Articles

THE OBSERVED PROPERTIES OF DWARF GALAXIES IN AND AROUND THE LOCAL GROUP

Published 2012 June 4 © 2012. National Research Council Canada. All rights reserved.
, , Citation Alan W. McConnachie 2012 AJ 144 4 DOI 10.1088/0004-6256/144/1/4

1538-3881/144/1/4

ABSTRACT

Positional, structural, and dynamical parameters for all dwarf galaxies in and around the Local Group are presented, and various aspects of our observational understanding of this volume-limited sample are discussed. Over 100 nearby galaxies that have distance estimates reliably placing them within 3 Mpc of the Sun are identified. This distance threshold samples dwarfs in a large range of environments, from the satellite systems of the MW and M31, to the quasi-isolated dwarfs in the outer regions of the Local Group, to the numerous isolated galaxies that are found in its surroundings. It extends to, but does not include, the galaxies associated with the next nearest groups, such as Maffei, Sculptor, and IC 342. Our basic knowledge of this important galactic subset and their resolved stellar populations will continue to improve dramatically over the coming years with existing and future observational capabilities, and they will continue to provide the most detailed information available on numerous aspects of dwarf galaxy formation and evolution. Basic observational parameters, such as distances, velocities, magnitudes, mean metallicities, as well as structural and dynamical characteristics, are collated, homogenized (as far as possible), and presented in tables that will be continually updated to provide a convenient and current online resource. As well as discussing the provenance of the tabulated values and possible uncertainties affecting their usage, the membership and spatial extent of the MW sub-group, M31 sub-group, and the Local Group are explored. The morphological diversity of the entire sample and notable sub-groups is discussed, and timescales are derived for the Local Group members in the context of their orbital/interaction histories. The scaling relations and mean stellar metallicity trends defined by the dwarfs are presented, and the origin of a possible "floor" in central surface brightness (and, more speculatively, stellar mean metallicity) at faint magnitudes is considered.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

There has been a veritable explosion of data, discoveries, and realizations in the past decade relating to the broad research area of the structure and content of nearby galaxies, particularly those in which it is possible to resolve the individual stars that contribute to the galaxy's luminosity, and to which this article exclusively refers. By necessity, the objects of the attention of this research are the Milky Way (MW), its satellites, and their neighbors within the so-called Local Group, but it is ever expanding to include the far richer sample of objects that fall within the loosely defined part of the universe referred to only as the Local Volume.

The rapid growth in interest in this field can be traced to two distinct developments that have occurred in tandem. The first is entirely observational: nearby galaxies subtend large areas on the sky and contain a high density of individually rather faint stellar sources. Thus, the advent of wide-field, digital, high-resolution, multiplexing capabilities on large telescopes (and the computational facilities to handle the resulting data-flow) has had a profound effect (and will continue to do so). The second is the result of an increased understanding of the nature of our immediate cosmic environment, facilitated in no small part by computational tools (hardware and software) that have allowed for a higher-resolution examination of the consequences of the prevailing cosmological paradigm. The discoveries and realizations that these advances have enabled are significant: the hierarchical nature of galaxy formation and the realization that some of the relics of that process may still be identifiable at z = 0; the quest for a better understanding of dark matter and the realization that the smallest galaxies nearest to us may be the darkest laboratories in the universe; the chemical evolution of the universe and the realization that the nucleosynthetic imprints of the earliest generations of star formation may still be found in the Galaxy and its satellites. These are but a few examples. A new phrase has developed to collectively describe the aforementioned research and its motivations, and which contrasts it with its higher-redshift cousins: "near-field cosmology."

The phrase "near-field cosmology" is relatively recent and dates from the Annual Review article of Freeman & Bland-Hawthorn (2002). As such, it is sometimes considered a new field within astrophysics as a whole, although this is grossly unfair to its early exponents. Indeed, what could be argued as two of the most important papers in astrophysics—those of Eggen et al. (1962) and Searle & Zinn (1978)—are quintessential examples of observational near-field cosmology. The former took a well-defined sample of stars for which high-quality spectroscopic and photometric data existed and from that motivated a collapse model for the formation of the Galaxy. The latter used complexities within the stellar populations of the Galactic globular cluster system to motivate a scenario where the Galaxy (or part of it) is built through the accretion of smaller stellar systems. The principle ideas from both papers continue to form critical aspects of modern galaxy formation theory.

Many more galaxies are now able to have their resolved stellar content examined than was possible at the time of these early studies. The content and structure of the Andromeda Galaxy (M31) provide an excellent comparison (and, increasingly, contrast) to our own Galaxy (e.g., Courteau et al. 2011; Huxor et al. 2011; Watkins et al. 2010; McConnachie et al. 2009; Yin et al. 2009; Ibata et al. 2007; Chapman et al. 2007; Kalirai et al. 2006; Koch et al. 2008, and references therein). Even excluding this prominent astronomical target, however, there are still nearly a hundred or so galaxies for which their individual stars can be brought into focus by astronomers wanting to unearth fossil signatures of galactic formation.

The overwhelming majority of nearby galaxies—indeed, the overwhelming majority of galaxies in the entire universe—are dwarf galaxies. Exactly how low a luminosity or how small a mass is required before a galaxy is deemed meager enough to warrant this popular designation is generally not well defined or accepted. Kormendy (1985) famously showed that when surface brightness is plotted against luminosity for elliptical galaxies, two relations are apparently formed, one for "giant" galaxies and one for "dwarf" galaxies. This has usually been interpreted as evidence that these systems should be separated into two distinct groups (in terms of different formation mechanisms). Some more recent studies, on the other hand, suggest that there is actually a continuous sequence connecting the dwarf and giant regimes, and that the change in the form of the scaling relations with luminosity does not necessitate fundamental differences between low- and high-luminosity systems (e.g., Graham & Guzmán 2003, and references therein). With these considerations in mind, and for the purpose of this article, I consider dwarfs to have absolute visual magnitudes fainter than MV ∼ −18 (also adopted by Grebel et al. 2003). In what follows, this places all nearby galaxies but the MW, the Large Magellanic Cloud, M31, M33, NGC 55, and NGC 300 in the dwarf category.

The closest galaxies to the MW are its own satellites, the most prominent of which are the Large and Small Magellanic Clouds (LMC/SMC), both of which are naked-eye objects. The earliest reference to the Magellanic Clouds that survives is by the Persian astronomer Abd al-Rahman al-Sufi in his 964 work, Book of Fixed Stars (which also includes the earliest recorded reference to the Andromeda Galaxy). A millennium or so later, Shapley (1938a) discovered "a large rich cluster with remarkable characteristics" in the constellation of Sculptor. He speculated that it was

  • ...a super-cluster of the globular type and of galactic dimensions; or a symmetrical Magellanic Cloud devoid of its characteristic bright stars, clusters, and luminous diffuse nebulosity; or a nearby spheroidal galaxy, highly resolved, and of abnormally low surface brightness. These phrases are merely different ways of describing the same thing and of pointing out the uniqueness of the object.

Over seventy years later, a large body of research still devotes itself to understanding the detailed properties and structural, chemical, and dynamical characteristics of this class of stellar system. The prototype described above is referred to as the Sculptor dwarf spheroidal (dSph) galaxy.

Excellent discussions of the observational properties of—and our astrophysical understanding of—the nearest (Local Group) dwarf galaxies can be found in a large number of dedicated review articles (e.g., Gallagher & Wyse 1994; Grebel 1997; Mateo 1998; van den Bergh 2000; Geisler et al. 2007; Tolstoy et al. 2009). The compilations of observational data presented in numerous works by S. van den Bergh, in particular van den Bergh (2000), and the Annual Review article by Mateo (1998) are particularly valuable resources for the astronomical community. Since their original publication, however, there have been notable advances (in part a product of the observational and theoretical factors described earlier). Not least of these is the more than doubling of the number of known galaxies in the Local Group. In the surroundings of the Local Group too, there have been significant new discoveries, and techniques and observations that used to be applicable only to the MW satellites or nearby Local Group galaxies can now be applied to these more distant relatives.

The ability to reach the resolved populations of ever more distant galaxies is important not just for the improvement in statistics that it offers, but also in terms of the increasing range of environments that it allows us to study. The overwhelming amount of detailed information regarding the stellar population properties of dwarf galaxies comes from studies of the MW satellite systems. Given the sensitivity of low-mass systems to both internal (e.g., feedback through star formation) and external (e.g., ram and tidal stripping) environmental mechanisms, this might be expected to bias our understanding toward conditions that hold only for the MW environment. Within the very nearby universe, however, there are a large number of environments able to be probed, from the surroundings of the MW and M31, to the more isolated, outlying members of the Local Group, to the very isolated dwarf galaxies that surround the Local Group, to the numerous nearby groups that populate and form our immediate neighborhood. A large number of dedicated surveys in the past few years are systematically examining galaxies in the Local Volume at a range of wavelengths, and which include among their targets some of the more distant galaxies in this sample. These surveys are obtaining precision insights into local galaxies that naturally compliment lower-resolution studies of their more numerous, but more distant, cousins (e.g., SINGS, Kennicutt et al. 2003; the Spitzer LVL, Dale et al. 2009; THINGS, Walter et al. 2008; FIGGS, Begum et al. 2008b; HST ANGRRR/ANGST, Dalcanton et al. 2009; the GALEX survey of galaxies in the Local Volume, Lee et al. 2011; the Hα survey of Kennicutt et al. 2008; and many others).

I have alluded to several motivations for pursuing studies of nearby galaxies, be they related to cosmic chemical processing, dynamical evolution, galactic formation mechanisms, environment, near-field cosmology, or a wealth of other research areas on which much can be written. It is not the purpose of this article to contribute to those discussions here; the recent Annual Review article by Tolstoy et al. (2009) provides an excellent, current overview of the status of research into many of these topics using Local Group galaxies. Rather, this article attempts to collate, homogenize, and critically review our observational understanding of aspects of the nearby populations of dwarf galaxies, with particular emphasis on the origins and uncertainties of some key observable parameters. I will discuss science topics only when such discussions relate directly to observational measurements, and those measurements, in my opinion, provide grounds for caution, or when such discussion provides a context for highlighting observational relationships between parameters. This article primarily focuses on the stellar properties of the galaxies (and derived properties, such as the implied dark matter content) and cites observations conducted mostly at ultraviolet–optical–near-infrared wavelengths, with less discussion on their gaseous and dust content. With the intense efforts currently underway to provide a more full characterization of these systems, I expect that some numbers provided in this article will be out of date by the time it is published. As such, tables presented herein will be continually updated to provide a convenient, online library of dwarf galaxy parameters for general reference and use by the community.

The layout of the article is as follows. Section 2 describes the selection of the sample and gives a general summary of my methodology and reasoning in the construction of the data set. Section 3 reviews the discovery space of the galaxies and discusses general issues relating to distances, velocities, and the membership (or otherwise) of larger-scale groups/sub-groups. Section 4 reviews the observed photometric and structural characteristics of the sample and discusses aspects of the scaling relations that they define. Section 5 reviews their masses (stellar, H i, and dynamical) and internal kinematics and discusses issues related to morphological classifications and our dynamical understanding of the sample. Section 6 presents a compilation of available mean stellar metallicities and discusses associated observational trends. Section 7 concludes and summarizes.

2. CONSTRUCTING THE DATA SET

The sample of galaxies discussed in this article consists of all known galaxies with distances determined from measurements of resolved stellar populations (usually Cepheids, RR Lyrae, tip of the red giant branch, TRGB, but also including horizontal branch level and main-sequence fitting; see Tammann et al. 2008 for a thorough discussion of the former three methods and references) that place them within 3 Mpc of the Sun. Consequently, this sample contains the satellite systems of two major galaxies (the MW and M31), the quasi-isolated, outlying members of the Local Group, and the nearby isolated dwarf galaxies that do not clearly belong to any major galaxy grouping. It is therefore expected that this sample may be valuable (as, indeed, various individual members and sub-sets have already proven valuable) for examining the role played by environment in dwarf galaxy evolution, as discussed in Section 1. There is also a practical motivation for the choice of distance limit, namely, that a much larger threshold will start to select members of nearby groups (the closest of which is the Maffei group at ∼3 Mpc, although its awkward location at low Galactic latitudes prevents accurate distances from being determined for its members; see Fingerhut et al. 2007) and would expand the scope of this article considerably.

Of course, there are no solid boundaries between galaxy groups, filaments, and the field, and some of the galaxies discussed herein are likely dynamically associated with neighboring structures like the Sculptor (e.g., Karachentsev et al. 2003a) or IC 342 (e.g., Karachentsev et al. 2003b) groups, for example. Nevertheless, at the time of writing, the adoption of these selection criteria identifies exactly 100 galaxies (not including the MW and M31), of which 73 are definite or very likely members of the Local Group. It is likely that our basic knowledge of all of these galaxies and their resolved stellar content can be improved dramatically over the coming years with existing observational capabilities, and the more distant galaxies provide stepping stones into the Local Volume for future resolved stellar population studies enabled by next-generation facilities like the 30 m class telescopes. It is with an eye to the latter considerations that the distance is not limited to the zero-velocity surface of the Local Group (RLG).

The ensuing discussion will focus on dwarf galaxies. Of the 102 galaxies that are listed in the subsequent tables, virtually no discussion will be given to the MW and M31, and numbers relating to M33, NGC 55, and NGC 300 are included for completeness only. The same caveat applies to the Magellanic Clouds, since here the research body and available data are so extensive and far exceed those that exists for most other dwarf-like systems that the reader is referred to the many review articles, books, and conferences that deal specifically with these bodies (e.g., Westerlund 1997; van Loon & Oliveira 2009, and references therein).

In compiling this catalog of galaxies, and in addition to the papers cited in each table, I have made extensive use of the NASA Extragalactic Database, the HyperLeda database (Paturel et al. 2003), Mateo (1998), van den Bergh (2000), and Karachentsev et al. (2004). The requirement that all galaxies have distances based on resolved stellar populations leads to the exclusion of a few galaxies that have low velocities that could potentially place them within 3 Mpc, but which lack any direct distance measurement (e.g., see Table 1 of Karachentsev et al. 2004). The low-latitude galaxy Dwingeloo 1 is also excluded by this criterion, since its distance of ∼2.8 Mpc is based on a Tully–Fisher estimate (Karachentsev et al. 2003b; this galaxy is very likely a member of the Maffei group). The (uncertain) TRGB measurement of MCG9-20-1 by Dalcanton et al. (2009) places it at a distance of ∼1.6 Mpc, although its radial velocity of v ∼ 954 km s−1 is very high if it is really this close, and the photometric data now seem to favor a larger distance (K. Gilbert & J. Dalcanton 2011, private communication), so it too has been excluded from the list. Finally, there are a few galaxies for which it is unclear if they lie closer or farther than 3 Mpc (e.g., UGCA 92, D = 3.01 ± 0.24 Mpc; Karachentsev et al. 2006). I exclude such galaxies, even if the quoted uncertainties would bring them within range of the distance cut. Of course, if future estimates should clearly indicate a closer distance, then the online version of the tables will be updated as appropriate.

In compiling these references, I have favored, where possible, papers that are based on analysis of the resolved stellar populations. In the interest of homogeneity and in an (ill-fated) attempt to minimize the inevitable systematic differences between measurements for different galaxies, I also generally try to limit the number of distinct publications, studies, and/or methodologies that contribute to the overall data set, favoring large surveys of multiple galaxies in preference to studies of individual galaxies. However, if there appear to be significant discrepancies in the literature over observed values, then these differences are indicated either in the footnotes and/or in the text. Infrequently, I have estimated uncertainties where none were given in the original publications.

My primary intent in compiling this data set is to provide useful information on the population of (dwarf) galaxies in the nearby universe. I have tried to ensure that the references collected herein—while not complete in any way—are generally recent enough and relevant enough that I hope they will be able to provide the reader with a starting point from which most of the germane literature can be traced.

3. BASIC PROPERTIES, POSITIONS, AND VELOCITIES

Table 1 lists basic information for all nearby galaxies that satisfy the selection criteria discussed in the previous section.

