<<

The Photoevaporation of Dwarf Galaxies during Reionization

The Harvard community has made this article openly available. Please share how this access benefits you. Your story matters

Citation Barkana, Rennan, and Abraham Loeb. 1999. “The Photoevaporation of Dwarf Galaxies during Reionization.” The Astrophysical Journal 523 (1): 54–65. https://doi.org/10.1086/307724.

Citable link http://nrs.harvard.edu/urn-3:HUL.InstRepos:41393254

Terms of Use This article was downloaded from Harvard University’s DASH repository, and is made available under the terms and conditions applicable to Other Posted Material, as set forth at http:// nrs.harvard.edu/urn-3:HUL.InstRepos:dash.current.terms-of- use#LAA THE ASTROPHYSICAL JOURNAL, 523:54È65, 1999 September 20 ( 1999. The American Astronomical Society. All rights reserved. Printed in U.S.A.

THE PHOTOEVAPORATION OF DWARF GALAXIES DURING REIONIZATION RENNAN BARKANA Institute for Advanced Study, Olden Lane, Princeton, NJ 08540; barkana=ias.edu AND ABRAHAM LOEB Astronomy Department, Harvard University, 60 Garden Street, Cambridge, MA 02138; aloeb=cfa.harvard.edu Received 1999 January 12; accepted 1999 May 5

ABSTRACT During the period of reionization, the universe was Ðlled with a cosmological background of ionizing radiation. By that time a signiÐcant fraction of the cosmic gas had already been incorporated into col- lapsed galactic halos with virial temperatures[104 K that were unable to cool efficiently. We show that photoionization of this gas by the fresh cosmic UV background boiled the gas out of the gravitational potential wells of its host halos. We calculate the photoionization heating of gas inside spherically sym- metric dark matter halos and assume that gas that is heated above its virial temperature is expelled. In popular cold dark matter models, the Press-Schechter halo abundance implies that D50%È90% of the collapsed gas was evaporated at reionization. The gas originated from halos below a threshold circular velocity of D10È15 km s~1. The resulting outÑows from the dwarf galaxy population at redshifts z \ 5È10 a†ected the metallicity and the thermal and hydrodynamic states of the surrounding inter- galactic medium. Our results suggest that stellar systems with a velocity dispersion[10 km s~1, such as globular clusters or the dwarf spheroidal galaxies of the Local Group, did not form directly through cosmological collapse at high redshifts. Subject headings: cosmology: theory È galaxies: formation È galaxies: halos È radiative transfer

