NEW LIMITS ON AN INTERMEDIATE-MASS BLACK HOLE IN OMEGA CENTAURI. II. DYNAMICAL MODELS*

and

Published 2010 January 27 © 2010. The American Astronomical Society. All rights reserved.
, , Citation Roeland P. van der Marel and Jay Anderson 2010 ApJ 710 1063 DOI 10.1088/0004-637X/710/2/1063

0004-637X/710/2/1063

ABSTRACT

We present a detailed dynamical analysis of the projected density and kinematical data available for the globular cluster ω Centauri. We solve the spherical anisotropic Jeans equation for a given density profile to predict the projected profiles of the rms velocity ${\bar{\sigma }}(R)$, in each of the three orthogonal coordinate directions (line of sight, proper motion radial, and proper motion tangential). The models allow for the presence of a central dark mass, such as a possible intermediate-mass black hole (IMBH). We fit the models to new Hubble Space Telescope star count and proper motion data near the cluster center presented in the companion paper, which is Paper I in this series, combined with existing ground-based measurements at larger radii. The projected density profile is consistent with being flat near the center, with an upper limit γ ≲ 0.07 on the central logarithmic slope. The rms proper motion profile is also consistent with being flat near the center. The velocity anisotropy profile, distance, and stellar mass-to-light ratio are all tightly constrained by the data and found to be in good agreement with previous determinations by van de Ven et al. To fit the kinematics, we consider anisotropic models with either a flat core (γ = 0) or a shallow cusp (γ = 0.05). Core models provide a good fit to the data with MBH = 0; cusp models require a dark mass. If the dark mass in cusp models is an IMBH, then MBH = (8.7 ± 2.9) × 103M; if it is a dark cluster, then its extent must be ≲0.16 pc. Isotropic models do not fit the observed proper motion anisotropy and yield spuriously high values for any central dark mass. These models do provide a good fit to the Gauss–Hermite moments of the observed proper motion distributions (h4 = −0.023 ± 0.004, h6 = 0.001 ± 0.004). There are no unusually fast-moving stars observed in the wings of the proper motion distribution, but we show that this does not strongly constrain the mass of any possible IMBH. The overall end result of the modeling is an upper limit to the mass of any possible IMBH in ω Centauri: MBH ≲ 1.2 × 104M at ∼1σ confidence (or ≲1.8 × 104M at ∼3σ confidence). The 1σ limit corresponds to MBH/Mtot ≲ 0.43%. We combine this with results for other clusters and discuss the implications for globular cluster IMBH demographics. Tighter limits will be needed to rule out or establish whether globular clusters follow the same black hole demographics correlations as galaxies. The arguments put forward by Noyola et al. to suspect an IMBH in ω Centauri are not confirmed by our study; the value of MBH they suggested is firmly ruled out.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Intermediate-mass black holes (IMBHs) are of interest in a variety of astrophysical contexts (see, e.g., the reviews by van der Marel 2004; Miller & Colbert 2004). Especially the possible presence of IMBHs in the centers of globular clusters has intrigued astronomers for decades (e.g., Bahcall & Wolf 1976). This subject area underwent a considerable resurgence in recent years due to a combination of two advances. First, theoretical modeling suggested plausible formation scenarios for IMBHs in the centers of globular clusters from realistic initial conditions (Portegies Zwart & McMillan 2002). Second, the quality of observational data sets, especially those obtained with the Hubble Space Telescope (HST), advanced to the point where it is possible to measure velocity dispersions at a spatial resolution comparable to the sphere of influence size for plausible IMBH masses. This led to the identification of three globular clusters with candidate IMBHs, namely M15, G1, and ω Centauri.

Gerssen et al. (2002) used line-of-sight (LOS) velocity data for individual stars in the core-collapsed cluster M15 (NGC 7078) to argue for the presence of ∼3000 M in dark material near the center, possibly in the form of an IMBH. Dull et al. (1997) and Baumgardt et al. (2003a) showed that this could be a plausible result of mass segregation, and need not be an IMBH. Moreover, Baumgardt et al. (2005) argued that the observed steep brightness profile of M15, typical of the class of core-collapsed clusters, is in fact not consistent with expectation for a relaxed distribution of stars around an IMBH. Ho et al. (2003) and Bash et al. (2008) failed to detect X-ray or radio emission coincident with the cluster center. This is not necessarily inconsistent with an IMBH, since the accretion rate and/or the accretion efficiency are likely to be low. Nonetheless, M15 does not appear to be as strong an IMBH candidate as once thought.