Table 1. Basic Information

(1) (2) (3) (4) (5) (6) (7) (8)
Galaxy Other Names     R.A. Decl. Original Publication Comments
        J2000 J2000    
The MW sub-group (in order of distance from the MW)
The Galaxy The MW G S(B)bc 17h45m40fs0 −29d00m28s ... Position refers to Sgr A*
Canis Major   G ???? 07h12m35fs0 −27d40m00s Martin et al. (2004a) MW disk substructure?
Sagittarius dSph   G dSph 18h55m19fs5 −30d32m43s Ibata et al. (1994) Tidally disrupting
Segue (I)   G dSph 10h07m04fs0 +16d04m55s Belokurov et al. (2007)  
Ursa Major II   G dSph 08h51m30fs0 +63d07m48s Zucker et al. (2006a)  
Bootes II   G dSph 13h58m00fs0 +12d51m00s Walsh et al. (2007)  
Segue II   G dSph 02h19m16fs0 +20d10m31s Belokurov et al. (2009)  
Willman 1 SDSS J1049+5103 G dSph 10h49m21fs0 +51d03m00s Willman et al. (2005a) Cluster?
Coma Berenices   G dSph 12h26m59fs0 +23d54m15s Belokurov et al. (2007)  
Bootes III   G dSph? 13h57m12fs0 +26d48m00s Grillmair (2009) Very diffuse. Tidal remnant?
LMC Nubecula Major G Irr 05h23m34fs5 −69d45m22s ...  
SMC Nubecula Minor G dIrr 00h52m44fs8 −72d49m43s ...  
  NGC 292            
Bootes (I)   G dSph 14h00m06fs0 +14d30m00s Belokurov et al. (2006)  
Draco UGC 10822 G dSph 17h20m12fs4 +57d54m55s Wilson (1955)  
  DDO 208            
Ursa Minor UGC 9749 G dSph 15h09m08fs5 +67d13m21s Wilson (1955)  
  DDO 199            
Sculptor   G dSph 01h00m09fs4 −33d42m33s Shapley (1938a) The prototypical dSph
Sextans (I)   G dSph 10h13m03fs0 −01d36m53s Irwin et al. (1990)  
Ursa Major (I)   G dSph 10h34m52fs8 +51d55m12s Willman et al. (2005b)  
Carina   G dSph 06h41m36fs7 −50d57m58s Cannon et al. (1977)  
Hercules   G dSph 16h31m02fs0 +12d47m30s Belokurov et al. (2007) Tidally disrupting? Remnant? Cluster?
Fornax   G dSph 02h39m59fs3 −34d26m57s Shapley (1938b)  
Leo IV   G dSph 11h32m57fs0 −00d32m00s Belokurov et al. (2007) Binary w/Leo V?
Canes Venatici II SDSS J1257+3419 G dSph 12h57m10fs0 +34d19m15s Sakamoto & Hasegawa (2006)  
            Belokurov et al. (2007)  
Leo V   G dSph 11h31m09fs6 +02d13m12s Belokurov et al. (2008) Cluster? Binary w/Leo IV?
Pisces II   G dSph 22h58m31fs0 +05d57m09s Belokurov et al. (2010) Awaiting spectr. confirmation
Canes Venatici (I)   G dSph 13h28m03fs5 +33d33m21s Zucker et al. (2006b)  
Leo II Leo B G dSph 11h13m28fs8 +22d09m06s Harrington & Wilson (1950)  
  UGC 6253            
  DDO 93            
Leo I UGC 5470 G/L dSph 10h08m28fs1 +12d18m23s Harrington & Wilson (1950)  
  DDO 74            
  Regulus Dwarf            
The M31 sub-group (in order of distance from M31)
Andromeda M31 A Sb 00h42m44fs3 +41d16m09s ...  
  NGC 224            
  UGC 454            
M32 NGC 221 A cE 00h42m41fs8 +40d51m55s Legentil (1755)  
  UGC 452            
Andromeda IX   A dSph 00h52m53fs0 +43d11m45s Zucker et al. (2004)  
NGC 205 M110 A dE/dSph 00h40m22fs1 +41d41m07s Messier (1798)  
  UGC 426            
Andromeda XVII   A dSph 00h37m07fs0 +44d19m20s Irwin et al. (2008)  
Andromeda I   A dSph 00h45m39fs8 +38d02m28s van den Bergh (1972)  
Andromeda XXVII   A dSph 00h37m27fs1 +45d23m13s Richardson et al. (2011) Tidally disrupting? Remnant?
Andromeda III   A dSph 00h35m33fs8 +36d29m52s van den Bergh (1972)  
Andromeda XXV   A dSph 00h30m08fs9 +46d51m07s Richardson et al. (2011)  
Andromeda XXVI   A dSph 00h23m45fs6 +47d54m58s Richardson et al. (2011)  
Andromeda XI   A dSph 00h46m20fs0 +33d48m05s Martin et al. (2006)  
Andromeda V   A dSph 01h10m17fs1 +47d37m41s Armandroff et al. (1998)  
Andromeda X   A dSph 01h06m33fs7 +44d48m16s Zucker et al. (2007)  
Andromeda XXIII   A dSph 01h29m21fs8 +38d43m08s Richardson et al. (2011)  
Andromeda XX   A dSph 00h07m30fs7 +35d07m56s McConnachie et al. (2008)  
Andromeda XII   A/L dSph 00h47m27fs0 +34d22m29s Martin et al. (2006) Unbound to M31?
NGC 147 UGC 326 A dE/dSph 00h33m12fs1 +48d30m32s Herschel (1833) Binary w/NGC 185? Tidal stream
  DDO 3            
Andromeda XXI   A dSph 23h54m47fs7 +42d28m15s Martin et al. (2009)  
Andromeda XIV   A/L dSph 00h51m35fs0 +29d41m49s Majewski et al. (2007) Unbound to M31? In Pisces
Andromeda XV   A dSph 01h14m18fs7 +38d07m03s Ibata et al. (2007)  
Andromeda XIII   A dSph 00h51m51fs0 +33d00m16s Martin et al. (2006) In Pisces
Andromeda II   A dSph 01h16m29fs8 +33d25m09s van den Bergh (1972) In Pisces
NGC 185 UGC 396 A dE/dSph 00h38m58fs0 +48d20m15s Herschel (1789) Binary w/NGC 147?
Andromeda XXIX   A dSph 23h58m55fs6 +30d45m20s Bell et al. (2011) In Pegasus
Andromeda XIX   A dSph 00h19m32fs1 +35d02m37s McConnachie et al. (2008)  
Triangulum M33 A Sc 01h33m50fs9 +30d39m37s a  
  NGC 598            
  UGC 1117            
Andromeda XXIV   A dSph 01h18m30fs0 +46d21m58s Richardson et al. (2011)  
Andromeda VII Casseopia dSph A dSph 23h26m31fs7 +50d40m33s Karachentsev & Karachentseva (1999)  
Andromeda XXII   A dSph 01h27m40fs0 +28d05m25s Martin et al. (2009) M33 satellite? In Pisces
IC 10 UGC 192 A dIrr 00h20m17fs3 +59d18m14s Swift (1888)  
LGS 3 (Local Group Pisces (I) A dIrr/dSph 01h03m55fs0 +21d53m06s Karachentseva (1976)  
Suspect 3)           Kowal et al. (1978)  
Andromeda VI Pegasus dSph A dSph 23h51m46fs3 +24d34m57s Karachentsev & Karachentseva (1999)  
            Armandroff et al. (1999)  
Andromeda XVI   A/L dSph 00h59m29fs8 +32d22m36s Ibata et al. (2007) In Pisces
The rest of the Local Group and its neighbors (in order of distance from the barycenter of the Local Group)
Andromeda XXVIII   A/L dSph? 22h32m41fs2 +31d12m58s Slater et al. (2011) In Pegasus
IC 1613 DDO 8 L dIrr 01h04m47fs8 +02d07m04s Wolf (1906)  
  UGC 668            
Phoenix   L/G dIrr/dSph 01h51m06fs3 −44d26m41s Schuster & West (1976)b  
NGC 6822 IC 4895 L/G dIrr 19h44m56fs6 −14d47m21s Barnard (1884) Polar ring morph.
  DDO 209            
  Barnard's Galaxy            
Cetus   L dSph 00h26m11fs0 −11d02m40s Whiting et al. (1999) Isolated dSph
Pegasus dIrr UGC 12613 L/A dIrr/dSph 23h28m36fs3 +14d44m35s A. G. Wilson. Reported in Holmberg (1958)  
  DDO 216            
Leo T   L/G dIrr/dSph 09h34m53fs4 +17d03m05s Irwin et al. (2007)  
WLM (Wolf-Lundmark- UGCA 444 L dIrr 00h01m58fs2 −15d27m39s Wolf (1910) Defines RLG
-Melotte) DDO 221         Melotte (1926), including K. E. Lundmark  
Leo A Leo III L dIrr 09h59m26fs5 +30d44m47s Zwicky (1942) Defines RLG
  UGC 5364            
  DDO 69            
Andromeda XVIII   L dSph 00h02m14fs5 +45d05m20s McConnachie et al. (2008) Isolated dSph
Aquarius DDO 210 L dIrr/dSph 20h46m51fs8 −12d50m53s van den Bergh (1959) Defines RLG
Tucana   L dSph 22h41m49fs6 −64d25m10s Lavery (1990)c Isolated dSph; defines RLG
Sagittarius dIrr UKS 1927-177 L dIrr 19h29m59fs0 −17d40m41s Cesarsky et al. (1977) Defines RLG
            Longmore et al. (1978)  
UGC 4879 VV 124 L dIrr/dSph 09h16m02fs2 +52d50m24s Kopylov et al. (2008)d Defines RLG
NGC 3109 DDO 236 N dIrr 10h03m06fs9 −26d09m35s Herschel (1847) Loose NGC 3109 subgroup. Interacting with Antlia?
  UGCA 194            
Sextans B UGC 5373 N dIrr 10h00m00fs1 +05d19m56s A. G. Wilson? See Holmberg (1958)e Loose NGC 3109 subgroup
  DDO 70            
Antlia   N dIrr 10h04m04fs1 −27d19m52s Whiting et al. (1997)f Loose NGC 3109 subgroup. Interacting with NGC 3109?
Sextans A UGCA 205 N dIrr 10h11m00fs8 −04d41m34s Zwicky (1942) Loose NGC 3109 sub-group
  DDO 75            
HIZSS 3(A)   N (d)Irr? 07h00m29fs3 −04d12m30s Henning et al. (2000)g Zone of obscuration. Binary with B?
HIZSS 3B   N (d)Irr? 07h00m29fs3 −04d12m30s Henning et al. (2000)h Zone of obscuration. Binary with A?
KKR 25   N dIrr/dSph 16h13m48fs0 +54d22m16s Karachentsev et al. (2001b)  
ESO 410- G 005 UKS 0013-324 N dIrr/dSph 00h15m31fs6 −32d10m48s Lauberts (1982) NGC 55 sub-group, forming "bridge" to Sculptor?
            Longmore et al. (1982)  
NGC 55   N Irr 00h14m53fs6 −39d11m48s Dunlop (1828) NGC 55 sub-group, forming "bridge" to Sculptor?
ESO 294- G 010   N dIrr/dSph 00h26m33fs4 −41d51m19s Lauberts (1982) NGC 55 sub-group, forming "bridge" to Sculptor?
NGC 300   N Sc 00h54m53fs5 −37d41m04s Dunlop (1828) NGC 55 sub-group, forming "bridge" to Sculptor?
IC 5152   N dIrr 22h02m41fs5 −51d17m47s D. Stewart. Reported in Pickering (1899)  
KKH 98   N dIrr 23h45m34fs0 +38d43m04s Karachentsev et al. (2001a)  
UKS 2323-326 UGCA 438 N dIrr 23h26m27fs5 −32d23m20s Longmore et al. (1978) NGC 55 sub-group, forming "bridge" to Sculptor?
KKR 3 KK 230 N dIrr 14h07m10fs5 +35d03m37s Karachentseva & Karachentsev (1998)i Member of the Canes Venatici I cloud
GR 8 UGC 8091 N dIrr 12h58m40fs4 +14d13m03s Reaves (1956)j Gibson Reaves gave this galaxy his initials
  VV 558            
  DDO 155            
UGC 9128 DDO 187 N dIrr 14h15m56fs5 +23d03m19s van den Bergh (1959)  
UGC 8508   N dIrr 13h30m44fs4 +54d54m36s Vorontsov-Vel'Yaminov & Krasnogorskaya (1962)  
IC 3104 ESO 020- G 004 N dIrr 12h18m46fs0 −79d43m34s D. Stewart. Reported in Pickering (1908)  
  UKS 1215-794            
DDO 125 UGC 7577 N dIrr 12h27m40fs9 +43d29m44s van den Bergh (1959)  
UGCA 86   N dIrr 03h59m48fs3 +67d08m19s Nilson (1974) Companion of IC 342
DDO 99 UGC 6817 N dIrr 11h50m53fs0 +38d52m49s van den Bergh (1959)  
IC 4662 ESO 102- G 014 N dIrr 17h47m08fs8 −64d38m30s R. Innes. Reported in Lunt (1902)  
DDO 190 UGC 9240 N dIrr 14h24m43fs4 +44d31m33s van den Bergh (1959)  
KKH 86   N dIrr 13h54m33fs5 +04d14m35s Karachentsev et al. (2001a)  
NGC 4163 UGC 7199 N dIrr 12h12m09fs1 +36d10m09s Herschel (1789)  
  NGC 4167k            
DDO 113 UGCA 276 N dIrr 12h14m57fs9 +36d13m08s van den Bergh (1959)  
  KDG 90            

Notes. aDiscovery credited to G. B. Hodierna (see Steinicke 2010). bOriginally thought to be a globular cluster. Canterna & Flower (1977) subsequently identified its galactic nature. cThis author first suggested that this object is a member of the Local Group, but it had previously been cataloged by Corwin et al. (1985). dThese authors calculate a new distance to UGC 4879 and show that it is on the periphery of the Local Group. While included in the "Atlas and Catalog of Interacting Galaxies" by Vorontsov-Velyaminov (1959), earlier references to this object have not been found. eHolmberg (1958) credits Wilson (1955) with discovery, but this object is not listed in this manuscript. fThese authors rediscovered this galaxy and showed that it was in the periphery of the Local Group, but it had earlier been cataloged by Corwin et al. (1985), Feitzinger & Galinski (1985), and Arp & Madore (1987). gHIZSS 3 resolved into two sources by Begum et al. (2005). Position corresponds to the "HIZSS3 system." hHIZSS 3 resolved into two sources by Begum et al. (2005). Position corresponds to the "HIZSS3 system." iSee also Karachentseva et al. (1999). jDiscovered on inspection of photographic plates from a survey discussed in Shane (1947). C. D. Shane speculated that some of the nebulae that were visible could be Local Group dwarf galaxies. kThe original coordinates of NGC 4167 are coincident with the coordinates of NGC 4163, but Sulentic & Tifft (1973) note that NGC 4167 is "non-existent."

Download table as:  ASCIITypeset images: 1 2 3 4

Column 1. Galaxy name.

Column 2. Common alternative names.

Column 3. Indicator whether they are associated with MW [G], M31 [A], the Local Group [L], or are nearby neighbors [N].

Column 4. Morphological (Hubble) type. The distinction between dwarf elliptical (dE) and dwarf spheroidal (dSph) is based solely on luminosity and is therefore somewhat arbitrary. Objects fainter than MV ∼ −18 are given a "d" (dwarf) prefix and are again somewhat arbitrary.

Columns 5 and 6. Celestial coordinates (J2000).

Column 7. Original publication. For more details relating to the discoveries of the NGC/IC galaxies, the reader is referred to Steinicke (2010).1

Column 8. Comments.

Table 2 lists position and velocity information for the galaxy sample.

