t = 1.0tdyn, the relation B − n is then: B = 5µG( n 500cm−3 )0.7 (9) Since the exponent is > 0.5, most of the dense cores are found to have magnetic pressure larger than thermal pressure, at early times. This is illustrated in Fig. 8, where we have plotted the Alfv´en velocity versus the gas density, after one dynamical time, for experiments A and B. In model A, one can see that, on average, regions with density larger than the mean are characterized by an Alfv´en velocity larger than the sound velocity, and therefore by a magnetic pressure larger than the thermal pressure. The lowering of the exponent in the B − n power law with time means that later on, the dominance of the magnetic field in these cores tends to be reduced. Although the flow is random and approximately isotropic, the kinematic alignment of B with v makes dense cores accrete mass along B at an increased rate. Therefore, the accretion of mass around dense cores, embedded in a random flow with β ≈ 1, is such that, while magnetic pressure becomes dominant over thermal pressure during the initial phase of fragmentation, later on magnetic pressure decreases to the level of thermal pressure, even in the absence of gravity, and on a timescale that is competitive with ambipolar diffusion. This mechanism could be relevant for the process of star formation. The main point we want to clarify in this paper is that the fact of finding some cloud cores, with sub-Alfv´enic velocity dispersions and with magnetic pressure larger than thermal pressure, does not necessarily mean that the dynamics of molecular clouds is dominated by MHD waves; those cores may be formed, in a few million years, in a supersonic and super-Alfv´enic flow, only marginally affected by the magnetic field (model A). The dynamics becomes strongly affected by the magnetic field only in some very dense regions, on small scales, and at early times. We stress that a key theoretical ingredient to the interpretation of OH Zeeman splitting data is the B − n relation. The fact that the exponent of the relation is > 0.5 for more than two dynamical times is the reason why sub-Alfv´enic cores can be found in the experiment. In this work we have shown that: 5. Conclusions • Supersonic motions in model A are relatively long-lived, with decay times comparable to the estimated life-time of molecular clouds. Even in the absence of an energy source, supersonic and super-Alfv´enic motions can persist as such, on a scale of 20 pc, for about 107 years. • Supersonic sub-Alfv´enic motions (model B) dissipate their energy almost as fast as super-Alfv´enic motions. The original motivation for the MHD wave model is therefore absent. -- 13 -- • Random supersonic motions produce a very intermittent probability distribution of the magnetic energy, with an exponential tail. Since the distribution is so intermittent, one could easily detect a field strength that is several times in excess of the mean strength. • Dense cores, with magnetic pressure larger than thermal pressure, and velocity dispersions smaller than vA, are found as the result of the evolution of supersonic and super-Alfv´enic flows (model A). • A power law statistical B − n relation is generated by supersonic motions, in model A, but not by MHD waves, in model B. The exponent of the relation is > 0.5 for about two dynamical times, which allows for the existence of cores with vA > CS. • The exponent in the B − n relation decreases with time, because magnetic field lines are stretched and partially aligned with the flow. The statistical importance of the magnetic pressure in dense cores is thus expected to decrease with time, even in the absence of gravity and ambipolar diffusion. • The topology of the density field generated by model A is mainly filamentary and clumpy. This is reminiscent of the morphology of molecular clouds. On the other hand, the density field produced by MHD waves in model B is mainly structured in sheets, perpendicular to the magnetic field. • The statistical properties of the density field generated by random supersonic motions (as in model A) are in agreement with the statistical properties of the gas density in dark clouds, as inferred from stellar extinction determinations. Model B does not give such a good agreement. The comparison between model A and model B shows that model A (super-Alfv´enic motions) provides a nice understanding of the dynamics and structure of molecular clouds. Moreover, it is consistent with the OH Zeeman detections. We conclude that the supersonic motions observed in dark clouds, on a scale of a few parsecs to several parsecs, can be understood, without any major problem, as super-Alfv´enic motions, although sub-Alfv´enic dense cores certainly exist. We thank Prof. L. Mestel and Prof. E. Zweibel for their comments on the manuscript. This work has been supported by the Danish National Research Foundation through its establishment of the Theoretical Astrophysics Center. Computing resources were provided by the Danish National Science Research Council, and by the French 'Centre National de Calcul Parall`ele en Science de la Terre'. REFERENCES Arny, T. 1971, ApJ, 169, 289 Arons, J. & Max, C. E. 1975, ApJ, 196, L77 Blitz, L. & Shu, F. H. 1980, ApJ, 238, 148 Bonazzola, S., Falgarone, E., Heyvaerts, J., Perault, M. & Puget, J. L. 1987, A&A, 172, 293 Bonazzola, S., Perault, M., Puget, J. L., Heyvaerts, J., Falgarone, E. & Panis, J. F. 1992, J. Fluid Mech., 245, 1 -- 14 -- Carlberg, R. G. & Pudritz, R. E. 1990, MNRAS, 247, 353 Chandrasekhar, S. 1958, Proc. Roy. Soc., 246, 301 Crutcher, R. M. & Kaz`es, I. 1983, A&A, 125, L23 Crutcher, R. M., Troland, T. H., Goodman, A. A., Heiles, C., Kaz`es, I. & Myers, P. C. 1993, ApJ, 407, 175 Crutcher, R. M., Troland, T. H., Lazareff, B. & Kaz`es, I. 1996, ApJ, 456, 217 Dame, T. M., Elmegreen, B. G., Cohen, R. S., Thaddeus, P. 1986, ApJ, 305, 892 Dudorov, A. E. 1991, Sov. A., 35(4), 342 Elmegreen, B. G. 1985, ApJ, 299, 196 Falgarone, E. & Puget, J. L. 1986, A&A, 162, 235 Falgarone, E., P´erault, M. 1987, in Physical Processes in Interstellar Clouds, ed. G. Morfil, M. Scholer (Dordrecht: Reidel), 59 Falgarone, E., Puget, J., -L. & P´erault, M. 1992, A&A, 257, 715 Falgarone, E., Lis, D. C., Phillips, T. G., Pouquet, A., Porter, D. H. & Woodward, P. R. 1994, ApJ, 436, 728 Fleck, R. C. Jr 1981, ApJ, 246, L151 Galsgaard, K. & Nordlund, A. 1996, J. of Geophys. Res., 101(A6),13445 Galsgaard, K. & Nordlund, A. 1997, J. of Geophys. Res., 102, 231 Gammie, C. F. & Ostriker, E. C. 1996, ApJ, 466, 814 Goldreich, P. & Kwan, J. 1974, ApJ, 189, 441 Goldsmith, P. F. & Arquilla, R. 1985, in Protostars and Planets II, University of Arizona Press, p. 137 Heiles, C., Stevens, M. 1986, ApJ, 301, 331 Heiles, C. 1987, D. J. Hollenbach & H. A. Thronson (eds.), Interstellar Processes, Reidel, p. 171 Houlahan, P. & Scalo, J. 1992, ApJ, 