Gebhardt et al. (2002, 2005) used the velocity dispersion measured from integrated light near the center of the M31 cluster G1 to argue for the presence of a (1.8 ± 0.5) × 104M dark mass at the cluster center. The relaxation time of G1 is too long to have produced such a concentration of dark mass due to mass segregation. Therefore, this result was taken as evidence of for the presence of an IMBH. Baumgardt et al. (2003b) disagreed with this interpretation, and argued that models with no dark mass (IMBH or otherwise) provided an equally acceptable fit. The crux of this disagreement lies in the extremely small size of the sphere of influence (GMBH2 ≈ 0farcs03) of the putative IMBH, which is less than the spatial resolution of the observations. Phrased differently, the IMBH induces only a small signature in the data. As a result, the interpretation of the data is sensitive to the exact details of the data–model comparison. The statistical significance of the IMBH detection is not high, especially if one takes into account the possibility of "black hole bias" (discussed later in Section 6.3). On the other hand, the possible presence of an IMBH in G1 recently gained further credence with the detection of weak X-ray and radio emission from near the cluster center (Pooley & Rappaport 2006; Kong 2007; Ulvestad et al. 2007). The emission properties are consistent with accretion onto an IMBH, but other explanations cannot be ruled out.

Noyola et al. (2008, hereafter NGB08) argued for the presence of an IMBH in the center of the cluster ω Centauri (NGC 5139), which is well known for its many enigmatic properties (e.g., Meylan 2002). The dynamics of this cluster were studied previously by van de Ven et al. (2006, hereafter vdV06) using sophisticated axisymmetric models. However, there was insufficient kinematical data at small radii to meaningfully constrain the possible presence of any dark mass. NGB08 performed integrated-light measurements of the LOS velocity dispersion for two 5'' × 5'' fields: one on the cluster center that they had determined and one at R = 14'' from that center. They obtained σlos = 23.0 ±  2.0 km s−1 and σlos = 18.6 ±  1.6 km s−1 for these fields, respectively. The increase in σlos toward the cluster center was interpreted to imply the presence of an IMBH of mass MBH = 4.0+0.75−1.0 × 104M. They also determined the surface brightness profile of ω Centauri from HST images, and inferred a shallow central cusp of logarithmic slope γ = 0.08 ± 0.03. This is consistent with theoretical predictions for the cusp induced by an IMBH (Baumgardt et al. 2005).

Both G1 and ω Centauri are somewhat atypical globular clusters. They are the most massive clusters of their parent galaxies, M31 and the Milky Way, respectively. Both clusters have been suggested to be the stripped nuclei of larger galaxies (e.g., Freeman 1993; Meylan et al. 2001). Indeed, the nuclei of galaxies have very similar structural properties to globular clusters (these nuclei are therefore also referred to as nuclear star clusters; e.g., Walcher et al. 2005). Accretion activity is known to exist in at least some of these nuclei (e.g., Seth et al. 2008). So if G1 and ω Centauri contain IMBHs, then these may simply be the low-mass tail of the distribution of supermassive black holes (SMBHs) known to exist in galactic nuclei. As such, they would not necessarily be telling us anything fundamentally new about the formation and evolution of globular clusters in general. Nonetheless, G1 and ω Centauri are probably the best current candidates for globular clusters with IMBHs. Validating the existence of their suggested IMBHs is therefore of fundamental importance.