Table 2. Positions and Velocities

(1) (2) (3) (4) (5) (6) (7) (8) (9) (10) (11)
Galaxy l b E(BV) (mM)o D v [DMW,vMW] [DM31,vM31] [DLG,vLG] Referencesa
          (kpc) (km s−1) (kpc, km s−1) (kpc, km s−1) (kpc, km s−1)  
The Galaxy                          
Canis Majorb 240.0 −8.0 0.264 14.29 0.30 7 1 87.0c 4.0 [13, −117] [786, −185] [394, −172] (1) (2)
Sagittarius dSph 5.6 −14.2 0.153 17.10 0.15 26 2 140.0 2.0 [18, 169] [791, 149] [400, 159] (3) (4)
Segue (I) 220.5 +50.4 0.031 16.80 0.20 23 2 208.5 0.9 [28, 113] [792, 56] [401, 69] (5) (6)
Ursa Major II 152.5 +37.4 0.094 17.50 0.30 32 4 −116.5 1.9 [38, −33] [771, 11] [380, −3] (7) (8)
Bootes II 353.7 +68.9 0.031 18.10 0.06 42 1 −117.0 5.2 [40, −116] [807, −174] [416, −156] (9) (10)
Segue II 149.4 −38.1 0.185 17.70 0.10 35 2 −39.2 2.5 [41, 43] [753, 141] [361, 112] (11)
Willman 1 158.6 +56.8 0.014 17.90 0.40 38 7 −12.3 2.5 [43, 35] [780, 44] [390, 41] (12) (13)
Coma Berenices 241.9 +83.6 0.017 18.20 0.20 44 4 98.1 0.9 [45, 82] [802, 32] [412, 46] (5) (8)
Bootes III 35.4 +75.4 0.021 18.35 0.1 47 2 197.5 3.8 [46, 240] [800, 207] [410, 219] (14) (15)
LMC 280.5 −32.9 0.926 18.52 0.09 51 2 262.2d 3.4 [50, 68] [811, −6] [421, 12] (16)e (17)
SMC 302.8 −44.3 0.419 19.03 0.12 64 4 145.6f 0.6 [61, 5] [811, −48] [422, −35] (18)g (19)
Bootes (I) 358.1 +69.6 0.017 19.11 0.08 66 2 99.0 2.1 [64, 106] [819, 51] [430, 69] (20) (13)
Draco 86.4 +34.7 0.027 19.40 0.17 76 6 −291.0 0.1 [76, −96] [754, −32] [366, −46] (21) (22)
Ursa Minor 105.0 +44.8 0.032 19.40 0.10 76 3 −246.9 0.1 [78, −85] [758, −31] [370, −44] (23)h (24)
Sculptor 287.5 −83.2 0.018 19.67 0.14 86 6 111.4 0.1 [86, 79] [765, 106] [379, 97] (25) (26)
Sextans (I) 243.5 +42.3 0.047 19.67 0.10 86 4 224.2 0.1 [89, 72] [839, −14] [450, 6] (27)i (26)
Ursa Major (I) 159.4 +54.4 0.020 19.93 0.10 97 4 −55.3 1.4 [102, −7] [777, 4] [391, 0] (28) (8)
Carina 260.1 −22.2 0.061 20.11 0.13 105 6 222.9 0.1 [107, 7] [841, −70] [454, −53] (29) (26)
Hercules 28.7 +36.9 0.063 20.60 0.20 132 12 45.2 1.1 [126, 145] [826, 133] [444, 141] (30) (31)
Fornax 237.1 −65.7 0.021 20.84 0.18 147 12 55.3 0.1 [149, −33] [772, −24] [394, −30] (29) (26)
Leo IV 264.4 +57.4 0.026 20.94 0.09 154 6 132.3 1.4 [155, 13] [899, −78] [513, −54] (32) (8)
Canes Venatici II 113.6 +82.7 0.010 21.02 0.06 160 4 −128.9 1.2 [161, −95] [837, −122] [458, −113] (33) (8)
Leo V 261.9 +58.5 0.027 21.25 0.12 178 10 173.3 3.1 [179, 59] [915, −30] [530, −7] (34)
Pisces II 79.2 −47.1 0.065 21.30:   182:   ... ... [181, —] [660, —] [285, —] (35)
Canes Venatici (I) 74.3 +79.8 0.014 21.69 0.10 218 10 30.9 0.6 [218, 78] [864, 54] [493, 62] (36) (8)
Leo II 220.2 +67.2 0.017 21.84 0.13 233 14 78.0 0.1 [236, 24] [901, −29] [529, −16] (37)j (22)
Leo I 226.0 +49.1 0.036 22.02 0.13 254 15 282.5 0.1 [258, 174] [922, 111] [551, 125] (38) (39)
Andromeda 121.2 −21.6 0.684 24.47 0.07 783 25 −300.0 4.0 [787, −122] [0, 0] [392, −34] (40) (41)
M32 121.2 −22.0 0.154 24.53 0.21 805 78 −199.0 6.0 [809, −22] [23, 100] [414, 67] (42) (43)
Andromeda IX 123.2 −19.7 0.075 24.42 0.07 766 25 −207.7 2.7 [770, −32] [40, 90] [375, 56] (40) (44)
NGC 205 120.7 −21.1 0.085 24.58 0.07 824 27 −246.0 1.0 [828, −67] [42, 56] [433, 22] (40) (45)
Andromeda XVII 120.2 −18.5 0.074 24.50 0.10 794 37 ... ... [798, —] [45, —] [404, —] (46)
Andromeda I 121.7 −24.8 0.053 24.36 0.07 745 24 −375.8 1.4 [749, −204] [58, −82] [355, −116] (40) (47)
Andromeda XXVII 120.4 −17.4 0.080 24.59 0.12 828 46 ... ... [832, —] [75, —] [439, —] (48)
Andromeda III 119.4 −26.3 0.056 24.37 0.07 748 24 −345.6 1.8 [752, −171] [75, −49] [360, −83] (40) (47)
Andromeda XXV 119.2 −15.9 0.101 24.55 0.12 813 45 ... ... [817, —] [88, —] [426, —] (48)
Andromeda XXVI 118.1 −14.7 0.110 24.41 0.12 762 42 ... ... [766, —] [103, —] [377, —] (48)
Andromeda XI 121.7 −29.1 0.080 24.40 +0.2−0.5 759 +70−175 −419.6 4.4 [763, −255] [104, −134] [374, −167] (44)
Andromeda V 126.2 −15.1 0.125 24.44 0.08 773 28 −403.0 4.0 [778, −229] [110, −109] [389, −143] (40) (49)
Andromeda X 125.8 −18.0 0.129 24.23 0.21 701 68 −163.8 1.2 [706, 8] [110, 128] [314, 95] (50) (47)
Andromeda XXIII 131.0 −23.6 0.066 24.43 0.13 769 46 ... ... [774, —] [127, —] [388, —] (48)
Andromeda XX 112.9 −26.9 0.059 24.52 +0.74−0.24 802 +273−89 ... ... [805, —] [129, —] [420, —] (51)
Andromeda XII 122.0 −28.5 0.110 24.70 0.30 871 120 −558.4 3.2 [875, −393] [133, −272] [485, −306] (44)
NGC 147 119.8 −14.3 0.172 24.15 0.09 676 28 −193.1 0.8 [680, −4] [142, 118] [292, 85] (40) (52)
Andromeda XXI 111.9 −19.2 0.093 24.67 0.13 859 51 ... ... [862, —] [150, —] [476, —] (53)
Andromeda XIV 123.0 −33.2 0.060 24.33 0.33 735 112 −481.0 2.0 [739, −326] [162, −208] [361, −240] (54) (47)
Andromeda XV 127.9 −24.5 0.047 24.00 0.20 631 58 −339.0 7.0 [636, −180] [174, −61] [247, −95] (55) (56)
Andromeda XIII 123.0 −29.9 0.082 24.80 +0.1−0.4 912 +42−168 −195.0 8.4 [916, −34] [180, 86] [528, 53] (44)
Andromeda II 128.9 −29.2 0.061 24.07 0.06 652 18 −193.6 1.0 [657, −44] [184, 73] [276, 40] (40) (47)
NGC 185 120.8 −14.5 0.184 23.95 0.09 617 26 −203.8 1.1 [621, −17] [187, 105] [234, 72] (40) (52)
Andromeda XXIX 109.8 −30.8 0.046 24.32 0.22 731 74 ... ... [734, —] [188, —] [363, —] (57)
Andromeda XIX 115.6 −27.4 0.062 24.85 0.13 933 56 ... ... [936, —] [189, —] [548, —] (51)
Triangulumk 133.6 −31.3 0.041 24.54 0.06 809 22 −179.2 1.7 [814, −45] [206, 69] [442, 37] (40) (58)
Andromeda XXIV 127.8 −16.3 0.083 23.89 0.12 600 33 ... ... [605, —] [208, —] [220, —] (48)
Andromeda VII 109.5 −10.0 0.194 24.41 0.10 762 35 −309.4 2.3 [765, −98] [218, 24] [401, −8] (40) (47)
Andromeda XXII 132.6 −34.1 0.079 24.50:l   794:   ... ... [799, —] [221, —] [432, —] (53)
IC 10 119.0 −3.3 1.568 24.50 0.12 794 44 −348.0 1.0 [798, −150] [252, −32] [440, −64] (59) (43)
LGS 3 126.8 −40.9 0.040 24.43 0.07 769 25 −286.5 0.3 [773, −155] [269, −43] [422, −74] (40) (60)
Andromeda VI 106.0 −36.3 0.063 24.47 0.07 783 25 −354.0 3.0 [785, −180] [269, −62] [435, −93] (40) (49)
Andromeda XVI 124.9 −30.5 0.067 23.60 0.20 525 48 −385.0 5.0 [529, −229] [279, −110] [153, −143] (55) (56)
Andromeda XXVIII 91.0 −22.9 0.096 24.10 +0.5−0.2 661 +152−61 ... ... [661, —] [368, —] [364, —] (61)
IC 1613 129.7 −60.6 0.025 24.39 0.12 755 42 −233.0 1.0 [758, −154] [520, −64] [517, −90] (62) (63)
Phoenix 272.2 −68.9 0.016 23.09 0.10 415 19 −13.0m 9.0 [415, −103] [868, −104] [556, −106] (64) (65)
NGC 6822 25.4 −18.4 0.231 23.31 0.08 459 17 −57.0 2.0 [452, 43] [897, 65] [595, 64] (66) (67)n
Cetus 101.5 −72.9 0.029 24.39 0.07 755 24 −87.0 2.0 [756, −27] [681, 46] [603, 26] (40) (68)
Pegasus dIrr 94.8 −43.6 0.068 24.82 0.07 920 30 −183.3 5.0 [921, −21] [474, 89] [618, 60] (40) (69)o
Leo T 214.9 +43.7 0.031 23.10 0.10 417 19 38.1 2.0 [422, −58] [991, −108] [651, −98] (70) (8)
WLM 75.9 −73.6 0.038 24.85 0.08 933 34 −130.0 1.0 [932, −73] [836, −6] [794, −24] (40) (71)
Leo A 196.9 +52.4 0.021 24.51 0.12 798 44 22.3 2.9 [803, −19] [1200, −46] [941, −41] (72) (73)
Andromeda XVIII 113.9 −16.9 0.106 25.66 0.13 1355 81 ... ... [1358, —] [591, —] [969, —] (51)
Aquarius 34.0 −31.3 0.051 25.15 0.08 1072 39 −140.7 2.5 [1066, −27] [1173, 17] [1053, 9] (40) (69)o
Tucana 322.9 −47.4 0.031 24.74 0.12 887 49 194.0p 4.3 [882, 99] [1355, 62] [1076, 73] (74) (75)
Sagittarius dIrr 21.1 −16.3 0.124 25.14 0.18 1067 88 −78.5 1.0 [1059, 8] [1356, 20] [1156, 22] (76) (60)q
UGC 4879 164.7 +42.9 0.016 25.67 0.04 1361 25 −70.0r 15.0 [1367, −27] [1395, −4] [1321, −12] (77) (78)
NGC 3109 262.1 +23.1 0.067 25.57 0.08 1300 48 403.0 2.0 [1301, 193] [1987, 83] [1633, 110] (79) (67)
Sextans B 233.2 +43.8 0.031 25.77 0.03 1426 20 304.0 1.0 [1430, 171] [1943, 97] [1659, 114] (59) (63)
Antlia 263.1 +22.3 0.079 25.65 0.10 1349 62 362.0 2.0 [1350, 151] [2039, 39] [1684, 66] (59) (80)
Sextans A 246.1 +39.9 0.045 25.78 0.08 1432 53 324.0 2.0 [1435, 163] [2027, 73] [1711, 94] (59) (67)
HIZSS 3(A) 217.7 +0.1 1.013 26.12 0.14 1675 108 288.0 2.5 [1682, 139] [1923, 105] [1760, 109] (81)s (82)
HIZSS 3B 217.7 +0.1 1.013 26.12 0.14 1675 108 322.6 1.4 [1682, 174] [1923, 140] [1760, 143] (81)s (82)
KKR 25 83.9 +44.4 0.009 26.40 0.07 1905 61 −139.5 1.0 [1904, 31] [1853, 77] [1838, 67] (59) (83)
ESO 410- G 005 357.8 −80.7 0.014 26.42 0.04 1923 35 ... ... [1922, —] [1862, —] [1852, —] (59)
NGC 55 332.9 −75.7 0.013 26.43 0.12 1932 107 129.0 2.0 [1930, 98] [1964, 117] [1909, 111] (84) (67)
ESO 294- G 010 320.4 −74.4 0.006 26.54 0.04 2032 37 117.0 5.0 [2030, 72] [2090, 85] [2023, 81] (59) (85)
NGC 300 299.2 −79.4 0.013 26.59 0.06 2080 57 146.0 2.0 [2079, 103] [2078, 122] [2042, 116] (59) (67)
IC 5152 343.9 −50.2 0.025 26.45 0.05 1950 45 122.0 2.0 [1945, 81] [2211, 68] [2047, 73] (59) (67)
KKH 98 109.1 −22.4 0.123 27.01 0.09 2523 105 −136.9 1.0 [2526, 60] [1762, 184] [2140, 152] (59) (83)
UKS 2323-326 11.9 −70.9 0.015 26.72 0.09 2208 92 62.0 5.0 [2205, 74] [2153, 107] [2145, 99] (59) (86)
KKR 3 63.7 +72.0 0.014 26.70 0.12 2188 121 63.3 1.8 [2187, 135] [2464, 121] [2297, 127] (59) (87)
GR 8 310.7 +77.0 0.026 26.69 0.12 2178 120 213.9 2.5 [2177, 182] [2699, 117] [2421, 136] (59) (69)o
UGC 9128 25.6 +70.5 0.023 26.80 0.04 2291 42 152.0 1.0 [2288, 195] [2686, 159] [2465, 172] (59) (63)
UGC 8508 111.1 +61.3 0.015 27.06 0.03 2582 36 56.0 5.0 [2583, 165] [2609, 184] [2566, 181] (88) (89)
IC 3104 301.4 −17.0 0.410 26.78 0.18 2270 188 429.0 4.0 [2266, 242] [2923, 144] [2588, 170] (90) (67)
DDO 125 137.8 +72.9 0.020 27.06 0.05 2582 59 194.9 0.2 [2584, 245] [2764, 237] [2646, 240] (88) (83)
UGCA 86 139.8 +10.6 0.938 27.36 0.17 2965 232 67.0 4.0 [2971, 209] [2387, 302] [2663, 275] (91) (41)
DDO 99 166.2 +72.7 0.026 27.07 0.14 2594 167 251.0 4.0 [2596, 272] [2824, 252] [2683, 257] (88) (89)
IC 4662 328.5 −17.8 0.070 26.94 0.17 2443 191 302.0 3.0 [2436, 192] [3027, 116] [2722, 139] (91) (67)
DDO 190 82.0 +64.5 0.012 27.23 0.03 2793 39 150.0 4.0 [2793, 256] [2917, 263] [2829, 263] (88) (92)
KKH 86 339.0 +62.6 0.027 27.06 0.16 2582 190 287.2 0.7 [2578, 259] [3157, 186] [2856, 209] (88) (83)
NGC 4163 163.2 +77.7 0.020 27.28 0.03 2858 39 165.0 5.0 [2860, 184] [3119, 159] [2966, 166] (88) (93)
DDO 113 161.1 +78.1 0.020 27.35 0.06 2951 82 284.0 6.0 [2953, 305] [3210, 280] [3058, 287] (88) (41)

aReferences: (1) Bellazzini et al. 2006; (2) Martin et al. 2005; (3) Monaco et al. 2004; (4) Ibata et al. 1994; (5) Belokurov et al. 2007; (6) Simon et al. 2011; (7) Zucker et al. 2006a; (8) Simon & Geha 2007; (9) Walsh et al. 2008; (10) Koch et al. 2009; (11) Belokurov et al. 2009; (12) Willman et al. 2006; (13) Martin et al. 2007; (14) Grillmair 2009; (15) Carlin et al. 2009; (16) Clementini et al. 2003; (17) van der Marel et al. 2002; (18) Udalski et al. 1999; (19) Harris & Zaritsky 2006; (20) Dall'Ora et al. 2006; (21) Bonanos et al. 2004; (22) Walker et al. 2007; (23) Carrera et al. 2002; (24) Walker et al. 2009c; (25) Pietrzyński et al. 2008; (26) Walker et al. 2009b; (27) Lee et al. 2009; (28) Okamoto et al. 2008; (29) Pietrzyński et al. 2009; (30) Coleman et al. 2007; (31) Adén et al. 2009a; (32) Moretti et al. 2009; (33) Greco et al. 2008; (34) Belokurov et al. 2008; (35) Belokurov et al. 2010; (36) Martin et al. 2008a; (37) Bellazzini et al. 2005; (38) Bellazzini et al. 2004b; (39) Mateo et al. 2008; (40) McConnachie et al. 2005; (41) de Vaucouleurs et al. 1991; (42) Fiorentino et al. 2010; (43) Huchra et al. 1999; (44) Collins et al. 2010; (45) Geha et al. 2006b; (46) Irwin et al. 2008; (47) Kalirai et al. 2010; (48) Richardson et al. 2011; (49) Evans et al. 2000; (50) Zucker et al. 2007; (51) McConnachie et al. 2008; (52) Geha et al. 2010; (53) Martin et al. 2009; (54) Majewski et al. 2007; (55) Ibata et al. 2007; (56) Letarte et al. 2009; (57) Bell et al. 2011; (58) Corbelli & Schneider 1997; (59) Tully et al. 2006; (60) Young & Lo 1997b; (61) Slater et al. 2011; (62) Bernard et al. 2010; (63) Hoffman et al. 1996; (64) Hidalgo et al. 2009; (65) Irwin & Tolstoy 2002; (66) Gieren et al. 2006; (67) Koribalski et al. 2004; (68) Lewis et al. 2007; (69) Young et al. 2003; (70) Irwin et al. 2007; (71) Leaman et al. 2009; (72) Dolphin et al. 2002; (73) Brown et al. 2007; (74) Bernard et al. 2009; (75) Fraternali et al. 2009; (76) Momany et al. 2002; (77) Jacobs et al. 2011; (78) Kopylov et al. 2008; (79) Soszyński et al. 2006; (80) Barnes & de Blok 2001; (81) Silva et al. 2005; (82) Begum et al. 2005; (83) Huchtmeier et al. 2003; (84) Gieren et al. 2008; (85) Jerjen et al. 1998; (86) da Costa et al. 1998; (87) Begum et al. 2006; (88) Dalcanton et al. 2009; (89) Begum et al. 2008b; (90) Karachentsev et al. 2002b; (91) Karachentsev et al. 2006; (92) Haynes et al. 1998; (93) Huchra et al. 1995. bSignificant distance/velocity gradients. cCorresponding to RC stars at 6 < D < 7 kpc. dCarbon stars. eSee their Figure 8 for an excellent summary. fRed giant branch stars. gDistance calculated using their LMC–SMC offset of 0.51 ± 0.03 for the adopted LMC distance modulus. hAlso Bellazzini et al. (2002), but note that these estimates are ∼0.2 mag larger than some previous values. icf. Dolphin (2002), which also estimates distances by comparing real and model CMDs. jNote this estimate is ∼0.2 mag larger than some previous values. kRange in distance estimates from 750 ≲ D ≲ 950. lAssumed to be at M31–M33 distance. mAlternative value in Gallart et al. (2001). nSee Demers et al. (2006) for velocities of 110 C-stars. oUncertainty due to intrinsic asymmetries is gas distribution, L. Young (2010, private communication). pOptical velocity suggests that Tucana is not associated with nearby H i cloud. qSee also Begum et al. (2006). rOptical spectroscopy suggests that vh ∼ −70 km s−1; however, H i analysis by Bellazzini et al. (2011) gives vh ∼ −25 km s−1. No clear reason for the discrepancy currently exists. sDistance calculated prior to realization that HIZSS 3 was two systems, and will most likely correspond more closely to the more massive A system.

Download table as:  ASCIITypeset images: 1 2 3

Column 1. Galaxy name.

Columns 2 and 3. Galactic coordinates (l, b).

Column 4. Foreground extinction, E(B − V): these correspond to values derived by Schlegel et al. (1998) from an all-sky COBE/DIRBE and IRAS/ISSA 100 μm map, at the coordinates of the centroid of each galaxy. While Schlegel et al. (1998) estimate that values for E(B − V) are generally accurate to 16%, a major (and sometimes overlooked, including by this author!) caveat for their use in the direction of nearby (large) galaxies is that the brightest galaxies at IRAS 60 μm2 were excised from these maps and replaced by the "most likely value" of the underlying 100 μm emission (see Section 4.2 of Schlegel et al. 1998). The resulting values are not, therefore, directly observed, and any variation across the area occupied by the galaxy is not traced by these maps. For this compilation, the affected galaxies are NGC 6822, IC 1613, M33, NGC 205, NGC 3109, NGC 55, and NGC 300. The LMC, SMC, and M31 (and by necessity also M32) were not excised.

Columns 5 and 6. Distance modulus, heliocentric distance: all distance moduli are based on resolved stellar population analysis (generally Cepheid, RR Lyrae, or TRGB measurements, but also horizontal branch level and main-sequence fitting). Note that many TRGB estimates do not include the formal uncertainty in the absolute magnitude of the TRGB (Bellazzini et al. 2001, 2004a; Rizzi et al. 2007).

Column 7. Heliocentric velocity: for most Local Group galaxies, the velocity corresponds to the optical velocity of the galaxy. For some of the more distant galaxies, velocities are measured from H i.

Columns 8–10. Distance–velocity pairs for the Galactocentric (MW), M31, and Local Group frames of reference: these are calculated by first assuming that the barycenter of the Local Group is located at the mid point of the vector connecting the MW and M31 (for an adopted M31 distance modulus of (mM)0 = 24.47; McConnachie et al. 2005). The latter galaxy is more luminous (e.g., Hammer et al. 2007, and references therein), but dynamical mass estimates for both of these galaxies have produced conflicting results regarding which is more massive (e.g., Little & Tremaine 1987; Kochanek 1996; Wilkinson & Evans 1999; Evans & Wilkinson 2000; Evans et al. 2000; Klypin et al. 2002; Ibata et al. 2004; Geehan et al. 2006; Watkins et al. 2010; McMillan 2011), a result nearly exclusively due to the lack of a sufficient number of dynamical tracers at large radii. Given this uncertainty, I conclude that assuming their masses to be roughly the same is reasonable, simple, and convenient. Velocities are converted to the Local Group frame using the solar apex derived by Karachentsev & Makarov (1996) and subtracting the component of the MW's velocity in the direction of each dwarf. M31-centric velocities are obtained by subtracting the component of the Local Group centric velocity of M31 in the direction of each dwarf. Galactocentric distances and velocities are calculated in the usual way, for an adopted Galactic rotation velocity at the Sun of 220 km s−1 at a radius of 8.5 kpc from the Galactic center.

Column 11. References.

Figure 1 shows Aitoff projections of the Galactic coordinates of all galaxies in the sample. The top panel shows only those objects identified as definite or likely MW satellites, the middle panel shows definite or likely M31 sub-group members (blue points) and quasi-isolated Local Group galaxies (green points), and the bottom panel shows galaxies surrounding the Local Group within a distance of 3 Mpc. Some confirmed members of all galaxy groups within 5 Mpc are also shown in the bottom panel (gray points) to highlight the location of the Local Group and its neighbors with respect to these nearby structures (Karachentsev 2005). In this respect, it is important to highlight the work of I. Karachentsev, B. Tully, and their colleagues in both the identification of nearby galaxies and groups and the determination of accurate distances to many of our neighbors in order to understand the structure of the Local Volume (e.g., see Karachentsev & Tikhonov 1994; Karachentsev et al. 2003a, 2004; Tully 1987; Tully et al. 2006, 2009).

Figure 1.

Figure 1. Aitoff projections of the Galactic coordinates of MW galactic satellites (top panel); the M31 sub-group (blue) and isolated Local Group galaxies (green; middle panel); the nearest galaxies to the Local Group that have distances based on resolved stellar populations that place them within 3 Mpc (magenta; bottom panel). The positions of nearby galaxy groups are indicated in gray in the bottom panel.

Standard image High-resolution image

3.1. Membership of the MW Sub-group

Figure 2 shows Galactocentric velocity versus distance for all galaxies within 600 kpc for which necessary data exist. Here, the Galactocentric velocity listed in Table 2 has been multiplied by a factor of $\sqrt{3}$ to provide a crude estimate of the unknown tangential velocities of these galaxies. Dashed curves show the escape velocity from a 1012M point mass (a reasonable approximation to the escape velocity curve of the MW for the purposes of this discussion). The vertical dashed line indicates the approximate cosmological virial radius of the MW, Rvir ∼ 300 kpc (Klypin et al. 2002; these authors define Rvir as the radius within which the mean density of the model dark matter halo is a factor Δ ≃ 340 larger than the critical density of the universe).

Figure 2.

Figure 2. Galactocentric velocity vs. distance for all galaxies in the proximity of the MW. The y-ordinate has been multiplied by a factor of $\sqrt{3}$ to account for the unknown tangential motions of the galaxies. Dashed curves show the escape velocity from a 1012M point mass. The vertical dashed line indicates the approximate location of the expected cosmological virial radius of the MW (Rvir ∼ 300 kpc).