1. INTRODUCTION The lack of a Gunn-Peterson trough and the detection of The formation of galaxies is one of the most important, Lya emission lines from sources out to redshifts z \ 5.6 yet unsolved, problems in cosmology. The properties of (Weymann et al. 1998; Dey et al. 1998; Spinrad et al. 1998; galactic dark matter halos are shaped by gravity alone and Hu, Cowie, & McMahon 1998) demonstrates that reioniza- have been rigorously parameterized in hierarchical cold tion due to the Ðrst generation of sources must have dark matter (CDM) cosmologies (e.g., Navarro, Frenk, & occurred at yet higher redshifts; otherwise, the damping White 1997, hereafter NFW). However, the complex pro- wing of Lya absorption by the neutral IGM would have cesses involving gas dynamics, chemistry and ionization, eliminated the Lya line in the observed spectrum of these and cooling and heating, which are responsible for the for- sources (Miralda-Escude 1998). Popular CDM models mation of from the baryons inside these halos, have predict that most of the intergalactic was ionized still not been fully explored theoretically. at a redshift8 [ z [ 15 (Gnedin & Ostriker 1997; Haiman Recent theoretical investigations of early structure forma- & Loeb 1998, 1998c). The end of the reionization phase tion in CDM models have led to a plausible picture of how transition resulted in the emergence of an intense UV back- the formation of the Ðrst cosmic structures leads to reioni- ground that Ðlled the universe and heated the IGM to tem- zation of the intergalactic medium (IGM). The bottom-up peratures of D1È2 ] 104 K (Haiman & Loeb 1999; hierarchy of CDM cosmologies implies that the Ðrst Miralda-Escude , Haehnelt, & Rees 1999). After ionizing the gaseous objects to form in the universe have a low mass, just rareÐed IGM in the voids and Ðlaments on large scales, the 4 above the cosmological Jeans mass of D10 M_ (see, e.g., cosmic UV background penetrated the denser regions Haiman, Thoul, & Loeb 1996b and references therein). The associated with the virialized gaseous halos of the Ðrst gen- virial temperature of these gas clouds is only a few hundred eration of objects. Since a major fraction of the collapsed K, and so their metal-poor primordial gas can cool only as gas had been incorporated by that time into halos with a 4 a result of the formation of molecular hydrogen, H2. virial temperature[10 K, photoionization heating by the However,H2 molecules are fragile and were easily photo- cosmic UV background could have evaporated much of this dissociated throughout the universe by trace amounts of gas back into the IGM. No such feedback was possible at starlight (Stecher & Williams 1967; Haiman, Rees, & Loeb earlier times, since the formation of internal UV sources 1996a) that were well below the level required for complete was suppressed by the lack of efficient cooling inside most reionization of the IGM. Following the prompt destruction of these objects. of their molecular hydrogen, the early low-mass objects The gas reservoir of dwarf galaxies with virial tem- maintained virialized gaseous halos that were unable to peratures[104 K (or equivalently a one-dimensional cool or fragment into stars. Most of the stars responsible for velocity dispersion[10 km s~1) could not be immediately the reionization of the universe formed in more massive replenished. The suppression of dwarf galaxy formation at 4 galaxies, with virial temperaturesTvir Z 10 K, where z [ 2 has been investigated both analytically (Rees 1986; cooling due to atomic transitions was possible. The corre- Efstathiou 1992) and with numerical simulations (Thoul & 8 sponding mass of these objects at z D 10 was D10 M_, Weinberg 1996; Quinn, Katz, & Efstathiou 1996; Wein- typical of dwarf galaxies. berg, Hernquist, & Katz 1997; Navarro & Steinmetz 1997). 54 PHOTOEVAPORATION OF DWARF GALAXIES 55 \ The dwarf galaxies that were prevented from forming after wherex r/rvir and c depends ondc for a given mass M. \ reionization could have eventually collected gas at z 1È2, We include the dependence of halo proÐles on)0 and )", when the UV background Ñux declined sufficiently (Babul the current contributions to ) from nonrelativistic matter & Rees 1992; Kepner, Babul, & Spergel 1997). The reverse and a cosmological constant, respectively (see Appendix A process during the much earlier reionization epoch has not for complete details). been addressed in the literature. (However, note that the Although the NFW proÐle provides a good approx- photoevaporation of gaseous halos was considered by imation to halo proÐles, there are indications that halos Bond, Szalay, & Silk (1988) as a model for Lya absorbers at may actually develop a core (e.g., Burkert 1995; Kravtsov et lower redshifts z D 4.) al. 1998; see, however, Moore et al. 1998). In order to In this paper we focus on the reverse process by which examine the sensitivity of the results to model assumptions, gas that had already settled into virialized halos by the time we consider several di†erent gas and dark matter proÐles, of reionization was evaporated back into the IGM because keeping the total gas fraction in the halo equal to the of the cosmic UV background that emerged Ðrst at that cosmological baryon fraction. The simplest case we con- epoch. The basic ingredients of our model are presented in sider is an equal NFW proÐle for the gas and the dark ° 2. In order to ascertain the importance of a self-shielded matter. In order to include a core, instead of the NFW gas core, we include a realistic, centrally concentrated dark proÐle of equation (3), we also consider the density proÐle of halo proÐle and also incorporate radiative transfer. Gener- the form Ðt by Burkert (1995) to dwarf galaxies, ally we Ðnd that self-shielding has a small e†ect on the total 3H2 ) d amount of evaporated gas, since only a minor fraction of o(r) \ 0 (1 ] z)3 0 c , (4) the gas halo is contained within the central core. Our 8nG )(z) (1 ] bx)[1 ] (bx)2] numerical results are described in ° 3. In particular, we show where b is the inverse core radius, and we setd by requiring the conditions in the highest mass halo that can be dis- c the mean overdensity to equal the appropriate value,* , in rupted at reionization. We also use the Press & Schechter c each cosmology (see Appendix A). We also consider two (1974) prescription for halo abundance to calculate the frac- cases where the dark matter follows an NFW proÐle but the tion of gas in the universe that undergoes the process of gas is in hydrostatic equilibrium with its density proÐle photoevaporation. Our versatile semianalytic approach has determined by its temperature distribution. In one case, we the advantage of being able to yield the dependence of the assume the gas is isothermal at the halo virial temperature, results on a wide range of reionization histories and cosmo- given by logical parameters. Clearly, the Ðnal state of the gas halo depends on its dynamical evolution during its photoevapo- kV 2 T \ c ration. We adopt a rough criterion for the evaporation of vir 2k gas based on its initial interaction with the ionizing back- B ground. The precision of our results could be tested in spe- k r 2 ) * 1 ] z 3 \ 36,100 A vir B 0 c A B K , (5) ciÐc cases by future numerical simulations. In ° 4 we discuss 0.6m h~1 kpc )(z) 200 10 the potential implications of our results for the state of the p IGM and for the early history of low-mass galaxies in the where k is the mean molecular weight as determined by local universe. Finally, we summarize our main conclusions ionization equilibrium, andmp is the proton mass. The in ° 5. spherical collapse simulations of Haiman et al. (1996b) Ðnd a postshock gas temperature of roughly twice the value 2. A MODEL FOR HALOS AT REIONIZATION given by equation (5), so we also compare with the result of \ We consider gas situated in a virialized dark matter halo. setting T 2Tvir. In the second case, we let the gas cool for We adopt the prescription for obtaining the density proÐles a time equal to the Hubble time at the redshift of interest, z. Gas above 104 K cools rapidly due to atomic cooling until of dark matter halos at various redshifts from the Appendix 4 of NFW, modiÐed to include the variation of the collapse it reaches a temperature near 10 K, where the cooling time rapidly diverges. In this case, hydrostatic equilibrium yields overdensity*c. Thus, a halo of mass M at redshift z is characterized by a virial radius, a highly compact gas cloud when the halo virial tem- perature is greater than 104 K. In reality, of course, a frac- M 1@3 ) * ~1@3 tion of the gas may fragment and form stars in these halos. r \ 0.756A B C 0 c D vir 108 h~1 M )(z) 200 However, this caveat hardly a†ects our results since only a _ small fraction of the gas which evaporates is contained in 1 ] z ~1 halos withT [ 104 K. Throughout most of our sub- ] A B h~1 kpc , (1) vir 10 sequent discussion we consider the simple case of identical NFW proÐles for both the dark matter and the gas, unless or a corresponding circular velocity, indicated otherwise. 