393, 172 Hyman, J. M. 1979, R. Vichnevetsky & R. S. Stepleman (eds.), in Adv. in Comp. Meth. for PDE's, p. 313 Kimura, T. & Tosa, M. 1993, ApJ, 406, 512 Kolmogorov, A. N. 1941, Dokl. Akad. Nauk. SSSR, 30, 301 -- 305. Lada, C. J., Lada, E. A., Clemens, D. P., & Bally, J. 1994, ApJ, 429, 694 Langer, W. D., Velusamy, T., Kuiper, T. B. H., Levin, S., Olsen, E., Migenes, V. 1995, ApJ, 453, 293 Larson, R. B. 1981, MNRAS, 194, 809 Lee, S., Lele, S. K., Moin, P. 1991, Phys. Fluids A 3, 657 -- 15 -- L´eorat, J., Passot, T., Pouquet, A. 1990, MNRAS, 243, 293 Leung, C. M., Kutner, M. L., Mead, K. N. 1982, ApJ, 262, 583 McKee, C. F. & Zweibel E. G. 1995, ApJ, 440, 686 Mestel, L. & Spitzer, L. 1956, MNRAS, 116, 503 Mestel, L. 1965, Q. Jl. R.A.S., 6, 161, 265 Morris, M., Zuckerman, B., Turner, B. E. & Palmer, P. 1974, ApJ, 192, L23 ıı Mouschovias, T. Ch. 1976a, ApJ, 206, 753 Mouschovias, T. Ch. 1976b, ApJ, 207, 141 Mouschovias, T. Ch. & Psaltis, D. 1995, ApJ, 444, L105 Myers, P. C. 1983, ApJ, 270, 105 Myers, P. C. & Goodman, A. A. 1988, ApJ, 326, L27 Myers, P. C. & Khersonsky, V. H. 1995, ApJ, 442, 186 Nordlund, A, Galsgaard, K. & Stein, R. F. 1994, in R, J, Rutten, C J. Schrijver (eds.), Solar Surface Magnetic Fields, Vol. 433, NATO ASI Series Nordlund, A, Stein, R. F. & Galsgaard, K. 1996, in J. Wazniewski (ed.), Proceedings from the PARA95 workshop, Vol 1041 of Lecture Notes in Computer Science, p. 450 Nordlund, A& Galsgaard, K. 1997, Journal of Computational Physics, (in preparation) Norman, C., Silk, J. 1980, ApJ, 238, 158 Padoan, P. 1995, MNRAS, 277, 377 Padoan, P., Jones, B. J. T. & Nordlund, A. 1997, ApJ, 474, 730 Padoan, P., Nordlund, A. & Jones, B. J. T. 1997, MNRAS, in press Parker, D. A. 1973, MNRAS, 163, 41 Parker, E. N. 1994, Spontaneous Current Sheets in Magnetic Fields, Oxford University Press Passot, T., Pouquet, A. 1987, J. Fluid Mech., 181, 441 Passot, T., Pouquet, A., Woodward, P. 1988, A&A, 197, 228 Passot, T., V´azquez-Semadeni, E. & Pouquet A. 1995, ApJ, 455, 536 Porter, D. H., Pouquet, A. & Woodward, P.R. 1994, Phys. Fluids, 6, 2133 Quiroga, R. J. 1983, Ap. Sp. Sci., 93, 37 Sanders, D. B., Scoville, N. Z., Solomon, P. M. 1985, ApJ, 289, 372 Scalo, J. M., Pumphrey, W. A. 1982, ApJ, 258, L29 -- 16 -- Scalo, J. M. 1985, in Protostars and Planets II, ed. D. C. Black and M. S. Mathews (Tucson: University of Arizona Press), 349 Spitzer, L. 1968, in Nebulae and Interstellar Matter, Stars and Stellar Systems, ed. B. Middlehurst, L. H. Aller, 7, 1, Univ. Chicago Press Stein, R. F., Galsgaard, K. & Nordlund, A. 1994, in J. D. B. et al (ed.), Proc. of the Cornelius Lanczos International Centenary Conference, Society for Industrial and Applied Mathematics, Philadelphia, p. 440 Stenholm, L. G. & Pudritz, R. E. 1993, ApJ, 416, 218 Strittmatter, P. A. 1966, MNRAS, 132, 359 Troland, T. H., Crutcher, R. M. & Kaz`es, I. 1986, ApJ, 304, L57 V´azquez-Semadeni, E. 1994, ApJ, 423, 681 V´azquez-Semadeni, E., Ballesteros-Paredes, J. & Rodr´iguez, L. F. 1997, ApJ, 474, 292 V´azquez-Semadeni, E., Passot, T. & Pouquet, A. 1995, ApJ, 441, 702 Velusamy, T., Kuiper, T. B. H., Langer, W. D., in preparation Zuckerman, B. & Evans, N. J. 1974, ApJ, 192, L152 Zuckerman, B. & Palmer, P. 1974, ARA&A, 12, 279 Zweibel, E. G. & Josafatsson, K. 1983, ApJ, 270, 511 Xie, T. 1997, ApJ, 475, L139 This preprint was prepared with the AAS LATEX macros v4.0. -- 17 -- Fig. 1. -- Time evolution of rms Mach number and rms number density. -- 18 -- Fig. 2. -- Time evolution of kinetic and magnetic energies, in units of the mean thermal energy. -- 19 -- Fig. 3. -- Distribution of magnetic energy in units of the mean gas pressure, after one dynamical time. The dotted vertical line marks the initial mean value of the magnetic energy, and the dashed line the mean value after one dynamical time. -- 20 -- Fig. 4. -- The relation B − n, after one dynamical time. The magnetic field is expressed in the numerical units, that is as an Alfv´en velocity (defined with the mean gas density), in units of the sound speed. The 1 − σ 'error' bars show the dispersion of values of B around the mean for each bin. The size of the bins is chosen in order to have the same number of measurements inside each bin. -- 21 -- Fig. 5. -- 3-D representation of the time evolution of the density field. Time evolves from left to right. The upper row of panels show model A, while the lower row shows model B. -- 22 -- Fig. 6. -- Histograms of the cosine of the angle between v and B. alignement, while in experiment B the two fields are mainly perpendicular to each other. In experiment A there is a partial -- 23 -- Fig. 7. -- Relation between the dispersion of the extinction and the mean extinction, on a regular grid superposed of the projection of the density field. The value of the extinction are measured at random positions, simulating the random position of a star behind the cloud. -- 24 -- Fig. 8. -- The Alfv´en velocity, in units of the sound velocity, versus the gas density. The dashed line is the B − n relation, that is to say the Alfv´en velocity defined with the mean density, instead of with the local density. In Experiment A, the magnetic pressure is on average larger than the gas pressure, in regions with gas density larger than the mean.
0810.1519
1
0810
2008-10-08T20:00:25
Suzaku Observations of Cyg X-1
[ "astro-ph" ]
We present highlights from a series of four simultaneous Suzaku/RXTE observations of the black hole candidate Cyg X-1. We briefly summarize several key results from our decade long RXTE monitoring campaign. We then comment on challenges of analyzing the Suzaku data, i.e., improving the aspect correction beyond that of the existing tools, and quantitatively assessing pileup. All of our Suzaku observations (one, by design) occurred at or very near orbital phase 0 (superior conjunction), and hence show evolution in color-color diagrams due to X-ray absorption by material from the wind of the secondary. We present simple partial absorption models for this evolution. We then compare the Suzaku and RXTE data, and explicitly divide the Fe line region into narrow and broad components. Both are required for the Suzaku data, and are seen to be consistent with the RXTE data. These Suzaku observations occurred near historically hard, low flux states. We present fits of the broad band spectra with a simple phenomenological broken powerlaw model, as well as a more physically motivated Comptonization model. Whereas the former class of models described nearly all of the RXTE campaign better than any physical model, here the latter model is slightly more successful. The Comptonization model, however, exhibits little evidence for a soft disk component, which formally corresponds to a small, inner disk radius. Whether this is physical, due to unmodeled absorption, or is a calibration issue, remains an open question.