Standard image High-resolution image

With the obvious caveat regarding the generally unknown tangential components of the velocities of the galaxies, Figure 2 can be used to define membership of the MW subgroup. In particular, all galaxies within 300 kpc of the MW are likely bound satellites. This is with the exception of Leo I, which may be unbound unless its tangential components are significantly overestimated (also see Mateo et al. 2008; Sohn et al. 2007; Zaritsky et al. 1989, and references therein). Leo T, NGC 6822, and Phoenix may all be bound in this simple model, but their large distance from the MW makes their membership of the MW subgroup ambiguous. I note that no potential satellite of the MW or M31 (Section 3.2) is clearly unbound as a result of the magnitude of its radial velocity alone; all the potential satellites with large radial velocities relative to their host may be bound if their tangential velocities are sufficiently small. For the purpose of the classification in Column 3 of Table 1, and in the forthcoming discussion, I only include as definite MW satellites those objects that are bound by the dashed curves in Figure 2 and which have DGRvir. It is interesting to note that the radial distribution of satellites is such that they are found at all radii out to ∼280 kpc from the MW, at which point there is a notable gap in the distribution until the next set of objects at >400 kpc. A similar gap is present at a similar radius for satellites around M31 (Section 3.2; but see the recent discovery of Andromeda XXVIII by Slater et al. 2011). This tentatively suggests that this gap could be used to observationally define the limiting radius of the sub-groups (or equivalently, the extent of the two host galaxies) and appears consistent with previous work and other galaxy groups (e.g., Cen A, Sculptor, IC 342: Karachentsev et al. 2002a, 2003a, 2003b; see also Grebel et al. 2003).

Pisces II is the most recently discovered satellite to the MW (Belokurov et al. 2010) and does not appear in Figure 2 since it currently lacks a velocity measurement. At the time of writing, 15 new satellites have been discovered since 2005, all of them using the Sloan Digital Sky Survey (SDSS) and hence confined to its footprint. Irwin (1994) originally argued that the MW sub-group was essentially complete at Galactic latitudes away from the disk for objects of comparable luminosity as (or brighter than) Sextans. The fact that only one of the new satellites (Canes Venatici) is comparable in luminosity to any of the previously known MW satellites backs up this claim. Indeed, Koposov et al. (2008) demonstrate that most of the newly discovered galaxies are at the limits of detection of SDSS, implying that—even within the SDSS footprint—a large number of galaxies still await discovery. This was explored further by Tollerud et al. (2008), who conclude that there may by many hundreds of unseen faint satellites.

It is worth emphasizing the problems caused by the Galactic disk in searching for satellites. Figure 3 shows the distribution of MW satellites with (absolute) latitude. The histogram shows the observed distribution of all 27 satellites, whereas the dashed line indicates a uniform distribution, normalized to match observations at |b| > 30°. There are three satellites at |b| < 30° (one of which is Canis Major, on which there is still considerable debate as to whether it is a dwarf galaxy or an unrelated disk substructure; e.g., Martin et al. 2004a, 2004b; Momany et al. 2004; Martínez-Delgado et al. 2005; Moitinho et al. 2006; Butler et al. 2007; López-Corredoira et al. 2007), a number that is very low considering this accounts for half of the sky. Obscuration by the Galactic disk, however, does not become a serious problem until |b| ≲ 20°. While some of the area probed by SDSS reaches to very low Galactic latitudes, the majority of the Legacy area is located at b > 30°, likely explaining the dearth of satellites at intermediate latitudes.

Figure 3.

Figure 3. Galactic latitude distribution of MW dwarf galaxies. The dashed curve shows an isotropic distribution normalized to match the current number of dwarfs known at |b| > 30° (24) and highlights the dearth of objects at low Galactic latitude.

Standard image High-resolution image

Two photometric surveys currently underway—the Pan-STARRS (Kaiser et al. 2002) PS1 survey in the north and SkyMapper (Keller et al. 2007) in the south—together will map the entire sky to equivalent or deeper magnitude limits to SDSS and consequently should reveal the presence of many more satellites. These should remove the spatial bias introduced by SDSS. In addition, other wide-field (but not all-hemisphere) surveys such as that made possible with the soon-to-be-commissioned Subaru/HyperSuprimeCam will allow for even fainter satellites to be detected (if they exist). In the more distant future, the Large Synoptic Survey Telescope will re-examine the southern hemisphere sky to even greater depth. Judging by recent discoveries, there is every reason to expect that large numbers of very faint satellites will be discovered by all these projects. However, objects behind the Galactic plane will likely remain hidden, and it is unlikely that (m)any bright (MV < −8) Galactic dwarf galaxies at higher Galactic latitudes remain to be discovered.

3.2. Membership of the M31 Sub-group

Figure 4 shows M31-centric velocity versus distance for all galaxies within 600 kpc of Andromeda for which these measurements exist. Again, the M31-centric velocity listed in Table 2 has been multiplied by a factor of $\sqrt{3}$ to account for the unknown tangential velocities of these galaxies. Following Figure 2, dashed curves show the escape velocity from a 1012M point mass, and the vertical dashed line indicates the approximate cosmological virial radius of M31, Rvir ∼ 300 kpc (e.g., Klypin et al. 2002).

Figure 4.

Figure 4. M31-centric velocity vs. distance for all galaxies in the proximity of M31. The y-ordinate has been multiplied by a factor of $\sqrt{3}$ to account for the unknown tangential motions of the galaxies. Dashed curves show the escape velocity from a 1012M point mass. The vertical dashed line indicates the approximate location of the expected cosmological virial radius of M31 (Rvir ∼ 300 kpc).

Standard image High-resolution image

Whereas only Leo I around the MW stands out as a possible unbound satellite from its radial velocity (Mateo et al. 2008), M31 has at least two close-in dwarf galaxies that are likely unbound, Andromeda XII (Chapman et al. 2007) and XIV (Majewski et al. 2007). The Pegasus dwarf irregular galaxy (DIG) and IC 1613 are both extremely distant from Andromeda but could arguably be dynamically associated. In this respect they are similar to NGC 6822, LeoT, and Phoenix around the MW. It is often overlooked that IC 10 and LGS 3 are clear M31 satellites (as close to M31 as the Andromeda VI and XVI dSphs). However, these two galaxies differ from the majority of satellites insofar as they both have non-negligible gas fractions and younger stellar populations. LGS 3 is generally classified as being morphologically akin to LeoT (a "transition-type" dwarf; see Section 5 for discussion), whereas IC 10 is generally classified as a low-mass dwarf irregular. Indeed, as a whole, the M31 satellite system contrasts with the MW satellites in the variety of morphological types present. Around the MW, only the LMC and SMC are not classified as dSphs. However, in addition to the dSph satellites, M31 also has a transition dwarf (LGS 3), a dwarf irregular (IC 10, although it is much lower luminosity than either of the Magellanic Clouds), three dwarf elliptical galaxies (NGC 205, 147, and 185), a compact elliptical (M32), and a low-mass spiral galaxy (M33).

In the past seven years we have seen a rapid growth in the number of known M31 satellites through focused wide-field studies of the environment around Andromeda. All the new dwarfs have been discovered as overdensities of red giant branch (RGB) stars. This, therefore, imposes a basic magnitude limit on the galaxies that can be found, since it requires galaxies to have enough stars that a reasonable number are passing through the RGB phase. An exact study of the selection effects for dwarf galaxy searches around M31 has yet to be published, but the practical limit of these searches appears to be around MV ∼ −6. The most extensive survey of M31's environs, by the Pan-Andromeda Archaeological Survey (PAndAS; McConnachie et al. 2009), provides complete spatial coverage out to a maximum projected radius from M31 of 150 kpc. As shown in McConnachie et al. (2009) and Richardson et al. (2011), the projected radial distribution of satellites shows no sign of declining within this radius. It is highly probable, therefore, that a large number of satellites remain to be discovered at large radius (recently verified by the discoveries of Andromeda XXVIII (Slater et al. 2011) and Andromeda XXIX (Bell et al. 2011) in the SDSS DR8). At small radius (≲ 60 kpc), there is also a deficit of galaxies due to the dominance of M31 stellar populations in this region, making it difficult to identify any low-luminosity satellites that may be superimposed in front of (or behind) the main body of M31.

Within both the MW and the M31 sub-group there are at least three potential dynamical associations of satellites ("sub-sub-groups"). Around the MW, Leo IV and V are in close proximity to each other (positions and velocities) and may well constitute a binary system (Belokurov et al. 2008; de Jong et al. 2010). Within the M31 sub-group, Andromeda XXII is only about 40 kpc in projection from M33, whereas it is over 200 kpc in projection from M31 (Martin et al. 2009). Thus, Andromeda XXII may actually be the only known satellite of M33 (the bright dSph Andromeda II is also considerably closer to M33 than M31, but in a region where the gravitational potential of M31 clearly dominates). A second possible pairing is found to the north of M31 and includes NGC 185 and 147. Van den Bergh (1998) postulated that NGC 185 and 147 were a binary system; they are separated by only 1° in projection and their distances imply a three-dimensional separation of ∼60 kpc. Their velocities differ by only ∼10 km s−1, and such a small separation in phase-space seems unlikely (but not impossible) if these galaxies are not, or never have been, a binary system.

3.3. Membership of the Local Group, and Identification of Nearby Neighbors

Figure 5 shows Local-Group-centric velocity versus distance from the barycenter of the Local Group for all galaxies in the sample for which these data exist. The Local-Group-centric velocity listed in Table 2 has been multiplied by a factor of $\sqrt{3}$ to account for the unknown tangential components. Dashed curves indicate the escape velocity from a point mass of 2 × 1012M. Dotted curves indicate the escape velocity from a point mass of 5 × 1012M, the approximate total dynamical mass of the Local Group as implied from the timing argument (e.g., Lynden-Bell 1981). Many of the points at DLG ≲ 500 kpc represent galaxies that are satellites of either the MW or M31. These systems are dynamically dominated by the gravitational potential of these massive hosts rather than the Local Group as a whole.

Figure 5.

Figure 5. Local-Group-centric velocity vs. distance for all galaxies in the proximity of the Local Group. The y-ordinate has been multiplied by a factor of $\sqrt{3}$ to account for the unknown tangential motions of the galaxies. Dashed curves show the escape velocity from a 2 × 1012M point mass. Dotted curves show the escape velocity from a 5 × 1012M point mass, as implied from the timing argument (Lynden-Bell 1981). The vertical dashed line indicates the zero-velocity surface of the Local Group (RLG = 1060 ± 70 kpc), derived here from the mean distance of the six highlighted galaxies that cluster around zero velocity.

Standard image High-resolution image

There is a clear separation in Figure 5 between outer, apparently bound, Local Group members (the outermost of which is UGC 4879) and the next nearest galaxies (the four members of the NGC 3109 grouping; see van den Bergh 1999). The rise in velocity with distance for all the nearby neighbors is due to the Hubble expansion, and the reader is referred to Karachentsev et al. (2009) for a beautiful analysis of the Hubble flow around the Local Group.

The six outermost Local Group members cluster around vLG = 0, and their mean velocity is consistent with zero (〈v〉 = 4 ± 16 km s−1, where the uncertainty is the random error in the mean). As such, their mean distance can be used as an estimate of the zero-velocity surface of the Local Group and is RLG = 1060 ± 70 kpc (where the uncertainty is the random error in the mean; cf. Courteau & van den Bergh 1999; Karachentsev et al. 2002b, 2009; and others). It is also worth pointing out that the velocity dispersion of the Local Group can be estimated using those Local Group galaxies that are far from either the MW or M31 and that have published velocities. The six Local Group galaxies identified in Figure 5, with the addition of IC 1613 and Cetus, are all more than 500 kpc from either of the two large galaxies and yield σLG = 49 ± 36 km s−1 (where the uncertainty is the random error in the dispersion). Note that there is a systematic error component in all these estimates caused by our choice for the location of the barycenter of the Local Group, although it has a generally weak effect on any values given here. The fact that the measured velocity dispersion for the entire Local Group is so low that it is of order the velocity dispersion of the gas in the SMC is peculiar and well noted (e.g., Sandage 1986, and references therein). It may result from the necessity to measure the dispersion from galaxies that are mostly at or near turnaround.

Not including satellites of the MW and M31, it is highly likely that the census of Local Group galaxies remains severely incomplete. Unlike recently discovered MW satellites, isolated Local Group galaxies are not nearby. Unlike the M31 sub-group, isolated Local Group galaxies do not cluster in a specific area of sky amenable to dedicated surveys. The middle panel of Figure 1 shows that 8 of the 13 (quasi-)isolated Local Group galaxies are all in one quadrant of the sky (the same one that contains M31). However, it is also not clear what the expected spatial distribution of non-satellite Local Group galaxies should be, and it is likely naive to expect an isotropic distribution. Heroic searches for new Local Group galaxies and nearby neighbors have been made by groups led by I. Karachentsev, V. Karachentseva (e.g., Karachentseva & Karachentsev 1998; Karachentseva et al. 1999; Karachentsev et al. 2000 and many others), and A. Whiting (Whiting et al. 1997, 1999) through the visual inspection of literally every plate of the second Palomar Observatory Sky Survey and the ESO/Science Research Council survey, and these remain the most comprehensive all-sky searches to date. Whiting et al. (2007) attempt to quantify their visual search of these plates; they estimate that their search is ∼80% complete in the region away from interference from the MW disk (more than 70% of the sky) to a surface brightness level approaching 26 mag arcsec−2. This translates to (at most) 1 or 2 currently uncataloged Local Group members down to this limit, away from the disk of the Galaxy. Current and future large-scale digital surveys of the sky—including blind H i surveys as well as optical studies—are potentially rich hunting grounds to try to improve the census of nearby galaxies, be they members of the Local Group or the nearby neighbors that help define its environment.

4. PHOTOMETRIC AND STRUCTURAL PARAMETERS

Table 3 presents global photometric and structural parameters for each galaxy in the sample.