1@2 We assume a mass fraction of Y \ 0.24, and \ AGMB Vc include it in the calculation of the ionization equilibrium rvir state of the gas as well as its cooling and heating (see, e.g., r ) * 1@2 1 ] z 3@2 Katz, Weinberg, & Hernquist 1996). We adopt the various \ 31.6A vir BC 0 c D A B km s~1 . (2) reaction and cooling rates from the literature, including the h~1 kpc )(z) 200 10 rates for collisional excitation and dielectronic recombi- The density proÐle of the halo is given by nation from Black (1981); the recombination rates from Verner & Ferland (1996), and the recombination cooling 3H2 ) d o(r) \ 0 (1 ] z)3 0 c , (3) rates from Ferland et al. (1992) with a Ðtting formula by J. 8nG )(z) cx(1 ] cx)2 Miralda-Escude (1998, private communication). Collisional 56 BARKANA & LOEB Vol. 523 ionization rates are adopted from Voronov (1997), with the with time (e.g., Haiman & Loeb 1998, 1999; Miralda- corresponding cooling rate for each atomic species given by Escude et al. 1999). The evolution of this process depends its ionization rate multiplied by its ionization potential. We on the characteristic clustering scale of ionizing sources and also include cooling by Bremsstrahlung emission with a their correlation with the inhomogeneities of the IGM. In Gaunt factor from Spitzer & Hart (1971) and by Compton particular, the process takes more time if the sources are scattering o† the microwave background (e.g., Shapiro & typically embedded in dense regions of the neutral IGM Kang 1987). that need to be ionized Ðrst before their radiation shines on In assessing the e†ect of reionization, we assume for sim- the rest of the IGM. However, in our analysis we do not plicity a sudden turn-on of an external radiation Ðeld with a need to consider these complications, since the total frac- speciÐc intensity per unit frequency, l: tion of evaporated gas in bound halos depends primarily on \ ~21 ~a ~2 ~1 ~1 ~1 the maximum intensity achieved at the end of the reioniza- Il,0 10 I21(z)(l/lL) ergs cm s sr Hz , tion epoch. (6) In computing the e†ect of the background radiation, we include self-shielding of the gas, which is important at the wherelL is the Lyman limit frequency. Our treatment of the high densities obtained in the core of high-redshift halos. response of the cloud to this radiation, as outlined below, is For this purpose, we include radiative transfer through the not expected to yield di†erent results with a more gradual halo gas and photoionization by the resulting anisotropic increase of the intensity with cosmic time. The external radiation Ðeld in the calculation of the ionization equi- intensityI21(z) is responsible for the reionization of the librium. We also include the fact that the ionizing spectrum IGM, and so we normalize it to have a Ðxed number of becomes harder at inner radii, since the outer gas layers ionizing per baryon in the universe. We deÐne the preferentially block photons with energies just above the ionizing density as Lyman limit. We neglect self-shielding due to helium atoms. Appendix B summarizes our simpliÐed treatment of the = 4nI p (l) \ P l,0 H I radiative transfer equations. nc dl , (7) lL hlc pH I(lL) Once the gas is heated throughout the halo, some frac- where the photoionization efficiency is weighted by the tion of it acquires a sufficiently high temperature that it I becomes unbound. This gas expands due to the resulting photoionization cross section of H ,pH I(l), above the Lyman limit. The mean baryon number density is pressure gradient and eventually evaporates back to the IGM. The pressure gradient force (per unit volume) 1 ] z 3 ) h2 k +(T o/k) competes with the gravitational force of n \ 2.25 ] 10~4A B A b B cm~1 . (8) B 2 b 10 0.02 oGM/r . Because of the density gradient, the ratio between the pressure force and the gravitational force is roughly the Throughout the paper we refer to proper densities rather ratio between the thermal energyDkB T and the gravita- than comoving densities. As our standard case we assume a tional binding energy DkGM/r (which isDkB Tvir at rvir) post-reionization ratio ofn /n \ 1, but we also consider the per particle. Thus, if the kinetic energy exceeds the potential \ c b \ e†ect of settingn /n 0.1. For example, a 1.8 and energy (or roughly ifT [ Tvir), the repulsive pressure gra- \ \c b \ \ \ nc/nb 1 yieldI21 1.0 at z 3 andI21 3.5 at z 5, dient force exceeds the attractive gravitational force and close to the values required to satisfy the Gunn-Peterson expels the gas on a dynamical time (or faster for halos with constraint at these redshifts (see, e.g., Efstathiou 1992). Note T ? Tvir). thatnc/nb Z 1 is required for the initial ionization of the gas We compare the thermal and gravitational energy (both in the universe (although this ratio may decline after of which are functions of radius) as a benchmark for decid- reionization). ing which gas shells are expelled from each halo. Note that We assume that the above uniform UV background illu- the infall of fresh IGM gas into the halo is also suppressed minates the outer surface of the gas cloud, located at the because of its excessive gas pressure, produced by the same virial radiusrvir, and penetrates from there into the cloud. photoionization heating process. The radiation photoionizes and heats the gas at each radius This situation stands in contrast to feedback due to to its equilibrium temperature, determined by equating the supernovae, which depends on the efficiency of converting heating and cooling rates. The latter assumption is justiÐed the mechanical energy of the supernovae into thermal by the fact that both the recombination time and the energy of the halo gas. The ability of supernovae to disrupt heating time are initially shorter than the dynamical time their host dwarf galaxies has been explored in a number of throughout the halo. At the outskirts of the halo the theoretical papers (e.g., Larson 1974; Dekel & Silk 1986; dynamics may start to change before the gas can be heated Vader 1986, 1987). However, numerical simulations (Mac up to its equilibrium temperature, but this simply means Low & Ferrara 1999) Ðnd that supernovae produce a hole that the gas starts expanding out of the halo during the in the gas distribution through which they expel the shock- process of photoheating. This outÑow should not alter the heated gas, leaving most of the cooler gas still bound. In the overall fraction of evaporated gas. case of reionization, on the other hand, energy is imparted The process of reionization is expected to be highly non- to the gas directly by the ionizing photons. A halo for which uniform because of the clustering of the ionizing sources a large fraction of the gas is unbound by reionization is thus and the clumpiness of the IGM. As time progresses, the H II prevented from further collapse and formation. regions around the ionizing sources overlap, and each halo When the gas in each halo is initially ionized, an ioniza- is exposed to ionizing radiation from an ever increasing tion shock front may be generated (see the discussion of Lya number of sources. While the external ionizing radiation absorbers by Donahue & Shull 1987). The dynamics of such may at Ðrst be dominated by a small number of sources, it a shock front have been investigated in the context of the quickly becomes more isotropic as its intensity builds up interstellar medium by Bertoldi & McKee (1990) and Ber- No. 1, 1999 PHOTOEVAPORATION OF DWARF GALAXIES 57 toldi (1989). Their results imply that the dynamics of gas in librium temperature for which the heating timetheat due to a a halo are not signiÐcantly a†ected by the shock front UV radiation Ðeld equals the cooling timetcool. The unless the thermal energy of the ionized gas is greater than decrease in the temperature at*b \ 10 is due to the its gravitational potential energy. Furthermore, since gas in increased importance of Compton cooling, which is pro- a halo is heated to the virial temperature even before reioni- portional to the gas density rather than its square. At a zation, the shock is weaker when the gas is ionized than a given density, gas is heated at reionization to the tem- typical shock in the interstellar medium. Also, as noted perature indicated by the solid curve, unless the net cooling above, the ionizing radiation reaching a given halo builds or heating time is too long. The dashed curves show the up in intensity over a considerable period of time. Thus we temperature where the net cooling or heating time equals do not expect the ionization shock associated with the Ðrst tH. By deÐnition, points on the solid curve have an inÐnite encounter of ionizing radiation to have a large e†ect on the net cooling or heating time, but there is also a substantial eventual fate of gas in the halo. regime at low*b where the net cooling or heating time is RESULTS greater thantH. However, this regime has only a minor 3. e†ect on halos, since the mean overdensity inside the virial We assume the most popular cosmology to date radius of a halo is of order 200. On the other hand, if gas \ \ leaves the halo and expands, it quickly enters the regime (Garnavich et al. 1998), with)0 0.3 and)" 0.7. We illustrate the e†ects of cosmological parameters by dis- where it cannot reach thermal equilibrium. \ \ Figure 2 presents an example for the structure of a halo playing the results also for)0 1, and for)0 0.3 and \ 2\ with an initial total mass of M \ 3 ] 107 M at z \ 8. We )" 0. The models all assume)b h 0.02 and a Hubble _ \ \ \ assume the same cosmological parameters as in Figure 1. constant h 0.5 if)0 1 and h 0.7 otherwise (where \ ~1 ~1 The bottom plot shows the baryon overdensity* versus H0 100 h km s Mpc ). b Figure 1 shows the temperature of the gas versus its bary- r/rvir, and reÑects our assumption of identical NFW proÐles for both the dark matter and the baryons. The middle plot onic overdensity*b relative to the cosmic average (see Efstathiou 1992). The curves are for z \ 8 and assume shows the neutral hydrogen fraction versusr/rvir, and the \ \ top plot shows the ratio of thermal energy per particle )0 0.3 and)" 0.7. We include intergalactic radiation \ \ [TE \ (3/2)k T (r)] to potential energy per particle with a Ñux given by equation (6) for a 1.8 and nc/nb 1. B \ [PE \ k o /(r) o , where /(r) is the gravitational potential] The dotted curve showstH tcool with no radiation Ðeld, versusr/r . The dashed curves assume an optically thin wheretH is the age of the universe, approximately equal to vir ] 9 ~1 ] ~3@2 ~1@2 halo, while the solid curves include radiative transfer and 6.5 10 h (1 z) )0 yr at high redshift. This curve indicates the temperature to which gas has time to self-shielding. The self-shielded neutral core is apparent cool through atomic transitions before reionization. This from the solid curves, but since the point where TE/PE \ 1 temperature is always near T \ 104 K since below this tem- occurs outside this core, the overall unbound fraction does perature the gas becomes mostly neutral and the cooling not depend strongly on the radiative transfer in this case. Its time is very long. It is likely that only atomic cooling is value is 67% assuming an optically thin halo and 64% when relevant before reionization since molecular hydrogen is easily destroyed by even a weak ionizing background (Haiman et al. 1996a). The solid curve shows the equi-