astro-ph
astro-ph
Suzaku Observations of Cyg X-1 Michael Nowak∗† MIT Kavli Institute, Cambridge, USA E-mail: [email protected] We present highlights from a series of four simultaneous Suzaku/Rossi X-ray Timing Explorer (RXTE) observations of the black hole candidate Cyg X-1. We first briefly summarize several key results from our decade long RXTE monitoring campaign (which to date contains over 250 observations). We then comment on challenges of analyzing the Suzaku data, i.e., improving the aspect correction beyond that of the existing tools, and quantitatively assessing pileup. All of our Suzaku observations (one, by design) occurred at or very near orbital phase 0 (superior conjunc- tion), and hence show evolution in color-color diagrams due to X-ray absorption by material from the wind of the secondary. We present simple partial absorption models for this evolution. We then compare the Suzaku and RXTE data, and explicitly divide the Fe line region into narrow and broad components. Both are required for the Suzaku data, and are seen to be consistent with the RXTE data. These Suzaku observations occurred near historically hard, low flux states. We present fits of the broad band spectra with a simple phenomenological broken powerlaw model, as well as a more physically motivated Comptonization model. Whereas the former class of models described nearly all of the RXTE campaign better than any physical model, here the latter model is slightly more successful. The Comptonization model, however, exhibits little evidence for a soft disk component, which formally corresponds to a small, inner disk radius. Whether this is physical, due to unmodeled absorption, or is a calibration issue, remains an open question. 8 0 0 2 t c O 8 ] h p - o r t s a [ 1 v 9 1 5 1 . 0 1 8 0 : v i X r a VII Microquasar Workshop: Microquasars and Beyond September 1 - 5, 2008 Foca, Izmir, Turkey ∗Speaker. †With the assistance of Paolo Coppi (Yale), John E. Davis (MIT Kavli Institute), Manfred Hanke (Universität Erlangen-Nuremberg), Sera Markoff (University of Amsterdam), Katja Pottschmidt (CRESST/GSFC/UMBC), Sarah Trowbridge (MIT Kavli Institute), and Jörn Wilms (Universität Erlangen-Nuremberg). Anything coherent and truthful herein is largely thanks to them, while any incoherent ramblings are wholly my own. This work was supported in part by NASA Grants SV3-73016 and NNX08AE23G. c(cid:13) Copyright owned by the author(s) under the terms of the Creative Commons Attribution-NonCommercial-ShareAlike Licence. http://pos.sissa.it/ Suzaku Observations of Cyg X-1 Michael Nowak 1. Summary of RXTE Campaign Cyg X-1 is one of the best studied of the galactic black hole candidates (BHC). It has been a persistent X-ray source, exhibiting, very broadly, spectrally hard to somewhat softer X-ray states. (Whereas the hard X-ray tail softens, it never completely vanishes for Cyg X-1.) In Cyg X-1 these spectral states occur over a factor of only a few in X-ray flux. A summary of the system properties can be found in [13], while a summary of the broad band spectral behavior can be found in [22], and references therein. Whereas Cyg X-1 exhibits an interesting variety of variability behavior (see [15, 16, 1, 7, 20], etc.) and radio behavior (see [5, 6, 14, 22], etc.) correlated with the X-ray spectra and flux, here we concentrate on the X-ray spectral properties. In Fig. 1, we present some of the spectral highlights from our RXTE monitoring campaign ([22]). First, we see that the broad band spectra are extremely well-described by a very sim- ple model: an absorbed, exponentially cutoff broken powerlaw, with a broad gaussian line. The spectral break always occurs near 10 keV, with the degree of the break increasing for softer (typi- cally brighter) spectra. More physically motivated Comptonization models ([9, 22]) and X-ray jet models ([11]) fit the data nearly as well. The former fit the soft X-ray power law with a combi- nation of disk photons and Comptonization, and the (exponentially cutoff) hard X-ray powerlaw with Comptonization and reflection. The jet models also utilize these same physical components, but also allow for jet synchrotron radiation (usually in the soft X-ray), and jet synchrotron self- Compton (usually in the hard X-ray). For neither the Comptonization nor jet models is there a single continuum component underlying the Fe Ka line region. Prior studies have ascribed the G 1 -- G 2 correlation predominantly to a hardness-reflection anti- correlation ([23, 9]). Although this effect is seen somewhat by our campaign ([22], Fig. 1), it is greatly reduced from prior claims (e.g., [9]). A major difference between these studies is that we have allowed the fitted seed photon temperature to be a free parameter, whereas other studies have not (i.e., [9]). Fig. 1 shows that to some extent, a hardness-disk flux anti-correlation subsumes a substantial fraction of any hardness-reflection anti-correlation. The ability of Suzaku to measure the soft X-ray spectral regime is therefore seen to be crucial in further constraining such models. For all of our RXTE spectral fits, whether broken power law, Comptonization, or jet, a broad gaussian is found near the expected position of the 6.4 keV, Fe Ka line. However, the width of this line, and its correlations with spectral hardness, do depend upon the exact fitted model. (Broken power laws tend to produce the narrowest lines; different Comptonization models, e.g., comptt vs. eqpair yield differing line widths [22].) Furthermore, RXTE lacks the spectral resolution to decompose the residuals in the line region into broad and narrow components. Suzaku, on the other hand, does have sufficient resolution to decompose the line into components, and again is seen to be crucial for further constraining spectral models. 2. Attitude Correction and Pileup Estimation for the Suzaku Observations Before we can describe and fit the Suzaku spectra, we must first make sure that we have minimized instrumental effects. Most important among these is the reduction of photon pileup, i.e., when more than one photon lands in the same or adjacent pixels in one CCD readout frame (see [4] for a technical description). These "piled" events are either lost (due to exceeding the 2 Suzaku Observations of Cyg X-1 Michael Nowak b 1.2 1.0 0.8 2 − 1 0.6 0.4 0.2 0.0 1.5 1037 ] 1 − s g r e [ k s i d L 2.0 2.5 1 3.0 3.5 Fdisk (lh/ls)−0.19 2p / 0.5 0.4 0.3 0.2 0.1 0.0 2 4 lh/ls 6 0 2 4 lh/ls 6 8 Figure 1: A brief summary of results from our monitoring campaign of Cyg X-1 ([22]). The upper left shows unfolded RXTE-PCA/HEXTE spectra (see also [14]), fit with a simple absorbed, exponentially cutoff broken powerlaw plus broad gaussian line. (This spectrum is among the softest seen in the campaign.) The amplitude of the powerlaw break is correlated with the soft powerlaw slope (softer corresponds to a greater break), as shown in the upper right. Solid points are the Cyg X-1 hard state (as defined by [18]), and hollow points are the Cyg X-1 soft state. The lower left shows results from fitting the eqpair Comptonization model of [3], which parameterizes spectral hardness with a ratio of coronal compactness to seed photon compactness (≡ ℓh/ℓs). Here the break between the soft and hard X-ray powerlaws are a combination of anti-correlations between the spectral hardness and reflection fraction (left) and amplitude of any additional, soft X-ray disk component (right). 