Table 3. Luminosity and Structural Parameters

(1) (2) (3) (4) (5) (6) (7) (8) (9) (10) (11)
Galaxy V rh μV θ epsilon MV rh μeff Referencesa Comments
      (') (M/□'') (°)         (pc) (M/□'')    
The Galaxy                                  
Canis Major −0.1 0.8 ... ... 24.0b 0.6 123:c   0.8:   −14.4d 0.8 ... ... ... (1) (2)  
Sagittarius dSph 3.6 0.3 342.00 12.00 25.2 0.3 102 2 0.64 0.02 −13.5 0.3 2587 219 26.0 (3) (4)  
Segue 15.3 0.8 4.40 +1.2−0.6 27.6 1.0 85 8 0.48 0.13 −1.5 0.8 29 +8−5 28.7 (5)  
Ursa Major II 13.3 0.5 16.00 1.00 27.9 0.6 98 4 0.63 0.05 −4.2 0.6 149 21 29.1 (5)  
Bootes II 15.4 0.9 4.20 1.40 28.1 1.6 145 55 0.21 0.21 −2.7 0.9 51 17 29.1 (5)  
Segue II 15.2 0.3 3.40 0.20 27.4 0.4 182 17 0.15 0.10 −2.5 0.3 35 3 28.6 (6)  
Willman 1 15.2 0.7 2.30 0.40 26.1 0.9 77 5 0.47 0.08 −2.7 0.8 25 6 27.2 (5)  
Coma Berenices 14.1 0.5 6.00 0.60 27.3 0.7 115 10 0.38 0.14 −4.1 0.5 77 10 28.4 (5)  
Bootes III 12.6 0.5 ... ... 31.3e 0.3 90:   0.50:   −5.8d 0.5 ... ... ... (7)  
LMC 0.4 0.1                 −18.1 0.1       (8)  
SMC 2.2 0.2                 −16.8 0.2       (8)  
Bootes 12.8 0.2 12.60 1.00 27.5 0.3 14 6 0.39 0.06 −6.3 0.2 242 21 28.7 (5)  
Draco 10.6 0.2 10.00 0.30 25.0 0.2 89 2 0.31 0.02 −8.8 0.3 221 19 26.1 (5)  
Ursa Minor 10.6 0.5 8.20 1.20 26.0 0.5 53 5 0.56 0.05 −8.8 0.5 181 27 25.2 (9)  
Sculptor 8.6 0.5 11.30 1.60 23.5 0.5 99 1 0.32 0.03 −11.1 0.5 283 45 24.3 (9)  
Sextans 10.4 0.5 27.80 1.20 27.1 0.5 56 5 0.35 0.05 −9.3 0.5 695 44 28.0 (9)  
Ursa Major 14.4 0.3 11.30 1.70 27.7 0.5 71 3 0.80 0.04 −5.5 0.3 319 50 28.8 (5)  
Carina 11.0 0.5 8.20 1.20 25.5 0.5 65 5 0.33 0.05 −9.1 0.5 250 39 26.0 (9)  
Hercules 14.0 0.3 8.60 +1.8−1.1 27.2 0.6 102 4 0.68 0.08 −6.6 0.4 330 +75−52 28.3 (5)  
Fornax 7.4 0.3 16.60 1.20 23.3 0.3 41 1 0.30 0.01 −13.4 0.3 710 77 24.0 (9)  
Leo IV 15.1 0.4 4.60 0.80 27.5 0.7 121 9 0.49 0.11 −5.8 0.4 206 37 28.6 (10)  
Canes Venatici II 16.1 0.5 1.60 0.30 26.1 0.7 177 9 0.52 0.11 −4.9 0.5 74 14 27.2 (5)  
Leo V 16.0 0.4 2.60 0.60 27.1 0.8 96 13 0.50 0.15 −5.2 0.4 135 32 28.2 (10)  
Pisces II 16.3 0.5 1.10 0.10 25.7 0.6 77 12 0.40 0.10 −5.0:   58 99 26.8 (11)  
Canes Venatici 13.1 0.2 8.90 0.40 27.1 0.2 70 4 0.39 0.03 −8.6 0.2 564 36 28.2 (5)  
Leo II 12.0 0.3 2.60 0.60 24.2 0.3 12 10 0.13 0.05 −9.8 0.3 176 42 24.8 (9)  
Leo I 10.0 0.3 3.40 0.30 22.6 0.3 79 3 0.21 0.03 −12.0 0.3 251 27 23.3 (9)  
Andromeda                                  
M32 8.1 0.1 0.47 0.05 11.1:f   159 2 0.25 0.02 −16.4 0.2 110 16 17.0 (8) (12)  
Andromeda IX 16.3 1.1 2.50 0.10 28.0 1.2 ... ... ... ... −8.1 1.1 557 29 29.2 (13) (14)  
NGC 205 8.1 0.1 2.46 0.10 15.4:g   28 5 0.43 0.10 −16.5 0.1 590 31 20.3 (8) (12)  
Andromeda XVII 15.8 0.4 1.24 0.08 25.7 0.5 122 7 0.27 0.06 −8.7 0.4 286 23 26.8 (15)  
Andromeda I 12.7 0.1 3.10 0.30 24.7 0.2 22 15 0.22 0.04 −11.7 0.1 672 69 25.8 (16)  
Andromeda XXVIIh 16.7 0.5 1.80 0.30 27.6 0.5 150 10 0.40 0.20 −7.9 0.5 434 76 28.3 (17)  
Andromeda III 14.4 0.3 2.20 0.20 24.8 0.2 136 3 0.52 0.02 −10.0 0.3 479 46 26.2 (16)  
Andromeda XXV 14.8 0.5 3.00 0.20 27.1 0.5 170 10 0.25 0.05 −9.7 0.5 709 61 27.8 (17)  
Andromeda XXVI 17.3 0.5 1.00 0.10 27.4 0.5 145 10 0.25 0.05 −7.1 0.5 222 25 27.9 (17)  
Andromeda XI 17.5 1.2 0.71 0.03 26.5 1.3 ... ... ... ... −6.9 1.3 157 +16−37 27.6 (13) (14)  
Andromeda V 15.3 0.2 1.40 0.20 25.3 0.2 32 10 0.18 0.05 −9.1 0.2 315 46 26.7 (16)  
Andromeda X 16.6 1.0 1.30 0.10 26.3 1.1 46 5 0.44 0.06 −7.6 1.0 265 33 27.4 (15)  
Andromeda XXIII 14.2 0.5 4.60 0.20 28.0 0.5 138 5 0.40 0.05 −10.2 0.5 1029 76 27.8 (17)  
Andromeda XX 18.2 0.8 0.53 +0.14−0.04 26.2 0.8 80 20 0.30 0.15 −6.3 +1.1−0.8 124 +53−17 27.3 (18)  
Andromeda XII 18.3 1.2 1.20 0.20 28.5 1.3 ... ... ... ... −6.4 1.2 304 66 29.6 (13) (14)  
NGC 147 9.5 0.1 3.17:   21.2:   25 3 0.41 0.02 −14.6 0.1 623:   22.3 (8) r25 = 6.6 ± 0.3 arcmins
Andromeda XXI 14.8 0.6 3.50 0.30 27.0 0.4 110 15 0.20 0.07 −9.9 0.6 875 91 28.2 (19)  
Andromeda XIV 15.9 0.5 1.70 0.80 27.2 0.6 ... ... 0.31 0.09 −8.4 0.6 363 180 27.5 (20)  
Andromeda XV 14.6 0.3 1.21 0.05 24.8 0.4 0:   0.00:   −9.4 0.4 222 22 25.9 (21)  
Andromeda XIII 18.1 1.2 0.78 0.08 27.3 1.3 ... ... ... ... −6.7 1.3 207 +23−44 28.4 (13) (14)  
Andromeda II 11.7 0.2 6.20 0.20 24.5 0.2 34 6 0.20 0.08 −12.4 0.2 1176 50 26.3 (16)  
NGC 185 9.2 0.1 2.55:   20.8:   35 3 0.15 0.01 −14.8 0.1 458:   21.9 (8) r25 = 5.8 ± 0.3 arcmins
Andromeda XXIX 16.0 0.3 1.70 0.20 27.6 0.4 51 8 0.35 0.06 −8.3 0.4 361 56 27.6 (22)  
Andromeda XIX 15.6 0.6 6.20 0.10 29.3 0.7 37 8 0.17 0.02 −9.2 0.6 1683 105 30.2 (18)  
Triangulum 5.7 0.1         23   0.41   −18.8 0.1       (8)  
Andromeda XXIV 16.3 0.5 2.10 0.10 27.8 0.5 5 10 0.25 0.05 −7.6 0.5 367 27 28.5 (17)  
Andromeda VII 11.8 0.3 3.50 0.10 23.2 0.2 94 8 0.13 0.04 −12.6 0.3 776 42 25.3 (16)  
Andromeda XXII 18.0 0.8 0.94 0.10 26.7 0.6 114 15 0.56 0.11 −6.5 9.9 217 99 27.9 (19)  
IC 10 9.5 0.2 2.65:   24.6 0.2 ... ... 0.19 0.02 −15.0 0.2 612:   22.3 (8) (23) r25 = 3.2 ± 0.1 arcmins
LGS 3 14.3 0.1 2.10 0.20 24.8 0.1 0:   0.20:   −10.1 0.1 470 47 26.6 (24)  
Andromeda VI 13.2 0.2 2.30 0.20 24.1 0.2 163 3 0.41 0.03 −11.3 0.2 524 49 25.3 (16)  
Andromeda XVI 14.4 0.3 0.89 0.05 23.9 0.4 0:   0.00:   −9.2 0.4 136 15 25.0 (21)  
Andromeda XXVIII 15.6 +0.4−0.9 1.11 0.21 26.3 +0.4−0.9 39 16 0.34 0.13 −8.5 +0.4−1.0 213 +64−45 26.3 (25)  
IC 1613 9.2 0.1 6.81:   22.8:   50 2 0.11 0.03 −15.2 0.2 1496:   24.1 (8) r25 = 8.1 ± 0.2 arcmins
Phoenix 13.2 0.4 3.76:   25.8:   5i 1 0.40j 0.10 −9.9 0.4 454:   26.4 (26) (27) rt = 15.8+4.3−2.8 arcmins; assumed c = 0.7
NGC 6822 8.1 0.2 2.65:   19.7:   330 10 0.24 0.05 −15.2 0.2 354:   20.8 (8) (28) r25 = 7.7 ± 0.2 arcmins
Cetus 13.2 0.2 3.20 0.10 25.0 0.2 63 3 0.33 0.06 −11.2 0.2 703 31 26.2 (16)  
Pegasus dIrr 12.6 0.2 2.10:   22.8:   120 2 0.46 0.02 −12.2 0.2 562:   24.4 (8)  
Leo T 15.1k 0.5 0.99 0.06 24.8 0.6 0:   0.00:   −8.0 0.5 120 9 26.0 (29)  
WLM 10.6 0.1 7.78:   23.7:   4 2 0.65 0.01 −14.2 0.1 2111:   24.8 (8)  
Leo A 12.4 0.2 2.15:   22.8:   114 5 0.40 0.03 −12.1 0.2 499:   24.4 (8) (30) r25 = 2.6 ± 0.1 arcmins
Andromeda XVIII 16.0 9.9 0.92 0.06 25.6:   ... ... ... ... −9.7:   363 32 26.7 (18) Lower limits on V and μV
Aquarius 14.5 0.1 1.47 0.04 23.6 0.2 99 1 0.50 0.10 −10.6 0.1 458 21 25.5 (31)  
Tucana 15.2 0.2 1.10 0.20 25.0 0.1 97 2 0.48 0.03 −9.5 0.2 284 54 25.6 (32)  
Sagittarius dIrr 13.6 0.2 0.91 0.05 23.9 0.2 90:   0.50:   −11.5 0.3 282 28 23.5 (33) rh from integral under central King profile
UGC 4879 13.2 0.2 0.41 0.04 21.2 0.2 84 11 0.44 0.04 −12.5 0.2 162 16 21.5 (34)  
NGC 3109 10.7 0.1 4.30 0.10 22.6 0.1 92 1 0.82 0.01 −14.9 0.1 1626 71 22.9 (35)  
Sextans B 11.3 0.2 1.06 0.10 21.9 0.3 110 2 0.31 0.03 −14.5 0.2 440 42 21.9 (8) (36)  
Antlia 15.2 0.2 1.20 0.12 23.9 0.2 135 5 0.40 0.04 −10.4 0.2 471 52 25.9 (37) (36)  
Sextans A 11.5 0.1 2.47:   22.8:   0 1 0.17 0.02 −14.3 0.1 1029:   24.1 (8) (38) r25 = 2.9 ± 0.1 arcmins
HIZSS 3(A) ... ... ... ... ... ... ... ... ... ... ... ... ... ... ...    
HIZSS 3B ... ... ... ... ... ... ... ... ... ... ... ... ... ... ...    
KKR 25 15.9 0.2 0.40:   24.0 0.2 ... ... 0.41 0.02 −10.5 0.2 222:   24.2 (39) r26.5 = 0.55 arcmins
ESO 410- G 005 14.9 0.3 0.50 0.05 22.2 0.2 57:   0.37:   −11.5 0.3 280:   23.8 (40) (36)  
NGC 55 7.9 0.1 5.16:   19.3:   108 2 0.83 0.01 −18.5 0.2 2900:   20.4 (8) r25 = 16.2 ± 0.4 arcmins
ESO 294- G 010 15.3 0.3 0.42 0.04 22.3 0.2 8:   0.51:   −11.2 0.3 248 24 23.5 (40) (36)  
NGC 300 8.1 0.1 5.00:   21.0:   111 2 0.29 0.01 −18.5 0.1 3025:   22.1 (8) r25 = 10.1 ± 0.3 arcmins
IC 5152 10.9:l   0.97:   20.1:   100 2 0.38 0.02 −15.6:   550:   21.2 (8) r25 = 2.6 ± 0.1 arcmins
KKH 98 15.2 0.3 0.64 0.06 22.8 0.2 −5 1 0.41 0.01 −11.8 0.3 470 48 24.5 (36) (38)  
UKS 2323-326 13.5 0.2 0.90 0.10 22.9 0.2 −60 4 0.10 0.01 −13.2 0.2 578 69 24.0 (41) (38)  
KKR 3 17.2 0.3 0.36 0.04 23.8 0.2 0 1 0.05 0.01 −9.5 0.3 229 28 25.8 (36) (38)  
GR 8 14.5 0.2 0.32 0.04 22.6 0.2 61 2 0.20 0.05 −12.2 0.2 203 28 22.7 (42) (43)  
UGC 9128 14.4 0.3 0.64 0.07 22.6 0.2 46 2 0.40 0.05 −12.4 0.3 427 47 23.8 (43) (36)  
UGC 8508 13.7 0.1 0.42 0.04 21.5 0.2 −60 2 0.45 0.05 −13.4 0.1 315 30 22.1 (43) (36)  
IC 3104 12.8:l   2.01:   23.3:   45 2 0.52 0.02 −14.0:   1327:   24.4 (8) r25 = 1.9 ± 0.1 arcmins
DDO 125 12.7 0.3 1.04 0.10 22.1 0.2 −68 4 0.41 0.01 −14.4 0.3 781 77 23.1 (36) (38)  
UGCA 86 14.2:l   0.94:   22.8:   25 1 0.32 0.03 −13.2:   811:   24.5 (44) (38) r25 = 2.3 ± 0.2 arcmins
DDO 99 13.9 0.1 0.90 0.09 22.9 0.2 70 4 0.29 0.01 −13.2 0.2 679 81 24.2 (42) (38)  
IC 4662 11.1 0.3 0.48 0.05 18.7 0.2 −69 4 0.27 0.01 −15.8 0.3 341 44 20.1 (36) (38)  
DDO 190 12.8 0.1 0.64 0.06 21.4 0.2 82 5 0.10 0.02 −14.4 0.1 520 49 22.6 (45) (36)  
KKH 86 17.1 0.3 0.28 0.03 23.2 0.2 −3 1 0.39 0.01 −10.0 0.3 210 27 24.7 (36) (38)  
NGC 4163 13.2 0.3 0.45 0.05 21.1 0.2 11 2 0.30 0.05 −14.1 0.3 374 42 22.0 (43) (36)  
DDO 113 16.4 0.3 0.70 0.07 24.0 0.2 0:   0.00 0.09 −11.0 0.3 601 62 26.5 (8) (36)  

aReferences: (1) Bellazzini et al. 2006; (2) Butler et al. 2007; (3) Mateo et al. 1998; (4) Majewski et al. 2003; (5) Martin et al. 2008b; (6) Belokurov et al. 2009; (7) Correnti et al. 2009; (8) de Vaucouleurs et al. 1991; (9) Irwin & Hatzidimitriou 1995; (10) de Jong et al. 2010; (11) Belokurov et al. 2010; (12) Choi et al. 2002; (13) Collins et al. 2010; (14) Collins 2011; (15) Brasseur et al. 2011b; (16) McConnachie & Irwin 2006; (17) Richardson et al. 2011; (18) McConnachie et al. 2008; (19) Martin et al. 2009; (20) Majewski et al. 2007; (21) Ibata et al. 2007; (22) Bell et al. 2011; (23) Sanna et al. 2010; (24) Lee 1995; (25) Slater et al. 2011; (26) van de Rydt et al. 1991; (27) Martínez-Delgado et al. 1999; (28) Dale et al. 2007; (29) de Jong et al. 2008; (30) Vansevičius et al. 2004; (31) McConnachie et al. 2006; (32) Saviane et al. 1996; (33) Lee & Kim 2000; (34) Bellazzini et al. 2011; (35) Hidalgo et al. 2008; (36) Sharina et al. 2008; (37) Aparicio et al. 1997; (38) Fingerhut et al. 2010; (39) Karachentsev et al. 2001b; (40) Loveday 1996; (41) Lee & Byun 1999; (42) Makarova 1999; (43) Vaduvescu et al. 2005; (44) Karachentsev et al. 1997; (45) Aparicio & Tikhonov 2000. bMeasured within central 2 × 2 deg. cBroadly aligned with the Galactic Plane. dDistance-independent absolute magnitude estimate. eMeasured within projected radius of 0.8 deg. fB magnitude, corresponding to "standard fit" in Table 1 of Choi et al. (2002). gB magnitude; see Table 1 of Geha et al. (2006b). hMeasurement of structural parameters complicated by presence of surrounding stream. iCorresponds to "outer component"; θinner = 95°. jCorresponds to "outer component"; epsiloninner = 0.4. kConverted from g, r. lB magnitude.

Download table as:  ASCIITypeset images: 1 2 3

Column 1. Galaxy name.

Column 2. Apparent magnitude in V (Vega magnitudes). In general, these have been corrected for foreground extinction (but not internal extinction), although in some cases the original papers are unclear as to whether extinction corrections were applied to the apparent magnitudes. For many of the fainter galaxies that have been resolved into stars, the magnitude may not be based on integrated light (often undetectable) and is instead estimated by measuring the luminosity of the brighter stars and by making assumptions regarding the luminosity function of stars below the detection limit.

Column 3. Radius (arcmin) containing half the light of the galaxy, measured on the semimajor axis. A vast array of different scale radii are used in the literature, and in cases where the original papers use other scale radii (e.g., core, exponential, half-brightness, R25, etc.), I have derived half-light radii by integrating under the appropriate profile, normalized to the apparent magnitude (and with the appropriate ellipticity if measured, otherwise assumed to be circular). In cases where the original papers did not fit a profile, I have assumed an exponential profile. The latter estimates are particularly uncertain, and in these cases I have also given the original scale radius in Column 11.

Column 4. Central surface brightness (mag arcsec−2). Where possible, I cite directly measured values, although in many cases (including in the original papers) this parameter is derived by normalizing the measured profile to the apparent magnitude.

Column 5. Position angle of major axis, measured in degrees east from north. For systems that show a change in this quantity as a function of radius, a mean value has been estimated.

Column 6. Ellipticity epsilon = 1 − b/a, where b is the semiminor axis and a is the semimajor axis. For systems that show a change in this quantity as a function of radius, a mean value has been estimated.

Column 7. Absolute visual magnitude, derived from Column 2 by subtraction of the distance modulus given in Table 2.

Column 8. Half-light radius in parsecs, derived from Column 3 assuming the distance modulus given in Table 2.

Column 9. The mean surface brightness within the isophote defined by the half-light radius, derived using values from Columns 2, 3, and 6. In cases where no ellipticity is measured, I have assumed circular isophotes.

Column 10. References.

Column 11. Comments.

For each galaxy I have tried to minimize the number of different papers that I cite to provide complete information, with the result that Table 3 references 45 papers from the last 21 years. Concern must be given to the systematic uncertainties that exist between measurements: they are necessarily complex, they are essentially unquantifiable, and they probably present the primary limitation for examination of the global properties of this population.

Despite the usefulness of parameters such as total luminosity and half-light radius, it should be strongly stressed that these numbers are wholly inadequate in reflecting the known complexity of galaxy structures. It is standard practice to decompose bright galaxies into components such as nucleus and envelope, or a bulge, a disk (several disks?), and a halo. More recently, however, a vast number of independent studies of galaxies included in Table 3 have shown that a full description of their structure requires decomposition of their surface brightness profile into multiple components (e.g., Irwin & Hatzidimitriou 1995; Martínez-Delgado et al. 1999; Lee et al. 1999; Lee & Byun 1999; Lee & Kim 2000; Vansevičius et al. 2004; McConnachie et al. 2007; Hidalgo et al. 2008, among many others). The addition of spectroscopic data (see Section 5) adds new complexity to the discussion as well, since dynamically distinct (e.g., Kleyna et al. 2003, 2004) and chemo-dynamically distinct (e.g., Tolstoy et al. 2004; Battaglia et al. 2006, 2008, 2011) populations of stars have been shown to reside within individual dwarf galaxies that do not necessarily always reveal themselves as features in the global surface brightness profiles. Thus, not only are integrated photometric properties such as rh and MV inadequate to describe the complex structures exhibited by many galaxies, but so too are single velocity dispersion or rotation measurements (Section 5; true for H i as well as stars; e.g., Lo et al. 1993; Young & Lo 1996, 1997a, 1997b; Young et al. 2003).

Figures 6 and 7 show basic scaling relations using the galaxy parameters given in Table 3, first studied in detail in Kormendy (1985). In Figure 6, I plot absolute magnitude against half-light radius. Here, I have converted the half-light radius given in Table 3 to the geometric mean half-light radius of the major and minor axes ($r = \sqrt{r_ar_b}$, where ra and rb are radii measured on the major and minor axes, respectively) to account for the presence of highly elliptical systems. The top panel includes as small dots the location of the Galactic globular clusters, from the data compiled by Harris (1996), and the dashed line shows the direction defined by points of constant surface brightness (averaged within the half-light radius). I have labeled obvious outlying points, and I have also labeled some of the least luminous (candidate) dwarf galaxies. The lower panel shows an expansion of the region −17 ⩽ MV ⩽ −7.

Figure 6.

Figure 6. Top panel: the absolute visual magnitude vs. half-light radius (geometric mean of the major and minor axes) for the galaxy sample (where the symbols and color-coding identify Galactic, M31, Local Group, and nearby galaxies, as explained in the key). Also indicated as small black dots are the location of the Galactic globular clusters, using the data compiled by Harris (1996). Obvious outliers to the overall trend displayed by the galaxies are highlighted by name, as are some of the least luminous (candidate) dwarf galaxies. The dashed line indicates the direction defined by points of constant surface brightness (averaged within the half-light radius). Bottom panel: same parameters as top panel showing an enlargement around −17 ⩽ MV ⩽ −7.

Standard image High-resolution image
Figure 7.

Figure 7. Surface brightness vs. absolute visual magnitude for the galaxy sample, where symbols have the same meaning as in Figure 6. The top panel uses central surface brightness, whereas the bottom panel uses the surface brightness averaged over the half-light radius. Obvious outliers to the overall trends are highlighted by name. Note that Bootes III is not present in the lower panel due to the lack of a measured value for rh.

Standard image High-resolution image

The observed relationship between absolute magnitude and half-light radius shown in the top panel of Figure 6 spans a factor of one million in luminosity. The galaxies seem well separated from the (Galactic) globular clusters in this two-dimensional projection, although the limited number of low-luminosity galaxies (and the fact that all of them are MW satellites) complicates interpretation of this region of the diagram. When MV and rh are considered separately, the one-dimensional distributions do overlap, such that the smallest, least luminous (candidate) dwarf galaxies are as faint as any known globular clusters and smaller than the most extended clusters. The distinction between these stellar systems in this parameter space becomes less clear when other types of compact stellar systems, such as ultracompact dwarfs and the nuclei of early-type galaxies, are included (e.g., see Figures 5 and 8 of Brodie et al. 2011).

Three galaxies are clear outliers in the top panel of Figure 6: Andromeda XIX is extremely extended for its luminosity (see also Figure 7), whereas the compact elliptical M32 is a known outlier in this projection due to its extreme concentration (see also Kormendy & Bender 2012, and references therein). Interestingly, the recently identified Local Group galaxy UGC 4879 (Kopylov et al. 2008) tends to the same side of the relation as M32, although not as extreme. Bellazzini et al. (2011) note that UGC 4879 has an unusual structure, insofar as there appear to be two flattened wings emanating from the central spheroid of stars, perhaps indicating the presence of a disk. More work is needed to determine the significance of the structure of UGC 4879 and its position in this diagram.