The (initial) IMBH evidence for M15, G1, and ω Centauri was all based on studies of LOS velocities. Such studies are complicated because they require spectroscopy in regions of high stellar density. Integrated-light measurements work well for a cluster as distant as G1 (Gebhardt et al. 2002, 2005), but the downside is that the sphere of influence in such clusters is poorly resolved. In closer Milky Way clusters, the limited number of stars (especially brighter ones) introduces significant shot noise in integrated-light measurements. For ω Centauri, this required special treatment of the data to counteract this (NGB08). For M15, HST/STIS was used to obtain spectroscopy of individual stars (van der Marel et al. 2002). However, the signal-to-noise requirements for a velocity measurement limit the data set to the brightest stars. In turn, this sets the smallest radius inside which a velocity dispersion can be measured.

A significant improvement in data quality is possible with proper motion measurements. Such studies do not require much observing time and allow measurement even of the more numerous faint stars. Moreover, the ratio of the velocity dispersions in the tangential and radial proper motion directions directly constrains the velocity-dispersion anisotropy of the system. This is important because of the well-known mass-anisotropy degeneracy for stellar systems (Binney & Mamon 1982). To determine accurate proper motions close to the center requires an observing platform with high spatial resolution and long-term stability. The HST provides such a platform. McNamara et al. (2003) measured proper motions for M15, which were modeled by van den Bosch et al. (2006) and Chakrabarty (2006). This confirmed the presence of dark mass near the cluster center, but did not shed new light on the issue of whether this is due to mass segregation or an IMBH. McLaughlin et al. (2006) obtained and modeled proper motions for 47 Tuc, and used these to set an upper limit of 1000–1500 M on the mass of any IMBH in that cluster.

In Anderson & van der Marel (2010, hereafter Paper I), we presented the results of a new observational HST study of ω Centauri. For the central few arcmin of the cluster, we determined the projected number density distribution of some 106 stars, as well as accurate proper motions of some 105 stars. The exact position of the cluster center was determined in three independent ways. First, we determined the center of the number density distribution using various methods, taking care to accurately account for incompleteness. Second, we determined the kinematical center, being the symmetry point of the rms proper motion velocity field. And third, we determined the center of unresolved light seen in Two Micron All Sky Survey data. All determinations agree to within 1''–2'', yielding a final estimate with an uncertainty of ∼1''.

The NGB08 center position, and hence the position of their central spectroscopic field, is at R = 12'' from the center determined in Paper I (and coincidentally, the second off-center spectroscopic field they studied is also at R = 12''). The differences between the centers determined in these two studies are likely due to a combination of subtle effects that biased the NGB08 analysis, as discussed in Paper I. As a result of this, the density cusp found by NGB08 was not measured around the actual cluster center. Also, we showed in Paper I that their use of unresolved light (as opposed to star counts) increased the impact of shot noise. In Paper I, we determined the number density profile around the actual center of ω Centauri. It was found to be well matched by a standard King model, even though a shallow cusp may not be ruled out.

In Paper I, we also determined the rms proper motions of stars detected in the same 5'' × 5'' fields studied spectroscopically by NGB08. For the canonical distance D = 4.8 kpc (vdV06), the results translate to one-dimensional values ${\bar{\sigma }}_{\rm pm}= 18.9 \pm 1.3 \;{\rm km}\;{\rm s}^{-1}$ and ${\bar{\sigma }}_{\rm pm}= 19.0 \pm 1.6 \;{\rm km}\;{\rm s}^{-1}$, for their "central" and "off-center" fields, respectively. So we did not confirm the NGB08 finding that one of their fields has a higher dispersion. The observation that there is no significant difference in proper motion dispersion between the two fields is consistent with our finding that they are both at the same radius R = 12'' from the actual cluster center.

Our results of Paper I did not confirm the arguments put forward by NGB08 to suspect the presence of an IMBH in ω Centauri. However, this does not mean that such an IMBH may not be present after all. The size and quality of our proper motion data set are superb. This provides the opportunity to study the central dynamics of ω Centauri at a level of detail that is unmatched by other clusters. We therefore present here a detailed study of the dynamics of ω Centauri, with the primary goal of exploiting the new data from Paper I to constrain the mass of any possible IMBH. We also include existing ground-based data in our data–model comparisons, to obtain the best possible results.