3 G G G (cid:181) W Suzaku Observations of Cyg X-1 Michael Nowak Figure 2: Left: An example of Suzaku attitude correction. The left half of the image shows a Suzaku PSF for Cyg X-1 using the standard attitude correction. The right half of the image shows the improvement with aeattcorr.sl. Right: Improving the attitude correction allows for a better estimate of pileup in the image. Here we show an image where discrete colors correspond to the minimum level of pileup in that region. The outer white circle is the extraction region, while the inner white circle is the excluded region (with pileup fractions as high as 30%). The average effective residual pileup level is less than 4%. event energy threshold, or migrating to bad "event grades"), or are read as a higher energy photon, therefore distorting the spectrum. (This distortion can be especially problematic at high energies, as the intrinsic photon count rate spectra tend to be (cid:181) E −1.7.) In order to assess the degree of pileup, we must first have the most accurate measure of the X-ray image. This in turn requires an accurate spacecraft attitude. For Suzaku, the latter is affected by "thermal wobbling" of the spacecraft, induced during day/night passages during the spacecraft orbit ([19]). Whereas a statistical correction has been developed to correct the spacecraft attitude ([19]), for bright sources (e.g., many AGN, and our Cyg X-1 data), the correction can be improved further by measuring the mean position of the point spread function (PSF) on short time scales, then refining the empirical correction of the spacecraft attitude. Such an algorithm, aeattcor.sl, has been developed by John Davis, and as shown in Fig. 2 it further improves the Cyg X-1 image. With this improved image, one can then calculate the mean counts per 3 × 3 pixel region per CCD readout frame, and thereby arrive at a quantitative assessment of the fractional pileup level in a given image. We have developed a tool (pileup_estimate.sl) to automate this process. It further allows one to create a region filter that excludes the central piled regions, and returns an estimate of the average pileup fraction in the remainder of the image. These tools have been made available publicly1. For each of our observations we exclude approximately 1/3 of the total counts from the center of the (attitude corrected) image. We estimate that any residual pileup fractions are <∼ 4%. 1http://space.mit.edu/CXC/software/suzaku/ 4 Suzaku Observations of Cyg X-1 Michael Nowak 5 . 2 2 2 5 . 1 1.5 2 3 2.5 1 Figure 3: As discussed by [22], RXTE observations of Cyg X-1 are well-fit by an exponentially cutoff, broken powerlaw. The low energy power law (G 1) is strongly correlated with the high energy powerlaw (G 2). The triangles correspond to the Cyg X-1 hard state, while the circles correspond to the Cyg X-1 soft state. The filled diamonds correspond to the RXTE observations simultaneous with our Suzaku observations. (Note that points are shown for eight individual color-color regions from the four separate observations.) 3. The Suzaku Observations Over the course of a year and a half, we have obtained four simultaneous Suzaku/RXTE ob- servations. (The most recent observation in April 2008 was also performed simultaneously with Chandra, XMM-Newton, INTEGRAL, and Swift.) In Fig. 3 we show the location in the G 1 -- G 2 diagram of broken powerlaw fits to the simultaneous RXTE data. We immediately see that these spectra were among the spectrally hardest observations of the past decade. Note that Fig. 3 shows eight, rather than just four points, as we have subdivided each observa- tion by its location on a color-color diagram defined by the Suzaku bandpasses (see Fig. 4). All of our observations show evolution towards the lower left corner of these diagrams, with three of the four observations rounding the corner and evolving to the right along the bottom of the diagram. These three observations (with durations of 0.2 in phase) overlapped with orbital phase 0 (i.e., superior conjunction; the April 2008 observation was specifically scheduled for that phase). The fourth observation covered orbital phase 0.2 -- 0.3. Thus all observations were subject to a series of dipping events due to absorption by the stellar wind of the secondary (see [2]). We can model the color-color diagrams by fitting a spectrum to the locus of points in the upper right hand of the plot, and then multiplying this spectrum by a range of additional column densities (above and beyond the Galactic column). This yields the downward/leftward evolution on the diagram. In order to obtain the evolution to the right along the bottom of the diagram, we need to presume a partial covering fraction of 80% (i.e., 20% uncovered) for at least the soft X-rays. Our best fit to the diagram, however, is for the middle X-rays to be 100% covered. The uncovered soft X-rays, rather than being wholly local to the source, is likely dominated by the dust scattering halo 5 G G Suzaku Observations of Cyg X-1 Michael Nowak 5 1 . 1 5 0 . ) V e k 9 − 3 ( / ) V e k 3 − 5 . 1 ( 0.2 0.4 0.6 0.8 (0.5−1.5 keV)/(1.5−3 keV) Figure 4: Color-color diagram for a Suzaku Cyg X-1 observation. The evolution to the lower left corner of the diagram is consistent with increased absorption, while the rightward extending tail indicates partial absorption. The lines show partial covering models where only the absorption is changed. From left to right, respectively, they represent: all channels having the same 82% covering fraction, the two high energy channels having the same (100%) covering fraction (the soft channel is 82% covered), and the highest energy channel being completely uncovered (the soft channel is 81% covered, the middle channel is 100% covered). Three of our Cyg X-1 observations occurred near orbital phase 0, and show similar color-color diagrams. (see [17]), a substantial fraction of which is contained within the 2 arcminute Suzaku PSF. We note that Chandra observations of such dipping are more consistent with a 95% covering fraction (see [8], and the contribution by M. Hanke in this volume). We have divided each observation into as many as three pieces - the locus of points in the upper right, the downward slope to the left, and the evolution to the right along the bottom of the diagram. These time intervals were defined by the Suzaku observations, and then RXTE spectra were extracted from the intersections with those time intervals. Throughout the rest of this work, we only consider spectra from the locus of points in the upper right. 4. The Fe Ka Line Region We next turn to the line region of the data. RXTE data have suggested a broad Fe line ([22]), but questions have remained as to contributions from narrow components, especially given the very modest resolution of RXTE2. First, we wish to address the consistency between the two detectors. We do this by fitting a simple phenomenological model (e.g., a powerlaw) simultaneously to the 3.5 -- 4.5 keV and 7.5 -- 8.5 keV data, and then noticing the ratio residuals in the full 3 -- 9 keV range. This procedure does not necessarily produce an accurate profile for any broad line in the 4 -- 8 keV region (it produces perhaps the "most optimistic" estimate of such a line), but it can show the consistency of the Suzaku and RXTE residuals. We show such ratio residuals in Fig. 5. 2Chandra resolves the narrow components of the line (see [12]). For Chandra, however, the broad components are difficult to assess with great accuracy (see [8], and the contribution by M. Hanke in this volume). 6 Suzaku Observations of Cyg X-1 Michael Nowak 5 1 . 1 1 1 . 5 0 . 1 o i t a R 1 5 9 . 