The lower panel of Figure 6 enlarges the region −17 ⩽ MV ⩽ −7 to demonstrate the very large scatter that exists in this regime (see below for discussion of the fainter systems). McConnachie & Irwin (2006) originally pointed out that, at a given luminosity, the M31 dwarf satellites were generally larger than the MW population. With a large number of new discoveries since this time, Brasseur et al. (2011a) demonstrate that the mean relations of the M31 and MW populations are statistically consistent with each other3. The limited number of points contributing to the relations for these different sub-groups makes interpretation difficult; for example, 4 out of 10 MW satellites have rh > 250 pc, whereas 20 out of 26 M31 satellites are this large. It may be that local environment contributes to determining the sizes of dwarf galaxies, but perhaps the most important conclusion that can be drawn from the lower panel of Figure 6 is the lack of a strong relation between MVrh in this regime; at any given luminosity, galaxy sizes vary by nearly an order of magnitude, and there is only a weak dependence on luminosity.

At even fainter magnitudes, the line of constant surface brightness (averaged within the half-light radius) included in the top panel of Figure 6 runs approximately parallel to the relation defined by the galaxy locus for MV ≳ −8 (and this also has a different slope than the weak relation at brighter magnitudes). This indicates that the smallest galaxies so far discovered all have similar surface brightnesses and further suggests that selection effects may be driving the form of the MVrh relation at faint magnitudes. To explore this further, I plot surface brightness against absolute magnitude in Figure 7. The top panel shows the central surface brightness, whereas the bottom panel uses the average surface brightness within the half-light radius. The former is in principle more susceptible to localized features (e.g., nuclei, individual bright stars, etc.) than the latter. I have labeled points that are obvious outliers in both panels (note that Bootes III is not present in the lower panel of Figure 7 due to the lack of a measured value for rh). These galaxies are all known or suspected to be undergoing tidal disruption.

There is a clear correlation between absolute magnitude and surface brightness in both panels of Figure 7 for galaxies more luminous than MV ∼ −9. Of considerable interest, however, is the change in the observed relationship between surface brightness and absolute magnitude within the dwarf regime. In particular, dwarf galaxies fainter than MV ≃ −8.5 do not continue to decrease in surface brightness with decreasing luminosity. Instead, the surface brightness levels off at faint magnitudes, as implied in Figure 6. The central surface brightness distribution for the 26 galaxies fainter than MV = −8.5 in Figure 7 can be characterized by a median surface brightness of 27.35 mag arcsec−2 and an inter-quartile range of 1.2 mag. Thus, the central surface brightness for galaxies fainter than MV = −8.5 appears remarkably constant given that the luminosity changes by a factor of ∼600 within the sample.

Surface brightness selection effects must be seriously considered in examining the correlations in Figure 7 (see, for example, earlier discussion by Impey & Bothun 1997, and references therein). Here, the work of Koposov et al. (2008) provides the most relevant reference. The selection functions are complex and depend on distance, magnitude, and surface brightness. However, Table 3 of Koposov et al. (2008) indicates that galaxies as faint as MV ∼ −4.4, at distances of ∼180 kpc, could be recovered even if their central surface brightness is as faint as 29.9 mag arcsec−2, more than 2 mag fainter than the median observed surface brightness at faint luminosities. There is also tentative evidence that the M31 satellites appear to behave in a similar way to the MW satellites (although they do not extend to the same low-luminosity limits). Searches for M31 satellites are subject to different (and as yet unquantified) selection effects compared to the MW satellites. Taken together, it appears possible that the "break" in the scaling relations at around MV ∼ −9 may be real. If so, it may imply that—as a result of either formation (e.g., inflow of gas into dark matter haloes) or subsequent feedback/evolution (e.g., star formation)—there is a low-density limit to the central stellar densities of dwarf galaxies. If the effect is instead due to selection, then we are only observing those low-luminosity galaxies with the highest surface brightness, indicating that there may be very significant scatter in this property at faint magnitudes. Regardless of this speculation, the behavior of surface brightness with luminosity shown in Figure 7 certainly requires further consideration and explanation.

5. MASSES AND KINEMATICS

Table 4 lists various information regarding the masses and kinematics of the galaxy sample.

Table 4. Masses and Kinematics

(1) (2) (3) (4) (5) (6) (7) (8) (9) (10)
Galaxy M σ* v*r MH i σH i vH ir Mdyn(⩽ rh) Referencesa Comments
  (106M) (km s−1) (km s−1) (106M) (km s−1) (km s−1) (106M)    
The Galaxy                          
Canis Major 49 20.0 3.0 ... ... ... ... ... ... ... ... (1) H i measurements complex due to location!
Sagittarius dSph 21 11.4 0.7 N/A N/A N/Ab N/A N/A N/A N/A 190 (2) (3) (4) (5)  
Segue (I) 0.00034 3.9c 0.8 N/A N/A N/A N/A N/A N/A N/A 0.26 (6) (5) Stellar velocity gradient constrained to <5 km s−1 at 90% confidence
Ursa Major II 0.0041 6.7 1.4 N/A N/A N/A N/A N/A N/A N/A 3.9 (7) (5) Velocity difference of 8.4 ± 1.4 km s−1 between stars in the east and west
Bootes II 0.0010 10.5 7.4 ... ... N/A N/A N/A N/A N/A 3.3 (8) (5)  
Segue II 0.00086 3.4 +2.5−1.2 ... ... ... ... ... ... ... 0.23 (9)  
Willman 1 0.0010 4.3 +2.3−1.3 ... ... N/A N/A N/A N/A N/A 0.27 (10) (5)  
Coma Berenices 0.0037 4.6 0.8 N/A N/A N/A N/A N/A N/A N/A 0.94 (7) (5) Stellar velocity gradient of 5.5 ± 1.2 km s−1 deg−1
Bootes III 0.017 14.0 3.2 ... ... ... ... ... ... ... ... (11)  
LMC 1500 20.2 0.5 49.8 15.9 460 15.8 0.2 63.0 3.0   (12) (13) (14) Stellar velocities based on carbon stars; H i boundaries between LMC, SMC, bridge, and interface regions not clearly defined
SMC 460 27.6 0.5 N/A N/A 460 22.0d 2.0 60.0 5.0   (15) (16) (14) Stellar velocities based on red giant branch stars; stellar velocity gradient of 8.3 km s−1 deg−1; H i boundaries between LMC, SMC, bridge, and interface regions not clearly defined
Bootes (I) 0.029 2.4 +0.9−0.5 ... ... N/A N/A N/A N/A N/A 0.81 (17) (5) Stellar velocity dispersion corresponds to dominant component, and there is evidence for a hotter component with σ* ∼ 9 km s−1
Draco 0.29 9.1 1.2 N/A N/A N/A N/A N/A N/A N/A 11 (18) (19) (5)  
Ursa Minor 0.29 9.5 1.2 N/A N/A N/A N/A N/A N/A N/A 9.5 (20) (19) (5)  
Sculptor 2.3 9.2 1.4 N/A N/A 0.22e N/A N/A N/A N/A 14 (21) (22) (23) (5) Stellar velocity gradient of 5.5 ± 0.5 km s−1 deg−1
Sextans (I) 0.44 7.9 1.3 N/A N/A N/A N/A N/A N/A N/A 25 (21) (22) (5)  
Ursa Major (I) 0.014 7.6 1.0 ... ... N/A N/A N/A N/A N/A 11 (7) (5)  
Carina 0.38 6.6 1.2 N/A N/A N/A N/A N/A N/A N/A 6.3 (21) (22) (5) Stellar velocity gradient of 2.5 ± 0.8 km s−1 deg−1
Hercules 0.037 3.7 0.9 ... ... N/A N/A N/A N/A N/A 2.6 (24) (5)  
Fornax 20 11.7 0.9 N/A N/A 0.17f N/A N/A N/A N/A 56 (21) (22) (25) (5) Stellar velocity gradient of 6.3 ± 0.2 km s−1 deg−1
Leo IV 0.019 3.3 1.7 ... ... N/A N/A N/A N/A N/A 1.3 (7) (5)  
Canes Venatici II 0.0079 4.6 1.0 ... ... N/A N/A N/A N/A N/A 0.91 (7) (5)  
Leo V 0.011 3.7 +2.3−1.4 ... ... N/A N/A N/A N/A N/A 1.1 (26) (27) (5)  
Pisces II 0.0086 ... ... ... ... ... ... ... ... ... ...    
Canes Venatici (I) 0.23 7.6 0.4 ... ... N/A N/A N/A N/A N/A 19 (7) (5)  
Leo II 0.74 6.6 0.7 ... ... N/A N/A N/A N/A N/A 4.6 (18) (5)  
Leo I 5.5 9.2 1.4 N/A N/A N/A N/A N/A N/A N/A 12 (28) (5)  
Andromeda                          
M32 320 92.0 5.0 55.0 3.0 N/A N/A N/A N/A N/A 540 (29) (5) Stellar velocity dispersion rises rapidly toward center and presents strong evidence for a supermassive black hole (e.g., see also van der Marel et al. 1997)
Andromeda IX 0.15 4.5 3.6 ... ... N/A N/A N/A N/A N/A 6.5 (30) (5)  
NGC 205 330 35.0 5.0 11.0 5.0 0.40 16.1g   N/A N/A 420 (31) (32) H i velocity gradient of ∼42 km s−1 measured over ∼3 arcmin
Andromeda XVII 0.26 ... ... ... ... N/A N/A N/A N/A N/A ... (5)  
Andromeda I 3.9 10.6 1.1 ... ... N/A N/A N/A N/A N/A 44 (33) (5)  
Andromeda XXVII 0.12 ... ... ... ... ... ... ... ... ... ...    
Andromeda III 0.83 4.7 1.8 ... ... N/A N/A N/A N/A N/A 6.1 (33) (34)h; (5)  
Andromeda XXV 0.68 ... ... ... ... ... ... ... ... ... ...    
Andromeda XXVI 0.060 ... ... ... ... ... ... ... ... ... ...    
Andromeda XI 0.049 ⩽4.6   ... ... N/A N/A N/A N/A N/A 1.9 (30) (5)  
Andromeda V 0.39 11.5 +5.3−4.4 ... ... N/A N/A N/A N/A N/A ... (35) (34) (5) Not associated with H i detection in Blitz & Robishaw (2000)
Andromeda X 0.096 3.9 1.2 ... ... N/A N/A N/A N/A N/A 2.3 (33) (5)  
Andromeda XXIII 1.1 ... ... ... ... ... ... ... ... ... ...    
Andromeda XX 0.029 ... ... ... ... ... ... ... ... ... ...    
Andromeda XII 0.031 2.6 +5.1−2.6 ... ... N/A N/A N/A N/A N/A 1.2 (30) (5)  
NGC 147 62 16.0 1.0 17.0 2.0 N/A N/A N/A N/A N/A 93 (36) (5)  
Andromeda XXI 0.76 ... ... ... ... ... ... ... ... ... ...    
Andromeda XIV 0.20 5.4 1.3 ... ... N/A N/A N/A N/A N/A 6.1 (33) (5)  
Andromeda XV 0.49 11.0 +7.0−5.0 ... ... N/A N/A N/A N/A N/A 16 (37) (5)  
Andromeda XIII 0.041 9.7 +8.9−4.5 ... ... N/A N/A N/A N/A N/A 11 (30) (5) Stellar velocity dispersion based on three stars
Andromeda II 7.6 7.3 0.8 ... ... N/A N/A N/A N/A N/A 36 (33) (5)  
NGC 185 68 24.0 1.0 15.0 5.0 0.11 15.3 0.8 N/Ai N/A 150 (36) (32)  
Andromeda XXIX 0.18 ... ... ... ... ... ... ... ... ... ...    
Andromeda XIX 0.43 ... ... ... ... ... ... ... ... ... ...    
Triangulum 2900                        
Andromeda XXIV 0.093 ... ... ... ... ... ... ... ... ... ...    
Andromeda VII 9.5 9.7 1.6 ... ... N/A N/A N/A N/A N/A 42 (33) (5)  
Andromeda XXII 0.034 ... ... ... ... ... ... ... ... ... ...    
IC 10 86 ... ... ... ... 50 8.0 1.0 34.0 5.0 ... (38) (39) Gaseous disk embedded in complex, extended H i distribution, (with a dispersion of around 30–40 km s−1)
LGS 3 0.96 7.9 +5.3−2.9 ... ... 0.38 9.1 0.3 ⩽5.0   17 (40) (41) Stellar velocity dispersion based on four stars
Andromeda VI 2.8 9.4 +3.2−2.4 ... ... N/A N/A N/A N/A N/A ... (35) (5)  
Andromeda XVI 0.41 ⩽10.0   ... ... N/A N/A N/A N/A N/A 7.9 (37) (5)  
Andromeda XXVIII 0.21 ... ... ... ... ... ... ... ... ... ...    
IC 1613 100 ... ... ... ... 65 25.0 3.0 5.9 1.0 ... (42) (43)  
Phoenix 0.77 ... ... ... ... 0.12 10.0 4.0 N/A N/A ... (44) H i cloud offset from optical center of Phoenix, proposed as supernova-driven outflow; H i velocity gradient of ∼10 km s−1 observed over ∼8 arcmin
NGC 6822 100 ... ... ... ... 130 5.8 1.9 47.0 2.0 ... (45) (46) Demers et al. (2006) obtain velocities for >100 AGB stars.
Cetus 2.6 17.0j 2.0 7.7 1.2 N/A N/A N/A N/A N/A 120 (47) (5) Tentative stellar velocity gradient across minor axis of 7.7 ± 1.2 km s−1; unassociated with nearby H i cloud
Pegasus dIrr 6.61 ... ... ... ... 5.9 <9.0   21.0:   ... (48) (49)  
Leo T 0.14 7.5 1.6 ... ... 0.28 6.9 1.0 N/A N/A 3.9 (7) (50)  
WLM 43 17.5 2.0 20.8 2.0 61 4.5 1.0 30.0 3.0 380 (51) (52)  
Leo A 6.0 9.3 1.3 N/A N/A 11 9.3 1.4 N/A N/A 25 (53) (54) Stellar kinematics derived using 10 young B supergiants + H ii regions; H i gradient of ∼1.3 km s−1 arcmin−1
Andromeda XVIII 0.63 ... ... ... ... ... ... ... ... ... ...    
Aquarius 1.6 ... ... ... ... 4.1 5.8 0.1 N/A N/A ... (48) H i velocity gradient of 6 km s−1 arcmin−1.
Tucana 0.56 15.8 +4.1−3.1 N/A N/A N/A N/A N/A N/A N/A 41 (55) (5) Stellar velocity gradient of ∼16 km s−1 over ∼2 × rh; unlikely associated with nearby H i cloud
Sagittarius dIrr 3.5 ... ... ... ... 8.8 10.0 1.0 N/A N/A ... (41)  
UGC 4879 8.3 ... ... ... ... 0.95 11.0 1.1 N/A N/A ... (56)  
NGC 3109 76 ... ... ... ... 450 10.0:   72.4 0.5 ... (57) (58) (59)  
Sextans B 52 ... ... ... ... 51 <18.0   71:     (60) (49)  
Antlia 1.3 ... ... ... ... 0.73 6.4 0.7 N/A N/A ... (58)  
Sextans A 44 ... ... ... ... 77 8.0 1.0 20.0 2.0 ... (61)  
HIZSS 3(A) ... ... ... ... ... 14 ... ... 42.0 4.0 ... (62)  
HIZSS 3B ... ... ... ... ... 2.6 ... ... N/A N/A ... (62) H i velocity gradient of ∼2.5 km s−1 arcmin−1
KKR 25 1.4 ... ... ... ... ... ... ... ... ... ...   See Huchtmeier et al. (2000) for attempted H i observations
ESO 410- G 005 3.5 ... ... ... ... 0.73 14.2 0.6 ... ... ... (63)  
NGC 55 2200 ... ... ... ... 1300 ... ... 86.0 6.0 ... (64) Castro et al. (2008) obtain stellar velocities for ∼200 massive blue stars and find that their kinematics closely trace the H i rotation curve
ESO 294- G 010 2.7 ... ... ... ... 0.34 9.7 0.9 ... ... ... (63)  
NGC 300 2100 ... ... ... ... 1800 13.0 5.0 98.8 3.1 ... (65) (66) H i velocity dispersion shows significant spatial variation
IC 5152 270 ... ... ... ... 87 ... ... ... ... ... (45) H i line widths available in Koribalski et al. (2004)
KKH 98 4.5 ... ... ... ... 6.6 ... ... ... ... ... (67) H i line widths available in Begum et al. (2008b) and Huchtmeier et al. (2003)
UKS 2323-326 17 ... ... ... ... 17: ... ... ... ... ... (68) See also Huchtmeier & Richter (1986); H i line widths available in Longmore et al. (1982) and Huchtmeier & Richter (1986)
KKR 3 0.54 ... ... ... ... 2.5 7.5 0.5 N/A N/A ... (69) H i velocity gradient of ∼3.3 km s−1 arcmin−1
GR 8 6.4 ... ... ... ... 11 7.0 0.2 N/A N/A ... (48) Lumpy H i velocity distribution, with some evidence of an overall velocity gradient
UGC 9128 7.8 ... ... ... ... 18 <15.0   56.0:   ... (49) See also Begum et al. (2008b)
UGC 8508 19 ... ... ... ... 29 ... ... 31.2 2.3 ... (67) (70)  
IC 3104 62 ... ... ... ... 13 ... ... ... ... ... (45) H i line widths available in Koribalski et al. (2004)
DDO 125 47 ... ... ... ... 35 7.3 1.5 11.2 2.7 ... (71) (72) H i rotational velocity measured at r = 2 arcmin; see also Huchtmeier et al. (2003), Begum et al. (2008b), and Swaters et al. (2009)
UGCA 86 16 ... ... ... ... 860 8.8 1.0 122.0 5.0 ... (73) H i structure consists of a rotating disk and a kinematically distinct, elongated "spur"
DDO 99 16 ... ... ... ... 52 ... ... ... ... ... (67) H i line widths available in Begum et al. (2008b)
IC 4662 190 ... ... ... ... 180 ... ... ... ... ... (45) H i line widths available in Koribalski et al. (2004)
DDO 190 51 ... ... ... ... 43 10.0 2.4 24.7 2.5 ... (71) (72) H i rotational velocity measured at r = 1.5 arcmin
KKH 86 0.82 ... ... ... ... 0.88 ... ... ... ... ... (74) H i line widths available in Karachentsev et al. (2001a)
NGC 4163 37 ... ... ... ... 7.9 <7.4   8.7:   ... (75)  
DDO 113 2.1 ... ... ... ... 48k ... ... ... ... ... (76) H i line widths available in Huchtmeier & Richter (1986)

aReferences: (1) Martin et al. 2005; (2) Ibata et al. 1994; (3) Ibata et al. 1997; (4) Peñarrubia et al. 2011; (5) Grcevich & Putman 2009; (6) Simon et al. 2011; (7) Simon & Geha 2007; (8) Koch et al. 2009; (9) Belokurov et al. 2009; (10) Martin et al. 2007; (11) Carlin et al. 2009; (12) Kim et al. 1998; (13) van der Marel et al. 2002; (14) Brüns et al. 2005; (15) Harris & Zaritsky 2006; (16) Stanimirović et al. 2004; (17) Koposov et al. 2011; (18) Walker et al. 2007; (19) Wilkinson et al. 2004; (20) Walker et al. 2009c; (21) Walker et al. 2009b; (22) Walker et al. 2008; (23) Carignan et al. 1998; (24) Adén et al. 2009a; (25) Bouchard et al. 2006; (26) Belokurov et al. 2008; (27) Walker et al. 2009a; (28) Mateo et al. 2008; (29) Bender et al. 1996; (30) Collins et al. 2010; (31) Geha et al. 2006b; (32) Young & Lo 1997a; (33) Kalirai et al. 2010; (34) Blitz & Robishaw 2000; (35) Collins et al. 2011; (36) Geha et al. 2010; (37) Letarte et al. 2009; (38) Shostak & Skillman 1989; (39) Wilcots & Miller 1998; (40) Cook et al. 1999; (41) Young & Lo 1997b; (42) Lake & Skillman 1989; (43) Silich et al. 2006; (44) Young et al. 2007; (45) Koribalski et al. 2004; (46) Weldrake et al. 2003; (47) Lewis et al. 2007; (48) Young et al. 2003; (49) Hoffman et al. 1996; (50) Ryan-Weber et al. 2008; (51) Kepley et al. 2007; (52) Leaman et al. 2009; (53) Brown et al. 2007; (54) Young & Lo 1996; (55) Fraternali et al. 2009; (56) Bellazzini et al. 2011; (57) Jobin & Carignan 1990; (58) Barnes & de Blok 2001; (59) Blais-Ouellette et al. 2001; (60) Epinat et al. 2008; (61) Skillman et al. 1988; (62) Begum et al. 2005; (63) Bouchard et al. 2005; (64) Puche et al. 1991; (65) Westmeier et al. 2011; (66) Hlavacek-Larrondo et al. 2011; (67) Begum et al. 2008b; (68) Longmore et al. 1982; (69) Begum et al. 2006; (70) Begum et al. 2008a; (71) Stil & Israel 2002a; (72) Stil & Israel 2002b; (73) Stil et al. 2005; (74) Karachentsev et al. 2001a; (75) Simpson & Gottesman 2000; (76) Huchtmeier & Richter 1986. bSee Putman et al. (2004) for possible detection of gas in the Sagittarius stream. cUpper limit (uncorrected for binaries) and with a possible member star removed that otherwise boosts the dispersion to ∼5.5 km s−1. dVaries from 5 to 40 km s−1 across face of SMC. eAmbiguous detection of two H i clouds offset from optical center that could belong to the Magellanic Stream or Sculptor Group. fAmbiguous H i detection of cloud offset from optical center that could be Galactic H i emission. gEstimated from mean velocities of seven prominent H i clumps. hSee their Note added in proof. iApparent north–south gradient of 20km s−1 due to two distinct H i clumps. jUncorrected for possible rotational signature. kCalculated following Huchtmeier & Richter (1988).