FIG. 2.ÈHalo structure after reionization, for a halo mass M \ 3 ] 107 \ M_ at z 8: baryon overdensity(*b) vs.r/rvir (bottom plot); neutral hydrogen fraction(nHI/nH) vs.r/rvir (middle plot); and the ratio of thermal FIG. 1.ÈTemperature T vs. baryon overdensity* : t \ t , no radi- energy (TE) to potential energy (PE) vs.r/r (top plot). The dashed curves \ b H cool vir ation Ðeld (dotted curve);theat tcool, with radiation (solid curve); and tH assume an optically thin halo, while the solid curves include radiative equals the net cooling/heating time, with radiation (dashed curves). The transfer and self-shielding. All the curves assume ) \ 0.3, ) \ 0.7, \ \ \ \ \ \ \ 0 " curves assume )0 0.3, )" 0.7, z 8, a 1.8, and nc/nb 1. a 1.8, and nc/nb 1. 58 BARKANA & LOEB Vol. 523 radiative transfer is included and only a fraction of the external photons make their way inside. Even when the opacity at the Lyman limit is large, some ionizing radiation still reaches the central parts of the halo because (1) the opacity drops quickly above the Lyman limit, and (2) the heated gas radiates ionizing photons inward. Figure 3 shows the unbound gas fraction after reioniza- tion as a function of the total halo mass. We assume ) \ \ \ 0 0.3,)" 0.7, andnc/nb 1. The three pairs of curves shown consist of a solid line (which includes radiative transfer) and a dashed line (which assumes an optically thin halo). From right to left, the Ðrst pair is for a \ 1.8 and z \ 8, the second is for a \ 5 and z \ 8, and the third is for a \ 1.8 and z \ 20. In each case the self-shielded core lowers the unbound fraction when we include radiative transfer (solid vs. dashed lines), particularly when the unbound fraction is sufficiently large that it includes part of the core itself. High-energy photons above the Lyman limit penetrate deep into the halo and heat the gas efficiently. Therefore, a steepening of the spectral slope from a \ 1.8 to a \ 5 decreases the temperature throughout the halo and lowers the unbound gas fraction. This is only partially com- pensated for by our UV Ñux normalization, which increases FIG. 4.ÈTotal halo mass for which 50% of the gas is unbound vs. reionization redshift. The solid line assumes a \ 1.8, and the dotted line I with increasing a so as to get the same density of ion- \ \ \ 21 assumes a 5, both for an)0 0.3, )" 0.7 cosmology. The other lines izing photons in equation (7). Increasing the reionization assume a \ 1.8 but di†erent cosmologies. The short-dashed line assumes redshift from z \ 8toz \ 20 increases the binding energy of ) \ 0.3, ) \ 0 and the long-dashed line assumes) \ 1. All assume 0 \ " 0 the gas, because the high-redshift halos are denser. nc/nb 1. Although the corresponding increase ofI21 with redshift (at a Ðxednc/nb) counteracts this change, the fraction of expelled gas is still reduced because of the deeper potential solid line assumes a \ 1.8 and the dotted line a \ 5, both \ \ \ wells of higher redshift halos. for)0 0.3 and)" 0.7. The other lines assume a 1.8 From plots similar to those shown in Figure 3, we Ðnd but di†erent cosmologies. The short-dashed line assumes the total halo mass at which the unbound gas fraction is ) \ 0.3, ) \ 0, and the long-dashed line assumes ) \ 1. 0 " \ 0 50%. We henceforth refer to this mass as the 50% mass. All assumenc/nb 1. Gas becomes unbound when its Figure 4 plots this mass as a function of the reionization thermal energy equals its potential binding energy. The redshift for di†erent spectra and cosmological models. The thermal energy depends on temperature, but the equi- librium temperature does not change much with redshift since we increase the UV Ñux normalization by the same (1 ] z)3 factor as the mean baryonic density. With this pre- scription for the UV Ñux, the 50% mass occurs at a value of the circular velocity that is roughly constant with redshift. Thus, for each curve, the change in mass with redshift is mostly due to the change in the characteristic halo density, which a†ects the relation between circular velocity and mass. The cosmological parameters have only a modest e†ect on the 50% mass and change it by up to 35% at a given redshift. Lowering)0 reduces the characteristic density of a halo of given mass, and so a higher mass is required in order to keep the gas bound. Adding a cosmological constant reduces the density further through*c (see eqs. [A1] and [A2]). For the three curves with a \ 1.8, the circular veloc- ity of the 50% mass equals 13 km s~1 at all redshifts, up to variations of a few percent. The spectral shape of the ionizing Ñux a†ects modestly the threshold circular velocity corresponding to the 50% mass, because assuming a steeper spectrum (i.e., with a larger a) reduces the gas temperature and thus requires a shallower potential to keep the gas bound. A higher Ñux normalization has the opposite e†ect of increasing the FIG. 3.ÈUnbound gas fraction vs. total halo mass. There are three threshold circular velocity. The left panel of Figure 5 shows pairs of curves, each consisting of a solid line (with radiative transfer) and a the variation of circular velocity with spectral shape, for two dashed line (without radiative transfer). From right to left, the three sets of \ \ curves correspond to a \ 1.8, z \ 8; a \ 5, z \ 8; and a \ 1.8, z \ 20. All normalizations(nc/nb 1 and nc/nb 0.1, solid and dashed \ \ \ the curves assume)0 0.3, )" 0.7, and nc/nb 1. curves, respectively). The right panel shows the complemen- No. 1, 1999 PHOTOEVAPORATION OF DWARF GALAXIES 59