0 4 6 8 Energy (keV) Figure 5: Fit residuals for RXTE (filled diamonds) and Suzaku (histogram), designed to emphasize features in the Fe Ka region. A simple power law spectrum was fit to the 3.5 -- 4.5 and 7.5 -- 8.5 keV spectra, and then the full 3.5 -- 9 keV spectra were noticed without refitting the model. The Suzaku data have been binned to half width half maximum (HWHM) of the spectral resolution. There is indeed very good agreement between the Suzaku and RXTE residuals, and both show broad deviations from a simple continuum, consistent with a relativistically broadened line. (We have not, however, included all the potential continuum complexity discussed above, i.e., disk, Comptonization, synchrotron, synchrotron self-Compton, and reflection). The Suzaku data also strongly suggest narrow components. A narrow emission feature is seen at 6.4 keV, and a possible absorption feature is seen at 6.7 keV. Based upon prior Chandra observations at orbital phase 0 ([8]), both features are expected. The emission and absorption features have comparable equivalent widths (absolute value ≈ 10 -- 30 eV), and hence do not form the main contribution to the broad residuals seen in the RXTE data (equivalent width ≈ 100-150 eV, if fitting just a single component). We see that at least three components are required to fit the broad line region residual in the RXTE data. In Fig. 6 we present such continuum plus multi-component line fits. Specifically, we include a relativistic diskline model, a narrow gaussian emission line, and a narrow gaussian absorption line. We show fits with and without an additional smeared edge. Both sets of fits require a broadened line component, with an inner radius ranging from 10 GM/c2 to 60 GM/c2. The equivalent width of the broad line component is in the 40 -- 130 eV range (and is 10% -- 30% lower if including a smeared edge). It is interesting to note that the peak in the blue wing of the line is obscured by the absorption line at 6.7 keV. We have found, however, that it is difficult to fit the width and location of the RXTE residuals without such a blue peak in the broad line component. 5. Broad Band Fits We now turn to the broad band spectra by considering models fit to the 0.7 -- 9 keV XIS data3, the 3 -- 22 keV RXTE-PCA data, and the 18 -- 200 keV RXTE-HEXTE data. As for our RXTE cam- 3For plotting purposes, all XIS spectra are summed, however, we fit each XIS data set individually. The HXD-PIN 7 Suzaku Observations of Cyg X-1 Michael Nowak 1 1 1 1 − − − − V V V V e e e e k k k k 1 1 1 1 − − − − s s s s 2 2 2 2 − − − − m m m m c c c c s s s s n n n n o o o o t t t t o o o o h h h h P P P P 2 2 2 2 V V V V e e e e k k k k o o i i t t a a R R 5 5 5 5 . . . . 3 3 3 3 3 3 3 3 5 5 5 5 . . . . 2 2 2 2 5 5 0 0 . . 1 1 1 1 5 5 9 9 . . 0 0 4 4 6 6 8 8 Energy (keV) Energy (keV) 1 1 1 1 − − − − V V V V e e e e k k k k 1 1 1 1 − − − − s s s s 2 2 2 2 − − − − m m m m c c c c s s s s n n n n o o o o t t t t o o o o h h h h P P P P 2 2 2 2 V V V V e e e e k k k k o o i i t t a a R R 5 5 5 5 . . . . 3 3 3 3 3 3 3 3 5 5 5 5 . . . . 2 2 2 2 5 5 0 0 . . 1 1 1 1 5 5 9 9 . . 0 0 4 4 6 6 8 8 Energy (keV) Energy (keV) Figure 6: A simultaneous fit to the RXTE/Suzaku data in the 3.5 -- 9 keV range, consisting of an absorbed disk plus powerlaw, a relativistically broadened and a narrow Fe Ka line, and a 6.7 keV absorption line (left), plus an additional smeared edge (right). (Each model component is shown individually.) Such complexity is required for the Suzaku data, and is seen to give very good consistency with the RXTE data. Note that the data here are unfolded without reference to any model and only rely on the detector responses. paign, we first consider simple phenomenological fits consisting of an absorbed, exponentially cut- off broken powerlaw. Instead of adding a single broad gaussian, we also include a narrow gaussian. Additionally, we include a soft disk component. An example fit is shown in Fig 7. Although this fit is relatively successful, several points are worth noting. For this, and all of our fits, there are sharp, ≈ 5% residuals in the 0.7 -- 1.2 keV region, i.e., near absorption edge fea- tures. We fit absorption with tbnew, a modification of the model of [21], which includes detailed structure at these edges. Altering absorption abundances, however, did not remove these residuals. We are unsure currently whether these residuals are instrumental (e.g., contaminant unmodeled in the Suzaku spectral response) or physical features (e.g., unmodeled line absorption or emission). The additional disk component is both of high temperature and low amplitude. Formally, low am- plitude corresponds to a small inner disk radius4 Given the high fitted disk temperature, here the disk amplitude implies an inner disk radius of only ≈ 0.1GM/c2. A different (more heavily piled up, even with removal of data from the center of the image) Suzaku observation of Cyg X-1 re- cently has been described with a more complex Comptonization model consisting of different low and high energy Comptonized components ([10]). Perhaps the low normalization/high temperature disk component is indicating the need for continuum models with such added complexity. In Fig. 7 we also show the same spectra fit with the eqpair Comptonization model, allowing for reflection, an unComptonized disk component tied to the temperature of the seed photons, and a second disk component with temperature left free (here, ≈ 100 eV). Overall, this fit is somewhat data are in good agreement with the 20 -- 70 keV RXTE-HEXTE data, and do not alter the fits. Consideration of the HXD-GSO background is a complex issue, and we defer discussion of these data to a forthcoming work. 4Discussions of the inner disk radius not receding as a BHC enters the low/hard state often have shown ratio residuals of the model -- excluding the disk component -- in order to illustrate their points. It is the relative weakness of this ratio, however, that formally indicates a small inner disk radius, i.e., a non-recessed disk. 8 Suzaku Observations of Cyg X-1 Michael Nowak 8 8 8 8 8 − − − − − 0 0 0 0 0 1 1 1 1 1 9 9 9 9 9 − − − − − 0 0 0 0 0 1 1 1 1 1 ) ) ) ) ) 1 1 1 1 1 − − − − − s s s s s 2 2 2 2 2 − − − − − m m m m m c c c c c s s s s s g g g g g r r r r r e e e e e ( ( ( ( ( Fn Fn Fn Fn Fn 5 0 . 1 1 o i t a R 8 8 8 8 8 8 8 − − − − − − − 0 0 0 0 0 0 0 1 1 1 1 1 1 1 9 9 9 9 9 9 9 − − − − − − − 0 0 0 0 0 0 0 1 1 1 1 1 1 1 ) ) ) ) ) ) ) 1 1 1 1 1 1 1 − − − − − − − s s s s s s s 2 2 2 2 2 2 2 − − − − − − − m m m m m m m c c c c c c c s s s s s s s g g g g g g g r r r r r r r e e e e e e e ( ( ( ( ( ( ( Fn Fn Fn Fn Fn Fn Fn 5 0 . 1 1 o i t a R 1 10 100 1 10 100 Energy (keV) Energy (keV) Figure 7: Broad-band Cyg X-1 spectra that have been unfolded solely using the detector response matrices (i.e., without regard to the fit model). The Suzaku data from all available XIS have been summed and binned to HWHM of the detector, and we show the simultaneous RXTE-PCA and RXTE-HEXTE data. (Model data have been "unfolded" identically to the source data.) Left: absorbed, exponentially cutoff broken powerlaw, plus additional disk, and broad and narrow line components. All (absorbed) model components are also shown individually, and we also show the broken powerlaw with and without the hardening above 10 keV. Right: The eqpair Comptonization model fits, including additional unComptonized disk components (one tied to the seed photon temperature, and one free), broad and narrow lines, and reflection. We also show the additional (absorbed) disk and line components individually, the (absorbed) seed photon spectrum (highest amplitude thermal bump in the figure), and the Compton spectrum without reflection. better than the broken powerlaw model. However, as for the broken powerlaw fit, any soft excess (here dominated by the seed photons for Comptonization) is both of high temperature, and low amplitude. (To date, for these historically hard "low state" observations, we have been unable to find a satisfactory fit with low seed photon temperatures.) Note also that the fitted reflection fraction is ≈ 0.3, i.e., counter to any expectations from a hardness-reflection fraction anti-correlation. We have simultaneous Chandra-HETG and XMM-RGS data for one of our observations, thus, we are exploring whether the inclusion of narrow absorption and emission components (e.g., from a highly ionized wind) can fundamentally alter the broad-band properties suggested by the above models. References [1] M. Axelsson, L. Borgonovo, and S. Larsson, Evolution of the 0.01-25 Hz power spectral components in Cygnus X-1, A&A 438, 999 -- 1012 (2005). [2] M. Bałuci´nska-Church, M. J. Church, P. A. Charles, F. Nagase, J. LaSala, and R. Barnard, The distribution of X-ray dips with orbital phase in Cygnus X-1, MNRAS 311, 861 -- 868 (2000). [3] Paolo Coppi, PASP Conference Series, 161, 375, 1999. [4] John E. Davis, Event pileup in charge coupled devices, ApJ 562, 575 -- 582 (2001). [5] E. Gallo, R. P. Fender, and G. G. Pooley, A universal radio-X-ray correlation in low/hard state black hole binaries, MNRAS 344, 60 -- 72 (2003). 