We restrict our modeling efforts in the present paper to spherical models. This is a simplification, given that ω Centauri is elongated and rotating (in fact, unusually so for a globular cluster). However, both the elongation (Geyer et al. 1983) and the rotation rate (vdV06) decrease toward the cluster center. Therefore, the assumption of sphericity is likely to be reasonable in the central region most relevant for constraining an IMBH. We find support for this throughout our paper by showing that spherical models yield an anisotropy profile, distance, and mass-to-light ratio that are in excellent agreement with the values obtained by vdV06. More sophisticated models, such as those of the type described by vdV06, van den Bosch et al. (2006), and Chaname et al. (2008), have the potential to constrain other aspects of the structure of ω Centauri, such as its inclination, rotation properties, and the presence of any kinematic subcomponents. However, spherical models are sufficient to obtain robust constraints on the central mass distribution.

The structure of this paper is as follows. Section 2 discusses the general modeling approach. Section 3 presents the data on star count and surface brightness profiles used for the analysis. Section 4 discusses parameterized fits to these data. Section 5 presents the kinematical data used for the analysis. Section 6 presents the kinematical data–model comparison, with a determination of the best-fitting model parameters. Section 7 discusses how the new IMBH results for ω Centauri fit in with our general understanding of IMBH demographics in globular clusters. Section 8 presents and discusses the conclusions.

2. MODELING APPROACH

2.1. Mass Density

Imaging observations of a cluster yield an estimate of μobs,X(R), the radial profile of the surface brightness in a photometric band X, averaged over concentric circles of radius R around the cluster center. This quantity is related to the projected intensity Iobs(R) according to

Equation (1)

where M☉,X is the absolute magnitude of the Sun in the given photometric band. Throughout this paper, units in which quantities are expressed are given in square brackets where relevant. In the presence of AX magnitudes of foreground extinction, the extinction-corrected profile is given by

Equation (2)

For a spherical cluster, I(R) is the LOS integration over the three-dimensional luminosity density j(r). This integration can be inverted through an Abel transform (e.g., Binney & Tremaine 1987) to yield

Equation (3)

To express this quantity in units of L pc−3, one needs to specify the cluster distance D. Sizes in arcsec and pc are related according to

Equation (4)

The mass density of stellar objects and their remnants is

Equation (5)

where [M/L](r) is the average mass-to-light ratio at each radius. Gradients in [M/L](r) can be thought of as arising from two separate ingredients: (1) mass segregation among luminous stars and (2) mass segregation of the heavy compact stellar remnants (heavy white dwarfs, neutron stars, and stellar-mass black holes) with respect to the luminous stars. The first ingredient generally causes a shallow gradient in [M/L](r) over the scale of the entire cluster (e.g., Gill et al. 2008). This has little effect on the model predictions in the area surrounding the very center, where we wish to search for the signature of an IMBH. By contrast, the second ingredient can in some clusters yield a strong increase in [M/L](r) near the very center (e.g., Dull et al. 1997; Baumgardt et al. 2003a) that mimics the presence of a dark sub-cluster.

Our modeling approach allows for exploration of models with arbitrary [M/L](r). However, we have found it useful and sufficient in the present context to restrict attention to a parameterized subset. Based on the preceding considerations, we write the mass density as

Equation (6)

Here $\tilde{\rho }(r)$ is the mass density obtained under the assumption of a constant mass-to-light ratio ϒ. The excess ρdark is treated mathematically as a separate component. For ρdark, we have found it convenient to use the parameterization

Equation (7)

corresponding to a Plummer density distribution. Here Mdark is the total excess mass in dark objects and adark is the characteristic scale length of their spatial distribution. The quantities ϒ, Mdark, and adark provide a convenient way for exploring a range of plausible [M/L](r) functions using only three parameters.