mass cuto† is given by the lowest mass halo in which gas has assembled by the reionization redshift. We adopt for this low-mass cuto† the linear Jeans mass, which we calcu- late following Peebles (1993, that bookÏs ° 6). The gas tem- perature in the universe follows the cosmic microwave ] background temperature down to a redshift 1 zt D 740() h2)2@5, at which the baryonic Jeans mass is 1.9 ] 5b 2 ~1@2 10 ()b h ) M_. After this redshift, the gas tem- perature goes down as (1 ] z)2, so the baryon Jeans mass ] ] 3@2 acquires a factor of[(1 z)/(1 zt)] . Until now we have considered baryons only, but if we add dark matter, then the mean density (or the corresponding gravitational force) is increased by) /) , which decreases the baryonic Jeans ~3@20 b mass by()0/)b) . The corresponding total halo mass is )0/)b times the baryonic mass. Thus the Jeans cuto† before reionization corresponds to a total halo mass of ) h2 ~1@2 ) h2 ~3@5 1 ] z 3@2 M \ 6.9 ] 103 A 0 B A b B A B M . J 0.2 0.02 10 _ (9) This value agrees with the numerical spherical collapse cal- culations of Haiman et al. (1996b). FIG. 5.ÈCircular halo velocity at which 50% of the gas is unbound, as We thus calculate the total fraction of gas in the universe a function of ionizing spectrum. Both panels showVc on the vertical axis. that is bound in preexisting halos, and the fraction of this The left panel varies the spectral slope a for two values of the normal- \ \ gas which then becomes unbound at reionization. In Figure ization,nc/nb 1 (solid curve) andnc/nb 0.1 (dashed curve). The right panel varies the normalization for two spectral slopes, a \ 1.8 (solid curve) 6 we show the fraction of the collapsed gas which evapo- \ \ \ rates as a function of the reionization redshift. The solid line and a 5(dashed curve). All curves assume an)0 0.3, )" 0.7 cosmol- ogy. assumes a \ 1.8, and the dotted line assumes a \ 5, both for \ \ \ )0 0.3, )" 0.7. The other lines assume a 1.8, the short-dashed line with) \ 0.3, ) \ 0 and the long- \ 0 " \ tary case of varying the spectral normalization, using two dashed line with)0 1. All assumenc/nb 1 and a primor- values for the spectral slope (a \ 1.8 and a \ 5, solid and dial n \ 1 (scale-invariant) power spectrum. In each case we \ dashed curves, respectively). All curves assume an )0 0.3, normalized the CDM power spectrum to the present cluster \ \ ~0.5 )" 0.7 cosmology. abundance,p8 0.5)0 (see, e.g., Pen 1998), wherep8 is Obviously, 50% is a fairly arbitrary choice for the unbound gas fraction at which halos evaporate. Figure 3 shows that for a given halo, the unbound gas fraction changes from 10% to 90% over a factor of D60 in mass, or a factor of D4 in velocity dispersion. When 50% of the gas is unbound, however, the rest of the gas is also substantially heated, and we expect the process of collapse and fragmen- tation to be inhibited. In the extreme case where the gas expands until a steady state is achieved where it is pressure conÐned by the IGM, less than 10% of the original gas is left inside the virial radius. However, continued infall of dark matter should limit the expansion. Numerical simula- tions may be used to deÐne more precisely the point at which gas halos are disrupted. Clearly, photoevaporation a†ects even halos with masses well above the 50% mass, although these halos do not completely evaporate. Note that it is also clear from Figure 3 that not including radi- ative transfer would have only a minor e†ect on the value of the 50% mass (typically D5%). Given the values of the unbound gas fraction in halos of di†erent masses, we can integrate to Ðnd the total gas frac- tion in the universe that becomes unbound at reionization. This calculation requires the abundance distribution of halos, which is readily provided by the Press-Schechter mass function for CDM cosmologies (relevant expressions FIG. 6.ÈFraction of the collapsed gas that evaporates from halos at reionization vs. the reionization redshift. The solid line assumes a \ 1.8, are given, e.g., in NFW). The high-mass cuto† in the inte- \ \ \ and the dotted line assumes a 5, both for an)0 0.3, )" 0.7 cosmol- gration is given by the lowest mass halo for which the ogy. The other lines assume a \ 1.8 but di†erent cosmologies. The short- unbound gas fraction is zero, since halos above this mass dashed line assumes) \ 0.3, ) \ 0, and the long-dashed line assumes \ 0 \ " are not signiÐcantly a†ected by the UV radiation. The low- )0 1. All assume nc/nb 1. 60 BARKANA & LOEB Vol. 523 the rms amplitude of mass Ñuctuations in spheres of radius Ðles, the total unbound gas fraction is 19.8%, and the halo ~1 ] 7 8 h Mpc. The fraction of collapsed gas that is unbound is mass that loses 50% of its baryons is 5.25 10 M_. If we D0.4È0.7 at z \ 6, and it increases with redshift. This frac- let the mass and the baryons follow the proÐle of equation tion clearly depends strongly on the halo abundance but is (4), the corresponding results are 20.0% and 5.31 ] 107 M \ ] 7 _ relatively insensitive to the spectral slope a of the ionizing for b 10 in equation (4) and 20.9% and 6.84 10 M_ for radiation. In hierarchical models, the characteristic mass b \ 5 (i.e., a larger core). With an NFW mass proÐle but gas (and binding energy) of virialized halos is smaller at higher in hydrostatic equilibrium at the virial temperature, the redshifts, and a larger fraction of the collapsed gas therefore unbound fraction is 19.2%, and the 50% mass is 4.33 ] 7 \ escapes once it is photoheated. Among the three cosmo- 10 M_. If we let the gas temperature beT 2Tvir, the logical models, the characteristic mass at a given redshift is unbound fraction is 22.0%, and the 50% mass is 1.18 \ \ \ ] 8 smallest for)0 1 and largest for )0 0.3, )" 0. 10 M_. For clouds of gas that condense by cooling for a In Figure 7 we show the total fraction of gas in the uni- Hubble time, the unbound fraction is 18.2%, and the 50% ] 7 verse that evaporates at reionization. The solid line assumes mass is 3.38 10 M_. We conclude that centrally concen- a \ 1.8, and the dotted line assumes a \ 5, both for ) \ trated gas clouds are in general more e†ective at retaining \ \ 0 0.3,)" 0.7. The other lines assume a 1.8, the short- their gas, but the e†ect on the overall unbound gas fraction dashed line with) \ 0.3, ) \ 0 and the long-dashed line in the universe is modest, even for large variations in the \ 0 "\ \ with)0 1. All assumenc/nb 1. For the di†erent cosmol- proÐle. If we return to the NFW proÐle but adopt f 0.01 ogies, the total unbound fraction goes up to 20%È25% if instead of f \ 0.5 in the NFW prescription for Ðnding the reionization occurs as late as z \ 6È7; in this case a sub- collapse redshift (see Appendix A), we Ðnd an unbound frac- ] 7 stantial fraction of the total gas in the universe undergoes tion of 20.3% and a 50% mass of 6.06 10 M_. Finally, the process of expulsion from halos. However, this fraction lowering) by a factor of 2 changes the unbound fraction b ] 7 typically decreases at higher redshifts. Although a higher to 19.0% and the 50% mass to 5.44 10 M_. Our predic- fraction of the collapsed gas evaporates at higher z (see Fig. tions appear to be robust against variations in the model 6), a smaller fraction of the gas in the universe lies in halos parameters. in the Ðrst place. The latter e†ect dominates except for the \ 4. IMPLICATIONS FOR THE IGM AND FOR LOW-REDSHIFT open model up to z D 7. As is well known, the )0 1 model produces late structure formation, and indeed the OBJECTS collapsed fraction decreases rapidly with redshift in this Our calculations show that a substantial fraction of gas cosmological model. Thelow-) models approach the in the universe may lie in virialized halos before reionization \ 0 )0 1 behavior at high z, but this occurs faster for the Ñat and that most of it evaporates out of the halos when it is model with a cosmological constant than for the open photoionized and heated at reionization. The resulting out- model with the same value of )0. Ñows of gas from halos may have interesting implications Changing the dark matter and gas proÐles as discussed in for the subsequent evolution of structure in the IGM. We ° 2 has a modest e†ect on the results. For example, with discuss some of these implications in this section. \ \ \ )0 0.3, )" 0.7, and z 8, and for our standard model In the pre-reionization epoch, a fraction of the gas in the where the gas and dark matter follow identical NFW pro- dense cores of halos may fragment and form stars. Some star formation is, of course, needed in order to produce the ionizing Ñux that leads to reionization. These Population III stars produce the Ðrst metals in the universe, and they may make a substantial contribution to the enrichment of the IGM. Numerical models by Mac Low & Ferrara (1999) suggest that feedback from supernovae is very efficient at expelling metals from dwarf galaxies of total mass 3.5 ] 108 M_, although it ejects only a small fraction of the inter- stellar medium in these hosts. Obviously, the metal expul- sion efficiency depends on the presence of clumps in the supernova ejecta (Franco et al. 1993) and on the supernova rate, the latter depending on the unknown star formation rate and the initial mass function of stars at high redshifts. Reionization provides an alternative method for expelling metals efficiently out of dwarf galaxies by directly photo- heating the gas in their halos, leading to its evaporation along with its metal content.1 Gas that falls into halos and is expelled at reionization attains a di†erent entropy than if it had stayed at the mean density of the universe. Gas that collapses into a halo is at a high overdensity when it is photoheated, and it is therefore at a lower entropy than if it were heated to the same tem-