9 n n n n n n n n n n n n Suzaku Observations of Cyg X-1 Michael Nowak [6] T. Gleissner, J. Wilms, G. G. Pooley, M. A. Nowak, K. Pottschmidt, S. Markoff, S. Heinz, M. Klein-Wolt, R. P. Fender, and R. Staubert, Long term variability of Cyg X-1. III. radio-X-ray correlations, A&A 425, 1061 -- 1068 (2004). [7] T. Gleissner, J. Wilms, K. Pottschmidt, P. Uttley, M. A. Nowak, and R. Staubert, Long term variability of Cyg X-1. II. the rms-flux relation, A&A 414, 1091 -- 1104 (2004). [8] M. Hanke, J. Wilms, M. A. Nowak, K. Pottschmidt, N. S. Schulz, and J. C. Lee, Chandra X-ray spectroscopy of the focused wind in the Cygnus X-1 system I. the non-dip spectrum in the low/hard state, ApJ (2008), in press. [9] A. Ibragimov, J. Poutanen, M. Gilfanov, A. A. Zdziarski, and C. R. Shrader, Broad-band spectra of Cygnus X-1 and correlations between spectral characteristics, MNRAS 362, 1435 -- 1450 (2005). [10] K. Makishima, H. Takahashi, S. Yamada, C. Done, A. Kubota, T. Dotani, K. Ebisawa, T. Itoh, S. Kitamoto, H. Negoro, Y. Ueda, and K. Yamaoka, Suzaku results on Cygnus X-1 in the low/hard state, PASJ 60, 585 -- (2008). [11] S. Markoff, M. Nowak, and J. Wilms, Going with the flow: Can the base of jets subsume the role of compact accretion disk coronae?, ApJ 635, 1203 -- 1216 (2005). [12] J. M. Miller, A. C. Fabian, R. Wijnands, R. A. Remillard, P. Wojdowski, N. S. Schulz, T. Di Matteo, H. L. Marshall, C. R. Canizares, D. Pooley, and W. H. G. Lewin, Resolving the composite Fe Ka Emission line in the galactic black hole Cygnus X-1 with Chandra, ApJ 578, 348 -- 356 (2002). [13] Michael A. Nowak, Brian A. Vaughan, Jörn Wilms, James Dove, and Mitchell C. Begelman, RXTE observations of Cygnus X -- 1: II. Timing analysis, ApJ 510, 874 -- 891 (1999). [14] Michael A. Nowak, Jörn Wilms, Sebastian Heinz, Guy Pooley, Katja Pottschmidt, and Stephane Corbel, Is the 'IR coincidence' just that?, ApJ 626, 1006 -- 1014 (2005). [15] K. Pottschmidt, J. Wilms, M. A. Nowak, W. A. Heindl, D. M. Smith, and R. Staubert, Temporal evolution of X-ray lags in Cygnus X-1, A&A 357, L17 -- L20 (2000). [16] K. Pottschmidt, J. Wilms, M. A. Nowak, G. G. Pooley, T. Gleissner, W. A. Heindl, D. M. Smith, R. Remillard, and R. Staubert, Long term variability of Cyg X-1 (1998 to 2001) I. systematic spectral-temporal correlations in the hard state, A&A 407, 1039 -- 1058 (2003). [17] P. Predehl and J. H. M. M. Schmitt, X-raying the interstellar medium: ROSAT observations of dust scattering halos, A&A 293, 889 -- 905 (1995). [18] R. A. Remillard and J. E. McClintock, X-ray properties of black-hole binaries, Annual Review of Astronomy and Astrophysics 44, 49 -- 92 (2006). [19] Y. Uchiyama, Y. Maeda, M. Ebara, R. Fujimoto, Y. Ishisaki, M. Ishida, R. Iizuka, M. Ushio, H. Inoue, S. Okada, H. Mori, and M. Ozaki, Restoring the Suzaku source position accuracy and point-spread function, PASJ 60, 35 -- 42 (2008). [20] P. Uttley, I. M. McHardy, and S. Vaughan, Non-linear X-ray variability in X-ray binaries and active galaxies, MNRAS 359, 345 -- 362 (2005). [21] J. Wilms, A. Allen, and R. McCray, On the absorption of X-rays in the interstellar medium, ApJ 542, 914 -- 924 (2000). [22] Jörn Wilms, Michael Nowak, Katja Pottschmidt, Guy G. Pooley, and Sonia Fritz, Long term variability of Cygnus X-1. IV. Spectral evolution 1999-2004, A&A 447, 245 -- 261 (2006). [23] Andrzej A. Zdziarski, Piotr Lubi´nski, and David A. Smith, Correlation between Compton reflection and X-ray slope in Seyferts and X-ray binaries, MNRAS 303, L11 -- L15 (1999). 10
astro-ph/9810347
1
9810
1998-10-21T19:01:34
The ROSAT Deep Survey IV. A distant lensing cluster of galaxies with a bright arc
[ "astro-ph" ]
An unusual double-lobed extended X-ray source (RX J105343+5735) is detected in the ROSAT ultra-deep HRI image of the Lockman Hole. The angular size of the source is 1.7 X 0.7 arcmin^2 and its X-ray flux is 2 X 10^-14 erg cm^-2 s^-1. R-band imaging from the Keck telescope revealed a marginal excess of galaxies brighter than R=24.5, but Keck LRIS spectroscopy of 24 objects around the X-ray centroid did not yield a significant number of concordant redshifts. The brightest galaxy close to the centre of the eastern emission peak appears to be a gravitationally lensed arc at z=2.570, suggesting that the X-ray object is associated with the lens, most likely a cluster of galaxies. Based on a comparison of lensing surface mass density, X-ray luminosity, morphology and galaxy magnitudes with clusters of known distance, we argue that RX J105343+5735 is a cluster at a redshift around 1. Future X-ray, ground-based optical/NIR and high resolution HST observations of the system will be able to clarify the nature of the object.
astro-ph
astro-ph
A&A manuscript no. (will be inserted by hand later) Your thesaurus codes are: 06 (11.09.3; 12.04.1; 12.07.1; 13.25.2) ASTRONOMY AND ASTROPHYSICS 30.1.2018 8 9 9 1 t c O 1 2 1 v 7 4 3 0 1 8 9 / h p - o r t s a : v i X r a The ROSAT Deep Survey IV. A distant lensing cluster of galaxies with a bright arc G. Hasinger1, R. Giacconi2, J.E. Gunn3, I. Lehmann1, M. Schmidt4, D.P. Schneider5, J. Trumper6, J. Wambsganss1, D. Woods1, and G. Zamorani7 1 Astrophysikalisches Institut Potsdam, An der Sternwarte 16, D-14482 Potsdam 2 European Southern Observatory, Karl-Schwarzschild-Str. 2, 85740 Garching bei Munchen, Germany 3 Princeton University Observatory, Princeton, NJ 08540, USA 4 California Institute of Technology, Pasadena, CA 91125, USA 5 Pennsylvania State University, University Park, PA 16802, USA 6 Max -- Planck -- Institut fur extraterrestrische Physik, Giessenbachstr. 1, 85740 Garching bei Munchen, Germany 7 Osservatorio Astronomico, via Zamboni 33, I-40126 Bologna, Italy Received 15 September 1998; accepted 21 October 1998 Abstract. An unusual double-lobed extended X-ray source (RX J105343+5735) is detected in the ROSAT ultra-deep HRI image of the Lockman Hole. The angular size of the source is 1.7 × 0.7 arcmin2 and its X-ray flux is 2 × 10−14 erg cm−2 s−1. R-band imaging from the Keck telescope revealed a marginal excess of galaxies brighter than R=24.5, but Keck LRIS spectroscopy of 24 objects around the X-ray centroid did not yield a significant num- ber of concordant redshifts. The brightest galaxy close to the centre of the eastern emission peak appears to be a gravitationally lensed arc at z=2.570, suggesting that the X-ray object is associated with the lens, most likely a clus- ter of galaxies. Based on a comparison of lensing surface mass density, X-ray luminosity, morphology and galaxy magnitudes with clusters of known distance, we argue that RX J105343+5735 is a cluster at a redshift around 1. Fu- ture X-ray, ground-based optical/NIR and high resolution HST observations of the system will be able to clarify the nature of the object. Key words: Galaxies: intergalactic medium; Cosmology: dark matter, gravitational lensing; X-rays: galaxies 1. Introduction and the evolution of structures in the universe. Surveys at the faintest X-ray fluxes (e.g. Hasinger et al. 1998, here- after paper I) predominantly find active galactic nuclei (Schmidt et al. 1998, hereafter paper II) often in conjunc- tion with starburst regions, which might also be responsi- ble for the absorption observed in the sources of the X-ray background (Fabian et al. 1998). Clusters of galaxies are the second most abundant class of objects in deep X-ray surveys. They are the largest bound structures in the universe, composed of dark matter condensations, hot X-ray emitting gas and galaxies. X-ray radiation is an efficient means to select clusters of galaxies, because the X-ray flux is proportional to the