By assuming that ϒ is independent of r, we neglect mass segregation among the luminous stars. For ω Centauri, there are in fact several pieces of evidence that it has not yet undergone complete mass segregation, consistent with its long half-mass relaxation time of ∼109.96±0.03 yr (McLaughlin & van der Marel 2005). First, Merritt et al. (1997) constructed non-parametric dynamical models for ω Centauri. They found that at the large radii sampled by their data, ρ(r) ∝ j(r) to within the error bars. Second, vdV06 constructed more general dynamical models for a wider range of data, and they too found that at all radii sampled by their study [M/L](r) is consistent with being constant. Third, Anderson (2002) studied luminosity functions (LFs) at different radii in ω Centauri using data obtained with HST. He found that variations in the LFs are less than would be expected if mass segregation had proceeded to the point of energy equipartition, and fourth, in Paper I we studied the proper motion dispersion of stars along the main sequence. This showed that lower mass stars do move faster, but by a smaller amount than would be expected for energy equipartition.

2.2. Gravitational Potential

The total gravitational potential of the system is

Equation (8)

where G is the gravitational constant and MBH is the mass of any central IMBH. The gravitational potential corresponding to $\tilde{\rho }(r)$ is

Equation (9)

The Plummer potential corresponding to ρdark is

Equation (10)

The potential Φtot could in principle be expanded to also include an extended halo of particulate dark matter, but that is more relevant for galaxies. There is no evidence that star clusters are embedded in dark halos. For the specific case of ω Centauri, vdV06 showed that there is no need for dark matter out to at least two half-light radii.

2.3. Intrinsic Velocity Moments

The Jeans equation for hydrostatic equilibrium in a spherical stellar system is (e.g., Binney & Tremaine 1987)

Equation (11)

where ν(r) is the number density of the stars under consideration. Because of the spherical symmetry, $\overline{v_{\theta }^2}= \overline{v_{\phi }^2}$. The function $\beta (r) \equiv 1 - \overline{v_{\rm tan}^2}/\overline{v_r^2}$ measures the anisotropy in the second velocity moments, where $\overline{v_{\rm tan}^2}\equiv \overline{v_{\theta }^2}= \overline{v_{\phi }^2}$. Models with β = 0 are isotropic, models with β < 0 are tangentially anisotropic, and models with 0 < β ⩽ 1 are radially anisotropic. The models may in principle be rotating if $\overline{v_{\phi }} \not= 0$. However, for the purposes of the present paper, we do not need to address how the second azimuthal velocity moment $\overline{v_{\phi }^2}\equiv \sigma _{\phi }^2 + \overline{v_{\phi }}^2$ splits into separate contributions from random motions and mean streaming.

The measured kinematics in a cluster are generally representative for only a subset of the stars (e.g., stars in that magnitude range for which kinematics can be measured). The number density of this subset is

Equation (12)

where λ(r) is the average luminosity of a star in the subset, and f(r) is the fraction of the total light that the subset contributes at each radius. Compact stellar remnants play no role in this equation, since they contribute neither to ν(r) (their kinematics are not probed) nor to j(r) (they emit negligible amounts of light). However, f(r) and λ(r) can vary slowly as a function of radius due to mass segregation among the luminous stars. The quantity λ(r) may also be affected by observational selection effects that vary as a function of radius. Nonetheless, variations in λ(r) will generally be smaller than those in f(r), as long as kinematics are measured for stars of similar brightness at each radius. Both f(r) and λ(r) are only expected to have a shallow gradient over the scale of the entire cluster. These gradients are much smaller than those in j(r), which can vary by orders of magnitude. Moreover, in the present paper, we are mostly concerned with the small area surrounding the very center. We therefore assume that

Equation (13)

which implies that, as before, we neglect mass segregation among the luminous stars.

It follows from the preceding that one can substitute j(r) for ν(r) in the Jeans equation (11), without actually needing to know the value of the constant in Equation (13). The Jeans equation is then a linear, first-order differential equation for $j \overline{v_r^2}$ with solution

Equation (14)

This is a special case of a more general solution for certain classes of axisymmetric systems (Bacon et al. 1983). We consider anisotropy functions of the form

Equation (15)

where ra > 0 is a characteristic anisotropy radius, β0 ⩽ 1 is the central value of β(r), and β ⩽ 1 is the asymptotic value at large radii. If β0 = β then β(r) is constant. With this β(r), Equation (14) reduces to

Equation (16)

which is always positive.