FIG. 7.ÈTotal fraction of gas in the universe which evaporates from halos at reionization, vs. the reionization redshift. The solid line assumes 1 Note that we have assumed zero metallicity in calculating cooling. \ \ \ \ a 1.8, and the dotted line assumes a 5, both for an )0 0.3, )" 0.7 Even if some metals had already been mixed into the IGM, the metallicity cosmology. The other lines assume a \ 1.8 but di†erent cosmologies. The of newly formed objects was likely too low to a†ect cooling since even at short-dashed line assumes) \ 0.3, ) \ 0 and the long-dashed line z D 3 the typical metallicity of the Lya forest has been observed to be \ 0 \ " assumes)0 1. All assume nc/nb 1. \0.01 solar (Songaila & Cowie 1996; Tytler et al. 1995). No. 1, 1999 PHOTOEVAPORATION OF DWARF GALAXIES 61 perature at the mean cosmic density. However, the overall Foltz 1998, and references therein) should probe di†erent change in the entropy density of the IGM is small for two H I column densities in galactic outÑow absorbers but reasons. First, even at z \ 6 only about 25% of the gas in similar column densities in the larger, more common the universe undergoes evaporation. Second, the gas absorbers. Follow-up observations with high spectroscopic remains in ionization equilibrium and is photoheated resolution could reveal the velocity Ðelds of these outÑows. during its initial expansion. For example, if z \ 6, ) \ 0.3, Although much of the gas in the universe evaporated at \ \ \ 0 )" 0.7, nc/nb 1, and a 1.8, then the recombination reionization, the underlying dark matter halos continued to time becomes longer than the dynamical time only when the evolve through infall and merging, and the heated gas may gas expands down to an overdensity of 26, at which point its have accumulated in these halos at lower redshifts. This temperature is 22,400 K compared with an initial latter process has been discussed by a number of authors, (nonequilibrium) temperature of 19,900 K for gas at the with an emphasis on the e†ect of reionization and the mean density. The resulting overall reduction in the entropy resulting heating of gas. Thoul & Weinberg (1996) found a is the same as would be produced by reducing the tem- reduction of D50% in the collapsed gas mass due to perature of the entire IGM by a factor of 1.6. This factor heating, for a halo ofV \ 50 km s~1 at z \ 2, and a com- c \ ~1 reduces to 1.4 if we increase z to 8 or increase a to 5. Note plete suppression of infall belowVc 30 km s . The e†ect that Haehnelt & Steinmetz (1998) showed that di†erences in is thus substantial on halos with virial temperatures well temperature by a factor of 3È4 result in possibly observable above the gas temperature. Their interpretation is that pres- di†erences in the Doppler parameter distribution of Lya sure support delays turnaround substantially and slows the absorption lines at redshifts 3È5. subsequent collapse. Indeed, as noted in ° 2, the ratio of the When the halos evaporate, recombinations in the gas pressure force to the gravitational force on the gas is could produce Lya lines or radiation from two-photon roughly equal to the ratio of its thermal energy to its poten- transitions to the ground state of hydrogen. However, a tial energy. For a given enclosed mass, the potential energy simple estimate shows that the resulting luminosity is too of a shell of gas increases as its radius decreases. Before small for direct detection unless these halos are illuminated collapse, each gas shell expands with the Hubble Ñow until by an internal ionizing source. In an externally illuminated, its expansion is halted and then reversed. Near turnaround, \ 8 z 6, 10 M_ halo, our calculations imply a total of the gas is weakly bound, and the pressure gradient may D1 ] 1050 recombinations per second. Note that the prevent collapse even for gas below the halo virial tem- number of recombinations is dominated by the high-density perature. On the other hand, gas that is already well within core, and if we did not include self-shielding we would the virial radius is tightly bound, which explains our lower ~1 obtain an overestimate by a factor of D15. If each recombi- value ofVc D 13 km s for halos that lose half their gas at nation releases one or two photons with a total energy of reionization. \ \ 10.2 eV, then for)0 0.3 and)" 0.7 the observed Ñux is Three-dimensional numerical simulations (Quinn et al. D5 ] 10~20 ergs s~1 cm~2. This Ñux is well below the 1996; Weinberg et al. 1997; Navarro & Steinmetz 1997) sensitivity of the planned Next Generation Space Telescope, have also explored the question of whether dwarf galaxies even if part of this Ñux is concentrated in a narrow line. could re-form atz Z 2. The heating by the UV background The photoionization heating of the gaseous halos of was found to suppress infall of gas into even larger halos ~1 dwarf galaxies resulted in outÑows with a characteristic (Vc D 75 km s ), depending on the redshift and on the velocity of D20È30 km s~1. These outÑows must have ionizing radiation intensity. Navarro & Steinmetz (1997) induced peculiar velocities of a comparable magnitude in noted that photoionization reduces the cooling efficiency of the IGM surrounding these galaxies. The e†ect of the out- gas at low densities, which suppresses further the late infall Ñows on the velocity Ðeld and entropy of the IGM at at redshifts below 2. We note that these various simulations z \ 5È10 could in principle be searched for in the absorp- assume an isotropic ionizing radiation Ðeld and do not cal- tion spectra of high-redshift sources, such as quasars. These culate radiative transfer. Photoevaporation of a gas cloud small-scale Ñuctuations in velocity and the resulting tem- has been calculated in a two-dimensional simulation perature Ñuctuations have been seen in recent simulations (Shapiro, Raga, & Mellema 1998), and methods are being by Bryan et al. (1998). However, the small halos responsible developed for incorporating radiative transfer into three- for these outÑows were only barely resolved even in these dimensional cosmological simulations (e.g., Abel, Norman, high-resolution simulations of a small volume. & Madau 1999; Razoumov & Scott 1999). The evaporating galaxies could contribute to the high Our results have interesting implications for the fate of column density end of the Lya forest (see Bond et al. 1988). gas in low-mass halos. Gas evaporates at reionization from \ ~1 For example, shortly after being photoionized, a z 8, 5 halos belowVc D 13 km s , or a velocity dispersion p D 10 ] 7 ~1 10 M_ halo has a neutral hydrogen column density of km s . A similar value of the velocity dispersion is also 2 ] 1016 cm~2 at an impact parameter of0.5r \ 0.66 kpc, required to reach a virial temperature of 104 K, allowing ] 17 ~2 ] 20 ~2vir ] 18 6 10 cm at0.25rvir, and 9 10 cm (or 9 10 atomic cooling and perhaps star formation before reioniza- cm~2 if we do not include self-shielding) at0.1r (assuming tion. Thus halos withp Z 10 km s~1 could have formed \ \ \ \ vir )0 0.3, )" 0.7, a 1.8, andnc/nb 1). These column stars before reionization. They would have kept their gas densities will decline as the gas expands out of the host after reionization and could have had ongoing star forma- galaxy. Abel & Mo (1998) have suggested that a large frac- tion subsequently. These halos were the likely sites of Popu- tion of the Lyman limit systems at z D 3 may correspond to lation III stars and could have been the progenitors of minihalos that survived reionization. Remnant absorbers dwarf galaxies in the local universe (seeMiralda-Escude & due to galactic outÑows can be distinguished from large- Rees 1998). On the other hand, halos withp [ 10 km s~1 scale absorbers in the IGM by their compactness. Close could not have cooled before reionization. Their warm gas lines of sight due to quasar pairs or gravitational lensed was completely evaporated from them at reionization and quasars (see, e.g., Crotts & Fang 1998; Petry, Impey, & could not have returned to them until very low redshifts, 62 BARKANA & LOEB Vol. 523 possiblyz [ 1, so that their stellar population should be host galaxy at a redshift much lower than their formation relatively young. redshift, they will survive disruption because of their high It is interesting to compare these predictions with the densities. The subsequent of gas could result from properties of dwarf spheroidal galaxies in the Local Group passages of the dwarf halos through the gaseous tidal tail of that have low central velocity dispersions. At Ðrst sight this a merger event or through the disk of the parent galaxy. In appears to be a difficult task. The dwarf galaxies vary this case, retention of cold, dense, and possibly metal- greatly in their properties, with many showing evidence for enriched gas against heating by the UV background multiple episodes of star formation as well as some very old requires a shallower potential well than accumulating warm stars (see the recent review by Mateo 1998). Another obs- gas from the IGM. Simulations of galaxy encounters tacle is the low temporal resolution of age indicators for old (Barnes & Hernquist 1992; Elmegreen, Kaufman, & Tho- \ \ stellar populations. For example, if)0 0.3 and )" 0.7, masson 1993) have found that dwarf galaxies could form then the age of the universe is 43% of its present age at but with small amounts of dark matter. However, the initial z \ 1 and 31% at z \ 1.5. Thus stars that formed at these conditions of these simulations assumed parent galaxies redshifts may already be D10 Gyr old at present, and are with a smooth dark matter distribution rather than clumpy difficult to distinguish from stars that formed at z [ 5. halos with dense subhalos inside them. Simulations by Nevertheless, one of our robust predictions is that most Klypin et al. (1999) suggest that galaxy halos may have early halos withp [ 10 km s~1 could not have formed stars large numbers of dark matter satellites, most of which have in the standard hierarchical scenario. Globular clusters no associated stars. If true, this implies that the dwarf spher- belong to one class of objects with such a low-velocity dis- oidal galaxies might be explained even if only a small frac- persion. Peebles & Dicke (1968) originally suggested that tion of dwarf dark halos accreted gas and formed stars. globular clusters may have formed at high redshifts, before A common origin for the Milky WayÏs dwarf satellites their parent galaxies. However, in current cosmological (and a number of halo globular clusters), as remnants of models, most mass Ñuctuations on globular cluster scales larger galaxies accreted by the Milky Way galaxy, has been were unable to cool e†ectively and fragment until z D 10 suggested on independent grounds. These satellites appear and were evaporated subsequently by reionization. We note to lie along two (e.g., Majewski 1994) or more (Lynden-Bell that Fall & Rees (1985) proposed an alternative formation & Lynden-Bell 1995; Fusi-Pecci et al. 1995) polar great scenario for globular clusters involving a thermal instability circles. The star formation history of the dwarf galaxies (e.g., inside galaxies with properties similar to those of the Milky Grebel 1998) constrains their merger history and implies Way. Globular clusters have also been observed to form in that the fragmentation responsible for their appearance galaxy mergers (e.g., Miller et al. 1997). It is still possible must have occurred early in order to be consistent with the that some of the very oldest and most metal-poor globular variation in stellar populations among the supposed frag- clusters originated fromz Z 10, before the UV background ments (Unavane, Wyse, & Gilmore 1996; Olszewski 1998). had become strong enough to destroy the molecular hydro- Observations of interacting galaxies (outside the Local gen in them. However, primeval globular clusters should Group) also suggest the formation of ““ tidal dwarf galaxies ÏÏ have retained their dark halos, but observations suggest (e.g., Duc & Mirabel 1997). that globular clusters are not embedded in dark matter Finally, there exists the possibility that the measured halos (Moore 1996; Heggie & Hut 1995). velocity dispersion of stars in the dwarf spheroidals under- Another related population is the nine dwarf spheroidals estimates the velocity dispersion of their dark halos. in the Local Group with central velocity dispersions p [ 10 Assuming that the stars are in equilibrium, their velocity km s~1, including Ðve below 7 km s~1 (e.g., Mateo 1998). In dispersion could be lower than that of the halo if the mass the hierarchical clustering scenario, the dark matter in a proÐle is shallower than isothermal beyond the stellar core present halo was most probably divided at reionization radius. As discussed in ° 2, halo proÐles are thought to vary among several progenitors, which have since merged. The from being shallow in a central core to being steeper than velocity dispersions of these progenitors were likely even isothermal at larger distances. The velocity dispersion and lower than that of the Ðnal halo. Thus the dwarf galaxies mass to light ratio of a dwarf spheroidal could also appear could not have formed stars at high redshifts, and their high if it is nonspherical or the stellar orbits are anisotropic. formation presents an intriguing puzzle. There are two pos- Some dwarf spheroidals may even not be dark matterÈ sible solutions to this puzzle, (1) the ionizing background dominated if they are tidally disrupted (e.g., Kroupa 1997). dropped dramatically at low redshifts, allowing the dwarf The observed properties of dwarf spheroidals require a ~3 galaxies to form atz [ 1, or (2) the measured stellar velocity central mass density of order 0.1M_ pc (e.g., Mateo dispersions of the dwarf galaxies are well below the velocity 1998), which is D7 ] 105 times the present critical density. dispersions of their dark matter halos. The stars therefore reside either in high-redshift halos or in Unlike globular clusters, the dwarf spheroidal galaxies the very central parts of low-redshift halos. Detailed obser- are dark matterÈdominated. The dark halo of a present-day vations of the velocity dispersion proÐles of these stars dwarf galaxy may have virialized at high redshifts but acc- could be used to discriminate between these possibilities. reted its gas at low redshift from the IGM. However, for CONCLUSIONS dark matter halos accumulating primordial gas, Kepner et 5. ] 4 al. (1997) found that even ifI21(z) declines as (1 z) below We have shown that the photoionizing background radi- z \ 3, only halos withV Z 20 km s~1 can form atomic ation that Ðlled the universe during reionization likely \ c ~1 hydrogen by z 1, andVc Z 25 km s is required to form boiled most of the virialized gas out of CDM halos at that molecular hydrogen. time. The evaporation process probably lasted of order a Alternatively, the dwarf dark halos could have accreted Hubble time because of the gradual increase in the UV cold gas at low redshift from a larger host galaxy rather background as the H II regions around individual sources than from the IGM. As long as the dwarf halos join their overlapped and percolated until the radiation Ðeld inside No. 1, 1999 PHOTOEVAPORATION OF DWARF GALAXIES 63 them grew up to its cosmic valueÈamounting to the full the value of the circular velocity threshold is nearly inde- contribution of sources from the entire Hubble volume. The pendent of redshift. The corresponding halo mass changes, 8 \ 7 \ precise reionization history depends on the unknown star however, from D10 M_ at z 5toD10 M_ at z 20, formation efficiency and the potential existence of mini- assuming a shallow ionizing spectrum. quasars in newly formed halos (Haiman & Loeb 1998). Based on these Ðndings, we expect that both globular The total fraction of the cosmic baryons which partici- clusters and Local Group dwarf galaxies with velocity dis- pate in the evaporation process depends on the reionization persions[10 km s~1 formed at low redshift, most probably redshift, the ionizing intensity, and the cosmological param- inside larger galaxies. The latter possibility has been sug- eters, but it is not very sensitive to the precise gas and dark gested previously for the Milky WayÏs dwarf satellites based matter proÐles of the halos. The central core of halos is on their location along polar great circles. typically shielded from the external ionizing radiation by the surrounding gas, but this core typically contains \20% of the halo gas and has only a weak e†ect on the global We are grateful to JordiMiralda-Escude , Chris McKee, behavior of the gas. We have found that halos are disrupted Roger Blandford, Lars Hernquist, David Spergel, and Jim up to a circular velocityV D 13 km s~1 for a shallow, Peebles for useful discussions. We also thank Renyue Cen c ~1 quasar-like spectrum, orVc D 11 km s for a stellar spec- and JordiMiralda-Escude for assistance with the reaction trum, assuming the photoionizing sources build up a and cooling rates. R. B. acknowledges support from Insti- density of ionizing photons comparable to the mean cosmo- tute Funds. This work was supported in part by the NASA logical density of baryons. At this photoionizing intensity, NAG 5-7039 grant (for A. L.).