square of the electron density. The highest observed redshifts of X- ray selected clusters to date are ∼0.8-0.9 (Henry et al., 1997; Rosati et al., 1998) and extended X-ray emission of clusters selected by other techniques has been reported out to z=1.27 (Stanford et al., 1997; Hattori et al., 1997). In this letter we report on the discovery of a faint double-lobed X-ray source, RX J105343+5735, detected in the ROSAT ultra-deep HRI image of the Lockman Hole (paper I). In section 2 and 3 we present the X-ray and optical observations, and in section 4 we discuss the properties of the object in the context of what we know about other galaxy clusters. We use a Hubble constant H0 = 50 km s−1 Mpc−1 and a deceleration parameter q0 = 0.5 throughout this paper. X-ray surveys efficiently select cosmic X-ray sources at dis- tances ranging from our immediate neighbourhood to the edge of the observable universe. By studying their proper- ties as a function of redshift we can probe the formation Send offprint requests to: G. Hasinger 2. X-ray observations The ROSAT Deep Survey (RDS) project is described in paper I. The RDS consists of a series of deep pointings of 207 ksec with the ROSAT PSPC and 1.31 Msec with 2 Bright arc cluster Table 1. X-ray sources detected near RX J105343+5735 RA (2000) Source 10h53m43.4s A+B 10h53m46.6s A 10h53m40.1s B 10h53m29.5s C 10h53m48.2s D a 10−14 erg cm−2 s−1 DEC (2000) 57d35m21s 57d35m17s 57d35m25s 57d35m38s 57d33m55s Fluxa 2 Type cluster 0.18 0.17 group QSO the ROSAT HRI in the direction of the "Lockman Hole", a line of sight with exceptionally low HI column density. A catalogue of 50 X-ray sources with fluxes brighter than 0.55 × 10−14 erg cm−2 s−1 has been published in paper I; the spectroscopic optical identifications are presented in paper II. The ultra-deep HRI pointing covers a solid angle of 0.126 deg2 inside the RDS. In the north-east quadrant of the HRI image (see paper I) is RX J105343+5735, a clearly extended X-ray source with a double-lobed structure. In Fig.1 we show X-ray contours of this source and its surroundings, superposed on an optical image. The extension of the source is about 1.7 and 0.6 arcmin in the E-W and N-S direction, respec- tively. We denote the Eastern lobe (A) and the Western lobe (B); both components are significantly extended. Two more X-ray sources are detected at a significance of ∼ 4σ: a possible group of galaxies at z=0.7 (C) and a QSO at z=2.572 (D). Details for the detected X-ray sources are given in Table 1. None of them appears in the previously published source list, because either their off-axis angle is too large (A, B) or their X-ray flux too low (C, D) for the selection criteria applied in paper I. The extended source is detected in the PSPC point- ing at an off-axis angle of ∼ 19 arcmin and therefore is not well resolved. With about 590 net photons detected in the PSPC a coarse spectrum could be obtained. We fixed the NH value to 1020 cm−2, consistent with the in- dependent determination of NH from a fit to the summed spectrum of all resolved sources in the Lockman Hole (Hasinger et al. 1993). Using a Raymond-Smith model with solar abundances (Z=1) we obtain acceptable fits (see χ2 in Table 3) to the data assuming various redshifts for the source. The results are not significantly different for Z=0.3. The flux observed in the 0.5-2.0 keV band is ∼ 2 × 10−14 erg cm−2 s−1 for all these fits. The derived temperatures range between 1.8 and 3.5 keV, the lumi- nosities (0.5-2 keV) between 1043 and 2.4 × 1044 erg s−1 for redshifts between 0.3 and 1.2 (see Table 3). 3. Optical observations An R-band image of the field of RX J105343+5735 was obtained on 1996 April 13 with the Low Resolution Imaging Spectrometer (LRIS; Oke et al. 1995) on the Keck I telescope. The exposure time was 300s and the seeing (FWHM) was 0.75". The image scale on the back- 9 4 . 3 1 0 1 - . 6 1 1 8 7 5 - Fig. 1. X-ray contours of a region about 3.6 × 3.6 arcmin2 in the ROSAT HRI ultra-deep pointing of the Lockman Hole, superposed on a Keck R-band exposure. Numbers indicate the objects for which we have obtained spectra. The insert in the upper right is a 20 × 20 arcsec2 zoom of the image close to the centre of lobe A. An arc-like feature (#2) is visible here. Fig. 2. Keck LRIS spectrum of the arc (#2 in Fig.1). The wavelengths of prominent UV metal absorption lines, red- shifted to z=2.570 are indicated. "S" denotes the position of a strong sky line. The spectrum is typical for a high redshift starforming galaxy. illuminated 2048 × 2048 Tektronix CCD is 0.215" pixel−1. The CCD image was de-biased and flatfielded in the stan- dard fashion using IRAF routines. A flatfield was gener- ated by performing a median of 8 different R images, all of which lacked bright stars and were taken over the course of the night. We were able to flatten the resultant images to better than ∼ 1%. Detection and photometric analysis of the galaxies and stars present in the program field was done using the SExtractor package (Bertin and Arnouts 1996). The faintest detections (3σ) have R ≈ 25.5. Bright arc cluster 3 Fig. 1 shows a section of the LRIS image underlying the HRI X-ray contours of the field. Close to the centroid of lobe A is an arc-like feature (#2). It has a length of ∼ 7" along the major axis and is not resolved along the minor axis. The arc has an integrated magnitude of R=21.4, a maximum surface brightness of µR = 22.9 mag arcsec−2 and some asymmetry along the major axis. There is diffuse emission around it with a surface brightness of µR ≈ 25.5 mag arcsec−2 as well as some faint galaxies. Spectra of 24 objects in this field were acquired with a multislit mask using LRIS on Keck II on 1998 March 19. The conditions were photometric, the seeing was ∼ 1" and the exposure time 3600s. The slits were all 1.4" wide and some of them tilted to maximize the number of galax- ies observed. The 300 line/mm grating produced spectra in the range 3800 to 8200 A with a resolution of ∼ 15A. Spectra were processed using standard MIDAS procedures for bias subtraction, flat-field correction, optimal extrac- tion and wavelength calibration. Relative flux correction and atmospheric absorption correction was done using the spectrophotometric standard BD +26 2606. Table 2 gives details about the observed galaxies using the numbering scheme given in Fig. 1 (#19-24 are outside the area covered by Fig. 1): coordinates, R magnitudes, redshifts and absolute R magnitudes calculated using K- corrections for the LRIS R-filter computed assuming the spectral energy distribution for a giant elliptical galaxy for objects with z < 1 and an assumed K-correction of 0 for the two high-redshift objects (#1 and #2). Apart from four concordant redshifts around 0.7 (#13,16,17,18), three of which are coincident with a discrete X-ray source (C in Fig. 1), possibly a group of galaxies, three galaxies at z=0.308 (#7,11,19), scattered throughout the image, and two galaxies at z=0.782 (#5,8), there is no obvious clustering of redshifts in this sample. The spectrum of the brightest object (#2) close to the centroid of lobe A is displayed in Fig. 2. It shows prominent absorption features of Lyα and low-ionization metal lines typical of the rest-frame UV emission of star- burst galaxies which are characteristic of a young stel- lar population. The spectrum is very similar to those of the UV-dropout galaxies detected by Steidel et al. (1996), which gives us confidence in the redshift determination at z = 2.570 ± 0.002. Its absolute magnitude (MR = −25.1) indicates significant magnification and its spectrum, ap- parent magnitude (R=21.4) and redshift (z=2.570) are all very similar to the gravitationally lensed galaxy at R=21.2, z=2.515 in the cluster A2218 (Ebbels et al, 1996), supporting the gravitational arc interpretation. Interest- ingly, the QSO (C) at z=2.572 ± 0.002 is at a projected distance of only 0.6 Mpc from the arc galaxy. 