The integral in Equation (16) can be used with the equations given for j(r) and Φtot(r) in Sections 2.1 and 2.2 to allow numerical evaluation of $\overline{v_r^2}(r)$. This requires knowledge of the observed surface brightness profile μobs,X(R) and the known extinction AX. It also requires choices for the scalar model quantities D, ϒ, MBH, Mdark, adark, β0, β, and ra, which can all be chosen to optimize the fit to the available kinematical data.

2.4. Projected Velocity Moments

To calculate quantities that are projected along the LOS, consider an arbitrary point P in the cluster. We adopt a Cartesian coordinate system (pmr, los, pmt) with three orthogonal axis and its origin at the cluster center C. The pmr-axis (proper motion radial) lies in the plane of the sky and points from the projected position of C to the projected position of P. The los-axis (LOS) points from the observer to the point C. The pmt-axis (proper motion tangential) lies in the plane of the sky, and is orthogonal to the pmr- and los-axes in a right-handed sense. The cylindrical polar coordinate system associated with (pmr, los, pmt) is (R, ϕ, z), and the spherical polar coordinate system is (r, θ, ϕ), with θ ∈ [ − π/2, π/2], and θ = π/2 corresponding to the pmt axis. The velocities in the (pmr, los, pmt) system satisfy

Equation (17)

Since $\overline{v_r v_{\phi }} = 0$ in a spherical system, the second moments are

Equation (18)

where we have used that $\overline{v_{\phi }^2}= \overline{v_{\theta }^2}= \overline{v_r^2}[1 - \beta (r)]$.

The projected second moments satisfy

Equation (19)

The derivations of these equations use the facts that cos2ϕ = (R/r)2 and $d\;{\rm los} = d \sqrt{r^2-R^2} = r \: dr /\sqrt{r^2-R^2}$. The factor of 2 arises from the identical contributions from points with los <0 and los >0. The angle brackets around the squared velocities on the left-hand side indicate the combined effect of averaging over velocity space and density-weighted averaging along the LOS. Formally speaking, the weighting is with number density if one is modeling discrete measurements, and with luminosity density if one is modeling integrated-light measurements. However, in the present paper, we are assuming that ν(r) ∝ j(r), so the two cases are equivalent. The integrals in Equation (19) can be evaluated numerically once the Jeans equation has been solved.

2.5. Projected Velocity Distributions

The normalized projected velocity distributions ${\cal L}(v,R)$ at projected radius R are generally different for the los, pmr, and pmt directions. The Jeans equation suffices to calculate the second velocity moments of these distributions. To calculate the distributions themselves, it is necessary to know the full phase-space distribution function. This is not generally straightforward (or unique) for anisotropic models. However, for an isotropic model, the distribution function depends only on the energy, and it can be uniquely calculated for given j(r) (being proportional to both ν(r) and $\tilde{\rho }(r)$ under the present assumptions) and Φtot(r), with the help of Eddington's equation (Binney & Tremaine 1987). Because of the isotropy, the distribution ${\cal L}_{\rm iso}(v,R)$ is independent of the choice of the velocity direction (los, pmr, or pmt) and satisfies, in analogy with Equation (19),

Equation (20)

Here rmax(v) is defined to satisfy $\Psi (r_{\rm max}) \equiv {1\over 2}v^2$, with Ψ ≡ −Φtot. The quantity ${\cal L}_{\rm loc}(v,r)$ is the local velocity distribution (before projection along the LOS) at radius r, which can be written with the help of Eddington's equation as (van der Marel 1994a)

Equation (21)

The integrals in Equations (20) and (21) can be used with the equations given for j(r) and Φtot(r) in Sections 2.1 and 2.2 to allow numerical evaluation of ${\cal L}_{\rm iso}(v,R)$. The result is a symmetric function of v (i.e., ${\cal L}_{\rm iso}(v,R) = {\cal L}_{\rm iso}(-v,R)$). This remains true if the distribution function is given a non-zero odd part in the angular momentum Lz around some axis (corresponding to a rotating model), provided that ${\cal L}_{\rm iso}(v,R)$ is then interpreted as the distribution averaged over a circle of radius R.