APPENDIX A

HALO PROFILE

We follow the prescription of NFW for obtaining the density proÐles of dark matter halos, but instead of adopting a constant overdensity of 200, we use the Ðtting formula of Bryan & Norman (1998) for the virial overdensity, \ 2] [ 2 *c 18n 82d 39d (A1) for a Ñat universe with a cosmological constant and \ 2] [ 2 *c 18n 60d 32d (A2) [ for an open universe, where d 4 )(z) 1. Given)0 and)", we deÐne ] 3 \ )0(1 z) )(z) ] 3] ] [ [ ] 2 . (A3) )0(1 z) )" (1 )0 )")(1 z) In equation (3), c is determined for a givendc by the relation * c3 d \ c . (A4) c 3 ln (1 ] c) [ c/(1 ] c) The characteristic density is given by 1 ] z 3 d \ C( f ))(z)A collB . (A5) c 1 ] z

For a given halo of mass M, the collapse redshiftzcoll is deÐned as the time at which a mass M/2 was Ðrst contained in progenitors more massive than some fraction f of M. This is computed using the extended Press-Schechter formalism (e.g., Lacey & Cole 1993). NFW Ðnd that f \ 0.01 Ðts their z \ 0 simulation results best. Since we are interested in high redshifts when mergers are very frequent, we adopt the more natural f \ 0.5 but also check the f \ 0.01 case. (For example, the survival \ ] 7 time of a z 8, 5 10 M_ halo before it merges is D30%È40% of the age of the universe at that redshift [Lacey & Cole 1993].) In both cases we adopt the normalization of NFW, which is C(0.5) \ 2 ] 104 and C(0.01) \ 3 ] 103.

APPENDIX B

RADIATIVE TRANSFER

We neglect atomic transitions of helium atoms in the radiative transfer calculation. We only consider halos for which kB T is well below the ionization energy of hydrogen, and so following Tajiri & Umemura (1998) we assume that recombinations to excited levels do not result in further ionizations. On the other hand, recombinations to the ground state result in the emission \ of ionizing photons all of which are in a narrow frequency band just above the Lyman limit frequencyl lL. We follow separately these emitted photons and the external incoming radiation. The external photons undergo absorption with an 64 BARKANA & LOEB Vol. 523 optical depth at the Lyman limit determined by dq lL \ p (l )n . (B1) ds H I L H I The emitted photons nearlL are propagated by the equation of radiative transfer, dI l \[p (l)n I ] g . (B2) ds H I H I l l \ \ Assuming all emitted photons are just abovel lL, we can setpH I(l) pH I(lL) in this equation and propagate the total number Ñux of ionizing photons, = P Il F1 4 dl . (B3) lL hl The emissivity term for this quantity is = P gl \ u dl aH I nH II ne , (B4) lL hl 4n wherea is the total recombination coefficient to all bound levels of hydrogen and u is the fraction of recombinations to the H I \ [ ground state. In terms of Table 5.2 of Spitzer (1978),u (/1 /2)//1. We Ðnd that a convenient Ðtting formula up to 64,000 K, accurate to 2%, is (with T in K) u \ 0.205 [ 0.0266 ln (T ) ] 0.0049 ln 2(T ) . (B5) When these photons are emitted, they carry away the kinetic energy of the absorbed electron. When the photons are reabsorbed at some distance from where they were emitted, they heat the gas with this extra energy. SincekB T > hlL, we do not need to compute the exact frequency distribution of these photons. Instead we solve a single radiative-transfer equation for the total Ñux of energy (above the ionization energy of hydrogen) in these photons, = P Il [ F2 4 (hl hlL) dl . (B6) lL hl The emissivity term for radiative transfer ofF2 is = ] ~11 [ P gl [ \ 2.07 10 s1(b) s2(b) ~3 ~1 ~1 (hl hlL) dl 1@2 nH II ne ergs cm s sr (B7) lL hl T 4n \ where b hlL/kT , T is in K and the functionss1 ands2 are given in Table 6.2 of Spitzer (1978). We Ðnd a Ðtting formula up to 64,000 K, accurate to 2% (with T in K): 040.78 , if T \ 103 K, s (T ) [ s (T ) \ 5 (B8) 1 2 60 [ 0.172 ] 0.255 ln (T ) [ 0.0171 ln 2(T ) , otherwise.

From each point we integrate along all lines of sight to ÐndqlL, F1, andF2 as a function of angle. Because of spherical symmetry, we do this only at each radius, and the angular dependence only involves h, the angle relative to the radial direction. We then integrate to Ðnd the photoionization rate. For each atomic species, the rate is 4n = \ P P Il ~1 !ci d) pi(l) dl s , (B9) 0 li hl whereli andpi(l) are the threshold frequency and cross section for photoionization of species i, given in Osterbrock (1989; see eq. [2.31]) for H I,HeI, and He II]. For the external photons the UV intensity isI e~ql, with the boundary intensity \ ~a ~3 l,0 Il,0 IlL,0(l/lL) as before, andql approximated asqlL(l/lL) . Sincepi(l) has the simple form of a sum of two power laws, the frequency integral in!ci can be done analytically, and only the angular integration is computed numerically (see the similar but simpler calculation of Tajiri & Umemura 1998). There is an additional contribution to photoionization for H I \ 4n only, from the emitted photons just abovel lL, given by/0 d)pi(lL)F1. The photoheating rate per unit volume is ni vi, whereni is the number density of species i and 4n = \ P P Il [ ~1 vi d) pi(l)(hl hlL) dl ergs s . (B10) 0 li hl The rate for the external UV radiation is calculated for each atomic species similarly to the calculation of! . The emitted 4n ci photons contribute tovH I an extra amount of /0 d)pi(lL)F2. REFERENCES Abel, T., & Mo, H. J. 1998, ApJ, 494, L151 Black, J. H. 1981, MNRAS, 197, 553 Abel, T., Norman, M. L., & Madau, P. 1999, ApJ, 531, in press Bond, J. R., Szalay, A. S., & Silk, J. 1988, ApJ, 324, 627 Babul, A., & Rees, M. 1992, MNRAS, 253, 31 Bryan, G. L., Machacek, M., Anninos, P., & Norman, M. L. 1998, ApJ, 517, Barnes, J. E., & Hernquist, L. 1992, Nature, 360, 715 13 Bertoldi, F. 1989, ApJ, 346, 735 Bryan, G., & Norman, M. 1998, ApJ, 495, 80 Bertoldi, F., & McKee, C. F. 1990, ApJ, 354, 529 Burkert, A. 1995, ApJ, 447, L25 No. 1, 1999 PHOTOEVAPORATION OF DWARF GALAXIES 65