4. Discussion The morphology, observed flux and lensing action of RX J105343+5735 strongly suggests that the X-ray source is # ∆R.A. ∆Dec. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 32 31 31 29 12 9 1 -9 -15 -24 -28 -45 -59 -76 -91 -108 -115 -119 -128 -138 -165 -167 -173 -278 -91 -11 1 -11 -6 -4 10 20 21 30 30 43 52 35 7 13 14 17 109 126 136 92 19 44 R 22.6 21.4 22.1 22.9 22.9 23.1 19.1 22.2 22.7 22.4 22.6 21.5 22.6 21.5 22.3 22.1 21.2 22.3 20.1 22.0 20.1 20.8 21.6 19.3 z 2.567 2.570 0.736 0.782 0.543: 0.308 0.782 0.581 0.097 0.308 0.045 0.697 0.602: 0.521 0.695 0.700 0.695 0.308 0.383 0.039: 0.595 0.815: 0.352 MR -23.9 -25.1 -23.4 -22.8 -20.8 -22.8 -23.5 -21.6 -16.6 -19.2 -15.7 -22.6 -22.9 -21.5 -23.1 -24.0 -22.9 -21.8 -20.6 -16.8 -23.6 -24.3 -23.0 Table 2. Parameters of galaxies with spectroscopy. #: number of object as indicated in Fig. 1; ∆R.A., ∆Dec.: distance ["] relative to the cluster centroid centroid at R.A. 10h53m43.4s, Dec. 57d35m21s (Epoch 2000); R: approximate R magnitude; z: redshift; MR: absolute R magnitude. Table 3. Cluster parameters for assumed redshifts red Redshift kT χ2 LX ML RcD 0.3 0.7 1.2 1.8 ± 0.5 2.6 ± 1.1 3.5 ± 1.1 keV 0.49 0.09 0.11 18.8 0.81 0.66 0.22 22.2 0.89 2.40 0.40 24.9 1044 erg/s 1014 M⊙ mag associated with a moderately luminous, distant cluster of galaxies. To quantify a possible excess of the projected density of galaxies compared to the field objects, we use the estimator discussed by Lidman and Peterson (1996, eqn. 1) which gives a value for the cluster contrast, σcl. Using the X-ray centroid as assumed cluster centre, we find a slight excess (σcl > 2.0) of galaxies with R=23.5- 24.5 for a 2' diameter aperture. The group observed at z=0.7 (galaxies #16, 17 and 18) shows a significant excess (σcl ≈ 3) of brighter galaxies (R < 23.5). No obvious enhancement of galaxies is observed near the QSO (#1). Since the redshift of the lens is unknown, we have pro- duced a series of simple models for various values of the lens redshift (see Table 3). For all considered redshifts the object fits onto the empirical luminosity-temperature re- lation for clusters of galaxies (see e.g. Ebeling et al., 1996). 4 Bright arc cluster We estimate that the lensed arc has a radius of curvature ΘE ≈ 7.5 ± 2". We can estimate the lensing mass inside this radius by assuming the arc radius corresponds to the Einstein ring radius for a mass ML. Assuming different redshifts we obtain the lensing masses in Table 3. Note that this mass corresponds to the core mass of one lobe only. In Table 3 we also give an estimate of the expected R magnitude for the brightest cluster galaxy RcD, assuming MR = −23 based on the compilation of Hoessel & Schnei- der (1985). We now discuss what would be the properties of the extended X-ray object for various assumed redshifts: z ≈ 0.3: Three galaxies (#7,11,19) at this redshift are found, including the 19th magnitude galaxy (#7) close to the symmetry axis. If the X-rays originate at this red- shift, the X-ray luminosity would correspond to a group of galaxies. The observed double structure could be sim- ilar to some nearby groups, which can have complicated morphologies (Mulchaey et al. 1996). In this case, object #7 could be the brightest galaxy of the group, but no other galaxy of similar magnitude is seen around it. The detection of an arc is an argument against a group in- terpretation. The lensing surface mass density would be a factor of ∼ 10 higher than that typically estimated for groups (Mulchaey et al., 1996). z ≈ 0.7: Four galaxies (#13,16,17,18) are detected at this redshift, but three of them are associated with source C, a very faint discrete X-ray source, possibly a group of galaxies (LX = 5 × 1042 erg/s). The X-ray luminos- ity would be consistent with a moderately rich cluster of galaxies. The elongated morphology is similar to other high-redshift clusters (Henry et al., 1997) and there is an interesting linear arrangement between A, B and C, which could indicate that all three mass condensations are in the same filamentary structure at z=0.7. But in this case A+B would have a factor of 10 higher X-ray luminosity than C, while C has clearly visible cluster galaxies in the right absolute magnitude range which are not seen in A+B. Therefore A+B would have to be a "dark" cluster with- out bright galaxies (Tucker et al., 1995). While this is an interesting possibility, we regard it as unlikely. z > 0.7: In this case all the measured properties are consistent with a normal, moderately rich cluster at very high redshift. The object would be similar to the lensing cluster 2016+112 at z=1 (Schneider et al. 1986; Hattori et al. 1997). In this scenario the brightest cluster galax- ies might just be visible in our R-band exposure (see Ta- ble 3) and we would not have a chance of detecting any cluster member within our spectroscopic limit of R ≈ 23. The lensing surface mass density falls among those for other clusters of galaxies. In this case the object would be one of the highest redshift clusters with X-ray emission. Assuming no evolution for the Rosati et al. (1998) clus- ter luminosity function, we would expect 0.15 clusters in the redshift range z=1-1.5 in our survey volume. We re- gard this as the most likely possibility which can be easily tested with future observations. Upcoming deep surveys with AXAF and XMM should have no problem detecting and studying a number of similar objects at high redshift. In conclusion, RX J105343+5735 is an intriguing ob- ject at any of the redshifts we can plausibly assign to the lens and well worth follow-up studies (e.g. NIR imaging, high-resolution imaging with HST, X-ray spectroscopy). In particular, it is the first detection of a gravitational arc which is optically brighter than any of the components of the lens. The detection of a QSO at the same redshift as the arc indicates that an enhanced density of background galaxies in this direction might have increased the likeli- hood to observe a lensed object. Acknowledgements. We thank an anonymous referee for help- ful comments. The ROSAT project is supported by the Bundesministerium fur Bildung, Forschung und Wissenschaft (BMBF), by the National Aeronautics and Space Administra- tion (NASA), and the Science and Engineering Research Coun- cil (SERC). The W. M. Keck Observatory is operated as a sci- entific partnership between the California Institute of Technol- ogy, the University of California, and the National Aeronautics and Space Administration. It was made possible by the gen- erous financial support of the W. M. Keck Foundation. This work was supported by DLR grant 50 OR 9403 5 (G.H., I.L.), National Science Foundation grant AST-95-09919 (D.P.S.) and ASI grant ARS-96-70 (G.Z.). References Bertin E. & Arnouts S. 1996, A&AS 117, 393 Ebbels T.M.D., Le Bourgne J.-F., Pello R., et al. 1996, MNRAS 281, L75 Ebeling H., Voges W., Bohringer H., et al. 1996, MNRAS 281, 799 Fabian A.C., Barcons X., Almaini O. & Iwasawa K. 1998, MN- RAS 297, L11 Hattori M., Ikebe Y., Asaoka I., et al. 1997, Nat 388, 146 Hasinger G., Burg R., Giacconi R., et al. 1993, A&A 275, 1 Hasinger G., Burg R., Giacconi R., et al. 1998, A&A 329, 482 (paper I) Henry J.P., Gioia I.M., Mullios C.R., et al. 1997, AJ 114, 1293 Hoessel J.G. & Schneider D.P. 1985, AJ 90, 1648 Lidman C.E. & Peterson B.A. 1996, AJ 112, 2454 Mulchaey J.S., Davis D.S., Mushotzky R.F., Burstein D. 1996, ApJ 456, 80 Oke J.B., Cohen J.G., Carr M., et al. 1995, PASP 107, 375 Rosati P., Della Ceca R., Norman C. & Giacconi R. 1998, ApJ 492, L21 Schneider D.P., Gunn J.E., Turner E.L., et al. 1986, AJ 91, 991 Schmidt M., Hasinger G., Gunn, J., et al. 1998, A&A, 329, 495 (paper II) Stanford S.A., Elston R., Eisenhardt P.R. et al. 1997, AJ 114, 2232 Steidel C.C., Giavalisco M., Pettini M., Dickinson M. & Adel- berger K. 1996, ApJ 462, L17 Tucker W.H., Tananbaum H. & Remillard R.A. 1995, ApJ 444, 532 This article was processed by the author using Springer-Verlag LaTEX A&A style file L-AA version 3.