Download table as:  ASCIITypeset images: 1 2 3 4

Column 1. Galaxy name.

Column 2. Mass of the galaxy in stellar masses, assuming a stellar mass-to-light ratio of 1. This has been adopted for simplicity and to allow easy scaling to whatever mass-to-light ratio is preferred (in particular, by adopting values estimated through near-infrared observations).

Column 3. The observed velocity dispersion of the stellar component and its uncertainty, generally based on multiple velocity measurements of individual (giant) stars, otherwise commented on in Column 10. Note that a single value for the velocity dispersion is often insufficient to describe the dynamics of the galaxy (e.g., multiple components and/or radial trends).

Column 4. The observed rotational velocity of the stellar component and its uncertainty. Where available, I quote the peak (observed) rotational velocity, otherwise the maximum observed rotational velocity.4 In general, I try to give the rotational velocity uncorrected for inclination or asymmetric drift, although it is sometimes unclear in the original papers whether these corrections have been applied. "N/A" indicates that a rotational signature has been looked for and none has been observed. Where only weak gradients in velocity are detected and/or the rotational signature is ambiguous, I comment on this in Column 10. In most cases, it is probably safest to interpret the absence of a rotational signature as the absence of any rotation with a magnitude equal to or greater than the magnitude of the velocity dispersion.

Column 5. Mass of H i in each galaxy. Here, I have scaled the values cited in the relevant papers to the distance given in Column 6 of Table 2. Where only an upper limit from a non-detection is available and that upper limit is small in comparison to the stellar mass, I give the H i mass as zero.

Column 6. The observed velocity dispersion of the H i component and its uncertainty. Note that the H i velocity dispersion is affected by local processes, such as heating due to star formation, and so caution is urged in interpreting these numbers. Note also that many galaxies are observed to contain multiple H i components (e.g., Lo et al. 1993; Young & Lo 1996, 1997a, 1997b; Young et al. 2003).

Column 7. The observed rotational velocity of the H i component and its uncertainty. Where available, I quote the peak (observed) rotational velocity, otherwise the maximum rotational velocity (see footnote 4). In general, I try to give the rotational velocity uncorrected for inclination or asymmetric drift, although it is sometimes unclear in the original papers whether these corrections have been applied. "N/A" indicates that a rotational signature has been looked for and none has been observed. In most cases, it is probably safest to interpret the absence of a rotational signature as the absence of any rotation with a magnitude equal to or greater than the magnitude of the velocity dispersion.

Column 8. Dynamical mass within the observed half-light radius. I have derived this quantity following the relation of Walker et al. (2009c), where Mdyn(⩽ rh) = 580 rh σ2 (see also Wolf et al. 2010) and so values are only given when both a stellar velocity dispersion and a half-light radius are available. Here, I adopt the value of the half-light radius measured on the semimajor axis. Dynamical masses can, of course, be estimated for all systems with measured structures and kinematics, but it is beyond the scope of this article to do so. While limited, the dynamical mass estimates in Column 8 are at least relatively homogeneous and hence comparable.

Column 9. References.

Column 10. Comments.

Figure 8 shows the mass of H i relative to visual luminosity for galaxies in the Local Group and its immediate vicinity, expressed in solar units, as a function of distance to either the MW or M31 (whichever is closer). Blue diamonds indicate galaxies with confirmed H i content. Orange arrows indicate the separations of gas-deficient galaxies, where the symbols for Sculptor and Fornax indicate that the presence of H i in these galaxies is ambiguous. Symbol size is proportional to absolute visual magnitude. The vertical dashed line indicates the approximate virial radius of the two host galaxies. Also indicated on the top axis is the "free-fall time," which is the time required for the galaxy to fall from rest at its current position and reach the host galaxy, acted upon only by the gravity of the giant galaxy,

Equation (1)

Dwarf galaxies with short free-fall timescales have likely had many chances to interact with the MW or M31, whereas galaxies with longer free-fall timescales likely have not had the time to complete many orbits. For systems where the free-fall time is of order or longer than a Hubble time (indicated with a dot-dashed line in Figure 8), it is reasonable to assume that interactions with giant galaxies have had a negligible effect on their evolution.

Figure 8.

Figure 8. H i fraction, expressed as MH i/LV in solar units, as a function of proximity to a giant galaxy. Blue diamonds indicate galaxies with confirmed H i content. Orange arrows indicate the separation of gas-deficient galaxies from either the MW or M31. The symbols for Sculptor and Fornax indicate that the presence of H i in these galaxies is ambiguous. Symbol size is proportional to absolute visual magnitude. The vertical dashed line indicates the approximate virial radius of the MW/M31. Also indicated on the top axis is the time required for the galaxy to accelerate from rest and reach the MW/M31. The dot-dashed line indicates the separation at which this time equals a Hubble time.

Standard image High-resolution image

Figure 8 shows the well-noted Local Group position–morphology relation, first highlighted by Einasto et al. (1974), in which the gas-deficient dSph galaxies are preferentially found in close proximity to either the MW or M31. Gas-rich dwarfs are preferentially located far from either of these two galaxies. Similar behavior has been noted in other galaxy groupings (e.g., Skillman et al. 2003b; Geha et al. 2006a; Bouchard et al. 2009, and references therein), and this has long been interpreted as evidence of the importance of external effects, such as tidal or ram-pressure stripping, on the evolution of dwarf galaxies. Indeed, many simulations show that the combined effects of these processes are in principle sufficient to remove the gas from dIrr-type galaxies (e.g., Mayer et al. 2006). However, Grebel et al. (2003) argue that dSph galaxies are too metal-rich for their luminosity in comparison to the (old) stellar populations in dIrr galaxies; if so, then this requires that the early metallicity evolution of the two morphological types (and not just their environments) were also different.

The visual appearance of dIrr galaxies is usually very different from that of dSph systems and reflects the presence of very young stellar populations and ongoing star formation in the former. Some dwarf galaxies are additionally classified as "dIrr/dSph": these so-called transition systems, such as DDO 210, distinguish themselves from dIrrs primarily through the absence of current star formation (formally, detectable H i but no detectable H ii regions), despite having very young stellar populations. While star formation in some of these galaxies may have ceased permanently, in others it may be possible that low-level star formation is continuing without the production of H ii regions (Fumigalli et al. 2011). Alternatively, ongoing, low-level star formation may be interrupted by periods of inactivity ("gasping"; see Tosi et al. (1991) for an example), in which case the distinction between a dIrr and a transition system merely reflects the moment we happen to be observing it.

The HST/ANGST survey (Dalcanton et al. 2009) has recently derived star formation histories (SFHs) for many nearby dwarf galaxies, including a large number of the more distant galaxies in this compilation. They show a tremendous diversity in the SFHs, and highlight that in a statistical sense there is essentially no difference between the SFHs of dIrr galaxies and the transition systems (Weisz et al. 2011). They suggest that many of the transition dwarfs are perhaps best treated as low mass dIrr galaxies, and if any of the transition dwarfs are deserving of the moniker it is likely the less isolated, low gas fraction systems. Weisz et al. (2011) also demonstrate that the "average" SFHs of dIrrs and dSphs in their sample are actually similar over most of cosmic time. It is only within the past few Gyr, and particularly within the past 1 Gyr, that the SFHs differ. However, the HST/ANGST survey generally samples only bright dSphs (comparable in luminosity to Fornax, for example), and low-luminosity systems comparable to Ursa Minor, Draco, and their ilk are either absent or underrepresented (due to the difficulty of finding such galaxies outside of the Local Group). The reader is referred to the HST/LCID survey (Monelli et al. 2010a, 2010b; Hidalgo et al. 2011, and references therein) that is presently investigating the possible connections between dIrr, transition, and dSph using isolated dwarfs within the Local Group.

A final distinction between dIrr and dSph galaxies is traditionally made based on their kinematics, although this is not necessarily well motivated from an observational perspective, particularly at faint magnitudes. It is generally argued that dIrr galaxies are rotationally supported systems, whereas dSph galaxies are pressure-supported systems. However, as is visible in Table 4, the majority of dynamical studies of dSph galaxies are based on resolved spectroscopy of individual (giant) stars. Until recently, essentially all dynamical studies of Local Group dIrr galaxies were based on H i, which generally shows significant rotation. However, recent work by Leaman et al. (2009) has shown that the evolved RGB population in WLM has (vr/σ) ≃ 1.2, whereas (vr/σ)H i ≃ 6.7. Thus, while it may be true that the stellar component of dIrrs shows significant rotation compared to dSphs, caution should be exercised in comparing H i kinematics to those of evolved stellar populations.

Figure 9 plots the stellar velocity dispersion given in Table 4 against the half-light radius for all galaxies for which this information exists. A broad correlation exists, such that colder systems have smaller scale sizes, although the scatter in this relationship is roughly an order of magnitude at a given scale radius. Measuring some of these values is challenging: Martin et al. (2008b) and Muñoz et al. (2012) both discuss the problems of determining structural parameters for galaxies in which few bright stars can be identified (generally the case for galaxies much fainter than MV ≃ −8). For velocity dispersions, too, the lack of a large number of bright spectroscopic targets and the low expected dispersions make these measurements particularly challenging. Typically, the velocity errors on individual stellar measurements that are used as a basis for deriving the overall dispersion are a few km s−1, which is of order the measured dispersion for some of the galaxies in Figure 9. One should also consider that the expected velocity dispersion of a faint system like Segue, due only to its baryonic mass, is exceptionally small, at around σ ≃ 0.1 km s−1 (McConnachie & Côté 2010).

Figure 9.

Figure 9. Stellar velocity dispersion vs. the (geometric-mean) half-light radius measured for all galaxies for which this information is available.

Standard image High-resolution image

There are a couple of primary sources of contamination when measuring velocity dispersions, all of which have been shown to bias the measured "intrinsic" dispersion of faint galaxies to higher values. Foreground contamination can have a significant effect. The first measurement of the velocity dispersion of the Hercules dwarf gave σ = 5.1 ± 0.9 km s−1 (Simon & Geha 2007). However, Adén et al. (2009a) obtained Stromgren photometry for stars in the direction of Hercules and were able to separate member stars from interloping dwarf stars in the halo of the Galaxy. Without doing this separation, Adén et al. (2009a) measure a value of σ = 7.3 ± 1.1 km s−1 (in broad agreement with Simon & Geha 2007). By additionally being able to reject foreground dwarfs, they find a considerably lower value of σ = 3.7 ± 0.9 km s−1 (see also Adén et al. 2009b).

Binary stars are a further source of known contamination when measuring the velocity dispersions of dwarf galaxies, where the reflex motion of an unresolved binary star is mistaken as a component of the intrinsic dispersion of the galaxy. This potential source of contamination was originally investigated prior to the identification of extremely faint dwarf galaxies to determine if it could inflate the measured velocity dispersion sufficiently to incorrectly infer the need for non-baryonic mass-to-light ratios. Numerous studies (e.g., Aaronson & Olszewski 1987; Mateo et al. 1993; Olszewski et al. 1996; Hargreaves et al. 1996) all concluded that, barring a pathological or extremely high binary population, binary stars were insufficient to boost a stellar system with σ ≲ 1 km s−1 to σ ≃ 8–10 km s−1, although some inflation of the dispersion is to be expected. Minor et al. (2010) have developed a methodology to statistically correct the velocity dispersions of dwarf galaxies for this binary inflation; their own study indicates that systems with intrinsic dispersions of 4–10 km s−1 are unlikely to have their dispersion inflated by more than 20% as a result of binaries. For systems with lower intrinsic dispersions than this, McConnachie & Côté (2010) argue that the inflation effect due to binaries could be significant, although this depends on the as-yet-unmeasured stellar binary fraction(s) in dwarf galaxies.

Recently, Koposov et al. (2011) developed a procedure for obtaining accurate radial velocities with demonstrably reliable uncertainties at faint magnitudes and conducted a multi-epoch spectroscopic study of the Bootes dwarf galaxy. Approximately 10% of the stars in their sample show evidence for velocity variability. Their data favor a two-component model for Bootes (see discussion in Section 4), where the central, dominant component has a velocity dispersion of σ = 2.4+0.9−0.5 km s−1. Their preferred value is notably lower than earlier values of order σ ≃ 6 km s−1 reported by Muñoz et al. (2006) and Martin et al. (2007) (although they are in statistical agreement). These results all emphasize the difficulties in obtaining reliable estimates of the intrinsic velocity dispersions of very faint dwarf galaxies: reliable (and small!) velocity uncertainties, multi-epoch data, and dwarf-giant separation are all essential components to any successful measurement.

With the above caveats on the basic dynamical data, Figure 10 shows the dynamical mass estimates for each dwarf galaxy within its half-light radius, as a function of absolute visual magnitude. Here, I have used the formalism by Walker et al. (2009c); Wolf et al. (2010) justify in detail why the mass within the half-light radius is in general able to be reliably measured for dispersion-supported spheroids. The left panel highlights by name each galaxy for which the necessary data exist, whereas the right panel shows the formal uncertainties on propagating the quoted uncertainties on σ and rh. Note that these estimates do not take into consideration errors introduced as a result of the dynamical state of the galaxy. For example, a few galaxies in Figure 10 show some rotation in addition to dispersion support (M32, NGC 205, NGC 147, NGC 185, Cetus, WLM), and other galaxies may or may not be in dynamical equilibrium. Further, the Walker et al. formalism assumes that the velocity dispersion profile is flat with radius (the Wolf et al. formalism adopts a luminosity-weighted mean velocity dispersion); however, for many galaxies shown in Figure 10 the form of the radial velocity dispersion profile is unknown. A full study of the dynamics of these galaxies is clearly outside the scope of this article, and the current estimates are provided solely to demonstrate the status of the observational data in terms of its basic implications regarding the mass content of the galaxies. Figure 11 shows the same data as Figure 10, except I have now normalized Mdyn by 0.5 L to estimate the implied mass-to-light ratio of each system within its half-light radius.

Figure 10.

Figure 10. Dynamical mass estimates within the half-light radius vs. absolute visual magnitude, for all galaxies for which stellar velocity dispersion and half-light radius information is available. These have been calculated using the relation from Walker et al. (2009c). Individual galaxies are identified in the left panel, and error bars are indicated in the right panel. Note that the errors were calculated by propagating the uncertainties on the stellar velocity dispersion and half-light radius in the usual way and do not account for any systematic uncertainties in either these quantities or the dynamical state of the galaxy.

Standard image High-resolution image
Figure 11.

Figure 11. Mass-to-light ratio, in solar units, calculated within the half-light radius for all dwarf galaxies for which the necessary data exist, as a function of absolute visual magnitude.

Standard image High-resolution image

Figures 10 and 11 demonstrate that there appears to be a well-defined relation consistent with a power law between Mdyn(⩽ rh) and MV that holds across a factor of nearly 1 million in luminosity, in agreement with the finding reported by Walker et al. (2009c). As discussed above, systematic errors in the measurements of the relevant quantities, as well as the necessary assumptions that must be made in converting these quantities to a mass estimate, mean that the precise position of any single point in these plots is open to considerable debate. Further, it should be stressed that observational selection effects such as that discussed for Figure 6 (which may be responsible for the precise form of the MVrh relationship at low luminosities) will also affect these figures due to the dependence of Mdyn on rh.

Measurements such as those on which Figures 10 and 11 are based have also been used to argue for a common mass scale for the (MW) dwarfs (Strigari et al. 2007). However, this is usually made with reference to the mass enclosed with an absolute radial scale (typically r = 300 pc), a quantity that is used because of its usefulness in dark-matter-only simulations, where there is no information regarding the size of the baryonic components of galaxies. However, given that the half-light radii of the dwarfs in and around the Local Group have 20 ≲ rh ≲ 2000 pc, from an observational perspective this is a far less well-defined quantity than Mdyn(⩽ rh).

6. MEAN STELLAR METALLICITIES

Table 5 lists the available data for the mean stellar metallicities, 〈[Fe/H]〉, of the galaxies.