Crotts, A. P. S., & Fang, Y. 1998, ApJ, 502, 16 Miller, B. W., Whitmore, B. C., Schweizer, F., & Fall, S. M. 1997, AJ, 114, Duc, P.-A., & Mirabel, I. F. 1997, in Proc. IAU Symp. 187, Cosmic Chemi- 2381 cal Evolution, ed. J. W. Truran & K. Nomoto (Dordrecht: Kluwer), in Miralda-Escude , J. 1998, ApJ, 501, 15 press Miralda-Escude, J., Haehnelt, M., & Rees, M. J. 1999, ApJ, submitted Dekel, A., & Silk, J. 1986, ApJ, 303, 39 (astro-ph/9812306) Dey, A., Spinrad, H., Stern, D., Graham, J. R., & Cha†ee, F. H. 1998, ApJ, Miralda-Escude, J., & Rees, M. J. 1998, ApJ, 497, 21 498, L93 Moore, B. 1996, ApJ, 461, L13 Donahue, M., & Shull, J. M. 1987, ApJ, 323, L13 Moore, B., Governato, F., Quinn, T., Stadel, J., & Lake, G. 1998, ApJ, 499, Elmegreen, B. G., Kaufman, M., & Thomasson, M. 1993, ApJ, 412, 90 L5 Efstathiou, G. 1992, MNRAS, 256, 43 Navarro, J. F., Frenk, C. S., & White, S. D. M. 1997, ApJ, 490, 493 (NFW) Fall, S. M., & Rees, M. J. 1985, ApJ, 298, 18 Navarro, J. F., & Steinmetz, M. 1997, ApJ, 478, 13 Ferland, G. J., Peterson, B. M., Horne, K., Welsh, W. F., & Nahar, S. N. Olszewski, E. W. 1998, in Galactic Halos: A UC Santa Cruz Workshop, 1992, ApJ, 387, 95 ed. D. Zaritski (San Francisco: ASP) Franco, J., Ferrara, A., Roczyska, M., Tenorio-Tagle, G., & Cox, D. P. Osterbrock, D. E. 1989, Astrophysics of Gaseous Nebulae and Active 1993, ApJ, 407, 100 Galactic Nuclei (Mill Valley: University Science Books) Fusi-Pecci, F., Ballazzini, M., Cacciari, C., & Ferraro, F. R. 1995, AJ, 100, Peebles, P. J. E. 1993, Principles of Physical Cosmology (Princeton: Prin- 1664 ceton Univ. Press) Garnavich, P. M., et al. 1998, ApJ, 509, 74 Peebles, P. J. E., & Dicke, R. H. 1968, ApJ, 154, 891 Gnedin, N. Y., & Ostriker, J. P. 1997, ApJ, 486, 581 Pen, U.-L. 1998, ApJ, 498, 60 Grebel, E. 1998, in IAU Symp. 192, The Stellar Content of Local Group Petry, C. E., Impey, C. D., & Foltz, C. B. 1998, ApJ, 494, 60 Galaxies, ed. P. Whitelock & R. Cannon (San Francisco: ASP) Press, W. H., & Schechter, P. 1974, ApJ, 187, 425 Haiman, Z., & Loeb, A. 1998, ApJ, 503, 505 Quinn, T., Katz, N., & Efstathiou, G. 1996, MNRAS, 278, L49 ÈÈÈ. 1999, ApJ, 519, 479 Razoumov, A., & Scott, D. 1999, MNRAS, submitted (astro-ph/9810425) ÈÈÈ. 1998c, in AIP Conf. Proc. 470, After the Dark Ages: When Gal- Rees, M. J. 1986, MNRAS, 218, 25 axies Were Young (the Universe at 2 \ z \ 5), ed. S. Holt & E. Smith Shapiro, P. R., & Kang, H. 1987, ApJ, 318, 32 (Woodbury: AIP), Shapiro, P. R., Raga, A. C., & Mellema, G. 1998, in Molecular Hydrogen in Haiman, Z., Rees, M., & Loeb, A. 1996a, ApJ, 476, 458 (erratum 484, 985 the Early Universe, Memorie Della Societa Astronomica Italiana, Vol. [1997]) 69, ed. E. Corbelli, D. Galli, & F. Palla (Florence: Soc. Astron. Italiana), Haiman, Z., Thoul, A. A., & Loeb, A. 1996b, ApJ, 464, 523 463 Haehnelt, M. G., & Steinmetz, M. 1998, MNRAS, 298, 21 Songaila, A., & Cowie, L. L. 1996, AJ, 112, 335 Heggie, D. C., & Hut, P. 1995, in IAU Symp. 174, Dynamical Evolution of Spinrad, H., Stern, D., Bunker, A., Dey, A., Lanzetta, K., Yahil, A., Star Clusters-Confrontation of Theory and Observations, ed. P. Hut & Pascarelle, S., &Ferna ndez-Soto, A. 1998, AJ, 116, 2617 J. Makino (Dordrecht: Kluwer), 303 Spitzer, L., Jr. 1978, Physical Processes in the Interstellar Medium (New Hu, E. M., Cowie, L. L., & McMahon, R. G. 1998, ApJ, 502, L99 York: Wiley) Katz, N., Weinberg, D. H., & Hernquist, L. 1996, ApJS, 105, 19 Spitzer, L., Jr., & Hart, M. H. 1971, ApJ, 166, 483 Kepner, J. V., Babul, A., & Spergel, D. N. 1997, ApJ, 487, 61 Stecher, T. P., & Williams, D. A. 1967, ApJ, 149, L29 Klypin, A. A., Kravtsov, A. V., Valenzuela, O., & Prada, F. 1999, ApJ, 522, Tajiri, Y., & Umemura, M. 1998, ApJ, 502, 59 82 Thoul, A. A., & Weinberg, D. H. 1996, ApJ, 465, 608 Kravtsov, A. V., Klypin, A. A., Bullock, J. S., & Primack, J. R. 1998, ApJ, Tytler, D., Fan, X.-M., Burles, S., Cottrell, L., Davis, C., Kirkman, D., & 502, 48 Zuo, L. 1995, in QSO Absorption Lines, ed. G. Meylan (Berlin: Kroupa, P. 1997, NewA, 2, 139 Springer), 289 Lacey, C. G., & Cole, S. M. 1993, MNRAS, 262, 627 Unavane, M., Wyse, R. F. G., & Gilmore, G. 1996, MNRAS, 278, 727 Larson, R. B. 1974, MNRAS, 271, 676L Vader, J. P. 1986, ApJ, 305, 669 Lynden-Bell, D., & Lynden-Bell, R. M. 1995, MNRAS, 275, 429 ÈÈÈ. 1987, ApJ, 317, 128 Mac Low, M., & Ferrara, A. 1999, ApJ, 513, 142 Verner, D. A., & Ferland, G. J. 1996, ApJS, 103, 467 Majewski, S. R. 1994, ApJ, 431, L17 Voronov, G. S. 1997, At. Data Nucl. Data Tables, 65, 1 Mateo, M. 1998, ARA&A, 36, 435 Weinberg, D. H., Hernquist, L., & Katz, N. 1997, ApJ, 477, 8