2.6. Observed Quantities

Equations (19) and (20) define velocity moments and velocity distributions at a fixed position on the projected plane of the sky. However, to determine these quantities observationally, it is necessary to gather observations over a finite area. In integrated-light measurements this happens naturally through the use of, e.g., apertures, fibers, or pixels along a long slit of finite width. In discrete measurements of individual stars this happens by binning the data during analysis. In either case, the corresponding model predictions are obtained as a density-weighted integral over the corresponding area in the projected (x, y)-plane of the sky. For example, the second LOS velocity moment over an aperture is

Equation (22)

Similar equations apply to the other velocity moments and to the velocity distribution ${\cal L}_{\rm iso}(v,R)$. In integrated-light measurements, one can also take the observational point-spread function (PSF) into account by convolving both I (x, y) and Iv2los〉(x, y) with the PSF before evaluation of the integrals in Equation (22).

In the following, we write

Equation (23)

It is also useful to consider the quantity

Equation (24)

which corresponds to the one-dimensional rms velocity averaged over all directions in the plane of the sky. Note that velocities in the plane of the sky are not directly accessible observationally. Instead of ${\bar{\sigma }}_{\rm pmr}$ and ${\bar{\sigma }}_{\rm pmt}$, we measure proper motions ${\bar{\Sigma }}_{\rm pmr}$ and ${\bar{\Sigma }}_{\rm pmt}$ that satisfy

Equation (25)

Transformation to velocities requires an assumption about the cluster distance D.

The overlying bar in the preceding equations is used to differentiate the rms velocity ${\bar{\sigma }}$ from the velocity dispersion σ. These quantities are equal only if the mean streaming is zero (which is always true in the pmr direction).

2.7. Optimizing and Assessing the Fit to the Data

In general, we will want to model LOS data points ${\bar{\sigma }}_{{\rm los},i} \pm \Delta {\bar{\sigma }}_{{\rm los},i}$ available for i = 1, ..., N, as well as proper motion data points ${\bar{\Sigma }}_{{\rm pm\ldots },j} \pm \Delta {\bar{\Sigma }}_{{\rm pm\ldots },j}$ available for j = 1, ..., M (with each proper motion data point referring to either the pmr or the pmt direction). To calculate model predictions, we use the known surface brightness profile μobs,X(R) and the extinction AX. The questions are then how to assess the fit to the data for given model parameters, and how to determine the parameters that provide adequate fits.

2.7.1. Distance and Mass-to-Light Ratio

Suppose that we have obtained model predictions ${\hat{\sigma }}_{{\rm los},i}$ and ${\hat{\Sigma }}_{{\rm pm\ldots },j}$ for the data points, using trial values ${\hat{D}}$ and ${\hat{\Upsilon }}$ for the distance and mass-to-light ratio, respectively. The fit can then be improved by scaling of the model predictions, which corresponds to rescaling of the mass and distance scales of the model. Let Flos and Fpm be factors that scale the model predictions in the LOS and proper motion directions, respectively. The best-fitting values of these scale factors and their error bars can be calculated using the statistics

Equation (26)

The best fit in each case is the value with the minimum χ2 (i.e., dχ2/dF... = 0), and the 68.3% confidence region corresponds to Δχ2 = χ2 − χ2min ⩽ 1. The inferred scalings imply that the distance and mass-to-light ratio are optimized for $D_{\rm fit} = {\hat{D}} F_{\rm los} / F_{\rm pm}$ and $\Upsilon _{\rm fit} = {\hat{\Upsilon }} F_{\rm los} F_{\rm pm}$. These fits apply at the given values of the other model parameters, some of which are also affected by the scaling. The model parameters MBH and Mdark are proportional to F3los/Fpm, and the model parameter adark is proportional to Flos/Fpm.

The error bars in Dfit and ϒfit (at fixed values of the other parameters) follow from propagation of the uncertainties in Flos and Fpm. The errors in Flos and Fpm are uncorrelated, but the errors in Dfit and ${\hat{\Upsilon }}$ can be highly correlated or anti-correlated. We will not generally quote the correlation coefficient in this paper, but nonetheless, this should be kept in mind where necessary.