Table 5. Mean Stellar Metallicities

(1) (2) (3) (4) (5)
Galaxy 〈[Fe/H]〉 Technique References Comments
The Galaxy          
Canis Major −0.5 0.2 CMD fitting Bellazzini et al. (2004c) Based primarily on MS stars
Sagittarius dSph −0.4 0.2 HRS Chou et al. (2007)a  
Segue (I) −2.72 0.40 MRS Norris et al. (2010) See also Simon et al. (2011)
Ursa Major II −2.47 0.06 MRS Kirby et al. (2008b, 2011)  
Bootes II −1.79 0.05 MRS (CaT) Koch et al. (2009)  
Segue II −2.00 0.25 HRS Belokurov et al. (2009)  
Willman 1 −2.1:   MRS Willman et al. (2011) Based on three stars
Coma Berenices −2.60 0.05 MRS Kirby et al. (2008b, 2011)  
Bootes III −2.1 0.2 MRS Carlin et al. (2009)  
LMC −0.5   MRS (CaT) Carrera et al. (2008) See also Cole et al. (2005)
SMC −1.00 0.02 MRS (CaT) Parisi et al. (2010)  
Bootes (I) −2.55 0.11 MRS Norris et al. (2010) See also Lai et al. (2011)
Draco −1.93 0.01 MRS Kirby et al. (2011) See also Winnick (2003)
Ursa Minor −2.13 0.01 MRS Kirby et al. (2011) See also Winnick (2003)
Sculptor −1.68 0.01 MRS Kirby et al. (2009, 2011) See also Battaglia et al. (2008)
Sextans (I) −1.93 0.01 MRS Kirby et al. (2011) See also Battaglia et al. (2011)
Ursa Major (I) −2.18 0.04 MRS Kirby et al. (2008b, 2011)  
Carina −1.72 0.01 MRS (CaT) Koch et al. (2006)  
Hercules −2.41 0.04 MRS Kirby et al. (2008b, 2011)  
Fornax −0.99 0.01 MRS Kirby et al. (2011) See also Battaglia et al. (2006)
Leo IV −2.54 0.07 MRS Kirby et al. (2008b, 2011) HRS of a single star in Simon et al. (2010)
Canes Venatici II −2.21 0.05 MRS Kirby et al. (2008b, 2011)  
Leo V −2.00 0.2 Isochrones de Jong et al. (2010)  
Pisces II −1.9:   Isochrones Belokurov et al. (2010)  
Canes Venatici (I) −1.98 0.01 MRS Kirby et al. (2008b, 2011)  
Leo II −1.62 0.01 MRS Kirby et al. (2011) See also Koch et al. (2007a)
Leo I −1.43 0.01 MRS Kirby et al. (2011) See also Koch et al. (2007b)
Andromeda          
M32 −0.25   Isochrones Grillmair et al. (1996) Likely significant age spread and gradients?
Andromeda IX −2.2 0.2 Co-added MRS (CaT) Collins et al. (2010)  
NGC 205 −0.8 0.2 Isochrones McConnachie et al. (2005)  
Andromeda XVII −1.9 0.2 Isochrones Brasseur et al. (2011b)  
Andromeda I −1.45 0.04 Isochrones Kalirai et al. (2010)  
Andromeda XXVII −1.7 0.2 Isochrones Richardson et al. (2011)  
Andromeda III −1.78 0.04 Isochrones Kalirai et al. (2010)  
Andromeda XXV −1.8 0.2 Isochrones Richardson et al. (2011)  
Andromeda XXVI −1.9 0.2 Isochrones Richardson et al. (2011)  
Andromeda XI −2.0 0.2 Co-added MRS (CaT) Collins et al. (2010)  
Andromeda V −1.6 0.3 Co-added LRS (CaT) Collins et al. (2011)  
Andromeda X −1.93 0.11 Isochrones Kalirai et al. (2010)  
Andromeda XXIII −1.8 0.2 Isochrones Richardson et al. (2011)  
Andromeda XX −1.5 0.1 Isochrones McConnachie et al. (2008)  
Andromeda XII −2.1 0.2 Co-added MRS (CaT) Collins et al. (2010)  
NGC 147 −1.1 0.1 MRS (CaT) Geha et al. (2010)  
Andromeda XXI −1.8 0.2 Isochrones Martin et al. (2009)  
Andromeda XIV −2.26 0.05 Isochrones Kalirai et al. (2010)  
Andromeda XV −1.8 0.2 Co-added MRS (CaT) Letarte et al. (2009)  
Andromeda XIII −1.9 0.2 Co-added MRS (CaT) Collins et al. (2010)  
Andromeda II −1.64 0.04 Isochrones Kalirai et al. (2010)  
NGC 185 −1.3 0.1 MRS (CaT) Geha et al. (2010)  
Andromeda XXIX −1.8:   Isochrones Bell et al. (2011)  
Andromeda XIX −1.9 0.1 Isochrones McConnachie et al. (2008)  
Triangulum          
Andromeda XXIV −1.8 0.2 Isochrones Richardson et al. (2011)  
Andromeda VII −1.40 0.30 Isochrones Kalirai et al. (2010)  
Andromeda XXII −1.8:   Isochrones Martin et al. (2009)  
IC 10 −1.28:   Isochrones/RGB color Tikhonov & Galazutdinova (2009)  
LGS 3 −2.10 0.22 RGB color Lee (1995)  
Andromeda VI −1.3 0.14 Co-added MRS (CaT) Collins et al. (2011)  
Andromeda XVI −2.1 0.2 Co-added MRS (CaT) Letarte et al. (2009)  
Andromeda XXVIII ... ...      
IC 1613 −1.6 0.2 CMD modeling Bernard et al. (2010) Derived in E. Skillman et al., (2012 in preparation)
Phoenix −1.37 0.2 RGB color Martínez-Delgado et al. (1999)  
NGC 6822 −1.0 0.5 LRS (CaT) Tolstoy et al. (2001)  
Cetus −1.9 0.10 MRS (CaT) Lewis et al. (2007)  
Pegasus dIrr −1.4 0.2 Isochrones McConnachie et al. (2005)  
Leo T −1.99 0.05 MRS Kirby et al. (2008b, 2011)  
WLM −1.27 0.04 MRS (CaT) Leaman et al. (2009)  
Leo A −1.4 0.2 CMD modeling Cole et al. (2007)  
Andromeda XVIII −1.8 0.1 Isochrones McConnachie et al. (2008)  
Aquarius −1.3 0.2 Isochrones McConnachie et al. (2006) Assumes an age of 4 Gyr; for an ancient population, find [Fe/H] ≃ −1.9
Tucana −1.95 0.15 MRS (CaT) Fraternali et al. (2009)  
Sagittarius dIrr −2.1 0.2 Isochrones Momany et al. (2002)  
UGC 4879 −1.5 0.2 Isochrones Bellazzini et al. (2011)  
NGC 3109 −1.84 0.2 RGB color Hidalgo et al. (2008)  
Sextans B ... ...      
Antlia −1.6 0.1 RGB color Aparicio et al. (1997)  
Sextans A −1.85:   RGB color Sakai et al. (1996) Adopted (VI)−3.5 = 1.3
HIZSS 3(A) ... ...      
HIZSS 3B ... ...      
KKR 25 −2.1 0.3 RGB color Karachentsev et al. (2001b)  
ESO 410- G 005 −1.93 0.2 RGB color Sharina et al. (2008)  
NGC 55          
ESO 294- G 010 −1.48 0.17 RGB color Sharina et al. (2008)  
NGC 300          
IC 5152 ... ...      
KKH 98 −1.94 0.25 RGB color Sharina et al. (2008)  
UKS 2323-326 −1.68 0.19 RGB color Sharina et al. (2008)  
KKR 3 −2.02 0.25 RGB color Sharina et al. (2008)  
GR 8 ... ...      
UGC 9128 −2.33 0.24 RGB color Sharina et al. (2008)  
UGC 8508 −1.91 0.19 RGB color Sharina et al. (2008)  
IC 3104 ... ...      
DDO 125 −1.73 0.17 RGB color Sharina et al. (2008)  
UGCA 86 ... ...      
DDO 99 −2.13 0.22 RGB color Sharina et al. (2008)  
IC 4662 −1.34 0.13 RGB color Sharina et al. (2008)  
DDO 190 −2.00 0.08 RGB color Aparicio & Tikhonov (2000)  
KKH 86 −2.33 0.29 RGB color Sharina et al. (2008)  
NGC 4163 −1.65 0.15 RGB color Sharina et al. (2008)  
DDO 113 −1.99 0.21 RGB color Sharina et al. (2008)  

Note. aAnd references therein.

Download table as:  ASCIITypeset images: 1 2

Column 1. Galaxy name.

Column 2. Mean stellar metallicity and error, generally derived from observations of mostly evolved giants, otherwise commented on in Column 5. Note that I quote the mean metallicity of all stars (as opposed to the median, mode, etc.) in the observed metallicity distribution function, and so this will be susceptible to any uncorrected biases in the observed sample. The mean stellar metallicity refers to the mean iron-peak metallicity expressed in logarithmic notation relative to the solar value.

Column 3. Technique used to estimate stellar metallicity. A discussion is given below.

Column 4. References.

Column 5. Comments.

Figure 12 shows mean stellar metallicity as a function of absolute visual magnitude, where by definition both axes are logarithmic quantities. In the left panel, all tabulated data from Table 5 are given, whereas in the right panel only data that use spectroscopically determined quantities are shown.

Figure 12.

Figure 12. Absolute visual magnitude vs. stellar metallicity measurement (as listed in Table 5), for all galaxies in the sample for which this information is available. Color-coding is the same as in previous figures. The left panel shows all measurements regardless of the techniques used to estimate metallicity, whereas the right panel only shows those measurements derived from spectroscopic observations of resolved stars (typically giants).

Standard image High-resolution image

Generally the most well-accepted method of determining [Fe/H] for a given star is from direct high-resolution spectroscopy (HRS) from which the strengths of iron lines can be measured directly. However, Table 5 lists estimates for 〈[Fe/H]〉 that stem from a variety of additional techniques.

  • 1.  
    RGB color. Da Costa & Armandroff (1990) derive a relation between the mean color of RGB stars in globular clusters measured at MV = −3 and the mean stellar metallicity of the cluster. A refined calibration (derived at MV = −3.5) using the same data was presented by Lee et al. (1993), and this has been used throughout the literature as a convenient means of estimating the mean stellar metallicity of a system when only photometry of bright stars is available. An obvious disadvantage, of course, is that it is calibrated on ancient stellar populations. Since RGB stars are subject to age and metallicity degeneracies, this technique may not be reliable if significantly younger stars are present. I note that in Table 5 and Figure 12 most of the non-Local-Group galaxies have their mean stellar metallicity determined in this way. There is a suggestion in Figure 12 that these estimates are low in comparison to Local Group galaxies with similar absolute visual magnitudes, but its cause or significance is difficult to determine.
  • 2.  
    Isochrones. The comparison of theoretical isochrones, evolutionary tracks, or globular cluster fiducials, to the CMDs of galaxies is usually heavily or entirely weighted toward the color of the RGB and so is, in principle, subject to the same systematic uncertainties regarding ages as discussed above. In addition, isochrones produced by different groups may have some systematic differences between the predicted RGB colors at any age. However, one potential advantage of isochrones over empirical RGB color relations is that, if additional information is available regarding ages (e.g., from the red clump or main-sequence turnoff), isochrones with appropriate ages can be used.
  • 3.  
    CMD fitting. The reader is referred to the many excellent papers on SFH analysis using resolved stellar populations for more details (e.g., Gallart et al. 2005; Tolstoy et al. 2009, and references therein). Here, age degeneracies will likely be much reduced since age information is available from elsewhere in the CMD (in particular the main-sequence turnoff region).
  • 4.  
    Calcium II triplet equivalent widths (low, moderate resolution spectroscopy). The strengths of the three CaT absorption lines at λ = 8498, 8542, and 8662 Å in RGB stars have been shown to correlate with [Fe/H] if the temperature and surface gravity of the star are able to be estimated through measurement of its absolute magnitude. The CaT is a relatively strong feature, and consequently it can be relatively easily measured using low- to moderate-resolution spectroscopy (L/MRS). Many authors (e.g., Rutledge et al. 1997) have developed calibrations between CaT equivalent width and [Fe/H], most recently Starkenburg et al. (2010). These authors also examine the form of the correlation at [Fe/H] ≲ −2.5. Using the new calibration, Starkenburg et al. (2010) demonstrate that the resulting estimates of [Fe/H] match those derived from HRS down to a limiting stellar metallicity of [Fe/H] ≃ −4. Many of the values listed in Table 5 use the earlier calibrations, which can potentially deviate from the new calibration for stellar metallicities lower than [Fe/H] ≃ −2, depending on the absolute magnitudes of the observed stars.
  • 5.  
    Spectral synthesis (moderate-resolution spectroscopy). Here, synthetic spectra spanning a range of key parameters (e.g., [Fe/H], [α/Fe], Teff, log g) are compared to the measured stellar spectrum to determine the most likely [Fe/H]. Kirby et al. (2008a) have developed this technique for specific use on the dwarf galaxies of the MW with considerable success. Here, the main advantages are that one uses more spectral information than relying only on specific indices. Since one uses moderate resolutions, fainter targets can be studied than is possible with HRS.

Clearly, the range of techniques from which 〈[Fe/H]〉 has been derived in Table 5 may produce numerous systematic differences between results. Recent discussions of some of these issues can be found in Kirby et al. (2008a) (spectral synthesis), Starkenburg et al. (2010) (CaT), and Lianou et al. (2011) (isochrones, CaT), and the reader is referred to these papers for more details. I note that several of the 〈[Fe/H]〉 values listed in Table 5 only quote the uncertainty from the random error in the mean (i.e., ${\sigma _{[{\rm Fe/H}]}}/{\sqrt{N}}$) and do not include the additional systematic uncertainties (which cannot be treated in the same way). For galaxies for which this is the case, an additional systematic error component must be included, typically a tenth of a dex or more.

In addition to the systematic uncertainties introduced by the measurement technique, an additional systematic effect enters the calculation of the mean stellar metallicities in Table 5 through the spatial extent of the stars contributing to the measurement. Dwarf galaxies are known to exhibit radial gradients in stellar populations, ages, and/or metallicity distributions (e.g., Harbeck et al. 2001; Tolstoy et al. 2004; Battaglia et al. 2006; Bernard et al. 2008, and references therein). Thus far, all observed metallicity gradients are such that a higher "mean" stellar metallicity is observed in the central regions of the dwarf than in the outer regions. Two galaxies with identical metallicity distribution functions and radial gradients will therefore be measured to have different "mean" stellar metallicities if there is any difference in the radial distribution of stars that contribute to the measurement. Given the range in radial scale associated with the galaxies in Table 5 and the unknown spatial structure of the stellar populations/ages/metallicities within many of these systems, this effect may be significant.

With the above caveats, both panels of Figure 12 show the well-known luminosity–mean stellar metallicity relation, whereby more luminous galaxies are generally more metal-rich in the mean than faint galaxies (see also Tremonti et al. 2004; Lee et al. 2006; Kirby et al. 2008b, and references therein). The relation is better defined in the right panel, which uses only spectroscopic measurements of 〈[Fe/H]〉; the degree to which the scatter in the left panel is a result of systematic errors (age effects, measurement techniques, etc.) is unknown. As pointed out by Kirby et al. (2008b), the luminosity–mean stellar metallicity relation at faint magnitudes (MV > −8) is entirely consistent with a continuation of the trend displayed for bright galaxies. However, it is also worth highlighting the dearth of data at faint magnitudes: if only galaxies fainter than MV = −8 are considered, then the data are equally consistent with a scatter around a floor in mean stellar metallicity at 〈[Fe/H]〉 ≃ −2.1. While more data are certainly needed, it is interesting to note that the magnitude at which one can start arguing for a floor in mean stellar metallicity is broadly the same point at which there is a break in the absolute magnitude–surface-brightness relation shown in Figure 7 (discussed in Section 4). While certainly speculative, this may suggest a direct link between the surface brightness of a galaxy (specifically, the density of baryons, rather than the total amount of baryons) and its mean stellar metallicity (see theoretical work by Dekel & Woo 2003; Revaz & Jablonka 2012, and references therein; see also Skillman et al. 2003a).

Many more parameters are required in order to fully define the metallicity distribution function of stars in any of the dwarf galaxies, and the reader is referred to Tolstoy et al. (2009) for a more complete discussion of ongoing observational efforts in this regard. As discussed in Tolstoy et al. (2009), detailed chemistry, such as abundance ratios of key elements, is now available for the nearest galaxies, with much more to come. In the near future, the combination of dynamical data with detailed information on the chemical evolution of galaxies is likely to continue to be a significant area of discovery and advance.

7. SUMMARY

Tables 15 provide observational parameters for approximately 100 nearby dwarf galaxies, including positional data; radial velocities; photometric and structural parameters; stellar, H i, and dynamical masses; internal kinematics (dispersions and rotation); and mean stellar metallicities. In addition to discussing the provenance of the values and possible sources of uncertainties affecting their usage, I also consider the membership and (limited?) spatial extent of the MW and M31 sub-groups; the morphological diversity of the M31 sub-group in comparison to the MW; the location of the zero-velocity surface of the Local Group; the timescales that can be associated with the orbital/interaction histories of the Local Group members; the scaling relations defined by the sample, including their behavior at the faintest magnitudes/surface brightnesses; and the luminosity–mean stellar metallicity relation, including its possible connection to the central surface brightness of the galaxies.

The dwarf galaxies discussed herein constitute all known satellite systems in the surroundings of the MW and M31, all quasi-isolated systems in the more distant reaches of the Local Group, and all other, generally very isolated, dwarf galaxies that surround the Local Group out to 3 Mpc, beyond which most of the known galaxies are members of Maffei, Sculptor, Canes Venatici, IC 342, M81, and the many other groups in the Local Volume found out to the Virgo Cluster and beyond. In some respects, these systems are stepping stones into the Local Volume, and we are fortunate that they present us with a large and diverse range of properties to explore. This article has purposefully focused entirely on observable, quantifiable properties of these galaxies, and it is right that the published tables likely already contain obsolete numbers or do not include the most recent discoveries.5,6 Our knowledge of many of these systems far exceeds that of other galaxies, and we will continue to be able to obtain ever more detailed observations of aspects of their structure, dynamics, stellar populations, and chemical signatures that are beyond the reach of observational programs of more distant galaxies. In the future as now, these are the galaxies for which we will know the most and so are perhaps destined to understand the least.

Updated versions of the core data available in Tables 15 for all dwarf galaxies within 3 Mpc are available at https://www.astrosci.ca/users/alan/Nearby_Dwarfs_Database.html. I thank Peter Stetson, Matthew Walker, Daniel Weisz, Sidney van den Bergh, and Else Starkenburg for careful readings of the manuscript and valuable comments; Ryan Leaman for his feedback, and for stressing the importance of radial metallicity gradients when considering mean stellar metallicities, which led to the addition of the relevant paragraph; and Josh Simon for highlighting relevant references. In addition, I am grateful to Vasily Belokurov, Michelle Collins, Patrick Côté, Stephanie Côté, Julianne Dalcanton, Tim Davidge, Marla Geha, Karoline Gilbert, Carl Grillmair, Mike Irwin, Bradley Jacobs, Ryan Leaman, Nicolas Martin, Emma Ryan-Weber, Evan Skillman, Peter Stetson, Brent Tully, Sidney van den Bergh, Matthew Walker, Daniel Weisz, and Lisa Young for their help in answering questions that came up during the preparation of this manuscript; to Wolfgang Steinicke and Harold Corwin Jr. for their invaluable assistance in tracing some of the early references relating to the original discoveries of these objects; and to the referees for their supportive comments and very helpful feedback that led to significant improvements in the text. Finally, thanks go to the organizers of the KITP program "First Galaxies and Faint Dwarfs: Clues to the Small Scale Structure of Cold Dark Matter," at which this article was completed. As such, this research was supported in part by the National Science Foundation under Grant No. NSF PHY11-25915.

Footnotes

  • Specifically, the 70 galaxies listed in Table 2 of Rice et al. (1988) with fluxes greater than or equal to 0.6 Jy.

  • Note added in Proof: See also the recent study by Tollerud et al. (2012).

  • For observations that do not extend to sufficiently large Galactocentric radius, the maximum observed rotational velocity will not necessarily equal the peak rotational velocity of the galaxy.

  • Updated versions of the core data available in Tables 15 for all dwarf galaxies within 3 Mpc are available at https://www.astrosci.ca/users/alan/Nearby_Dwarfs_Database.html.

  • Note added in Proof: For example, see Papers by Tollerud et al. (2012) and Kirby et al. (2012).

Please wait… references are loading.
10.1088/0004-6256/144/1/4