The use of scalings as described here provides a means of reducing the parameter space of the models by two. Also, it directly yields the best-fit distance and mass-to-light ratio if the remaining parameters can be determined or eliminated independently. The parameters MBH, Mdark, and adark can be determined as described in Section 2.7.2. They primarily affect the data points near the cluster center. So they can also be effectively eliminated by restricting the definition of the χ2 quantities in Equation (26) to the data points at large radius. The anisotropy profile β(r) can be determined as described in Section 2.7.3.

2.7.2. Dark Mass

Any dark component in our models can be either a cluster with parameters (Mdark, adark) or an IMBH of mass MBH. The latter is mathematically (but not astrophysically) equivalent to a cluster with parameters (Mdark = MBH, adark = 0). It is therefore not possible to discriminate between an IMBH and a dark cluster of sufficiently small size (within a resolution determined by the properties of the data). Conversely, we only need to explore the (Mdark, adark) parameter space, and this will automatically tell us (at adark = 0) what the limits on an IMBH are. In the following, we will use the term "dark mass" to refer generically to a dark component that could be either a dark cluster as in Equation (7) or an IMBH. We have not explored models that have both a dark cluster and an IMBH, although that would not have been difficult.

At given β(r), we start by calculating initial guesses for ${\hat{D}}$ and ${\hat{\Upsilon }}$ by fitting only the data at large radii. We then calculate for a range of trial values $({\hat{M}}_{\rm dark},{\hat{a}}_{\rm dark})$ the model predictions ${\hat{\sigma }}_{{\rm los},i}$ and ${\hat{\Sigma }}_{{\rm pm\ldots },j}$ for all the data points. For each pair of trial values, we calculate Dfit and ϒfit as in Section 2.7.1 by using the quantities χ2los and χ2pm, including observations at all distances from the cluster center. We then map the contours of χ2 ≡ χ2los + χ2pm in the two-dimensional (Mdark, adark) parameter space. The best fit (Mdark,fit, adark,fit) is obtained as the point of minimum χ2. The uncertainties of the fit follow from standard Δχ2 statistics, with the 68.3% confidence region corresponding to Δχ2 = χ2 − χ2min ⩽ 2.3 (Press et al. 1992). If there are combinations with Mdark = 0 within the confidence region, then a dark mass is not required to fit the data, for the assumed β(r). If there are combinations with Mdark ≠ 0 and adark = 0 within the confidence region, then the presence of an IMBH is consistent with the data for the assumed β(r). If we assume that the dark mass is an IMBH, then its 68.3% confidence region is given by Δχ2 ⩽ 1 at adark = 0.

The presence of an IMBH not only increases the rms velocity toward the center, but it also changes the shape of the velocity distribution. The distribution measured through an aperture around the center will have power-law wings that are significantly broader than those of a Gaussian distribution (van der Marel 1994b). This leads to the possibility of detecting very high velocity stars (Drukier & Bailyn 2003). It is therefore useful to determine and analyze the observed velocity distribution of the stars around the center to further assess the plausibility of the presence of an IMBH. The observed distribution can be compared to the predicted distribution ${\cal L}_{\rm iso, ap}(v)$ that can be calculated for an isotropic model. It is also useful to calculate the Gauss–Hermite moments (van der Marel & Franx 1993) of the observed velocity distribution as a function of distance from the center. A sufficiently massive IMBH causes an increase in h4 toward the center (van der Marel 1994b). The absence of broad wings or a central increase in h4 can be used to obtain an upper limit to the mass of any IMBH.

2.7.3. Anisotropy Profile

Binney & Mamon (1982) showed that there are many different combinations of β(r) and the enclosed mass M(<r) in a spherical system that can produce the same observed projected rms LOS velocity profile ${\bar{\sigma }}_{\rm los}(R)$. The key to constraining the mass distribution, and hence the possible presence of an IMBH, is therefore to constrain the anisotropy profile β(r). This is not possible when only rms LOS velocities are available. However, β(r) can be constrained when information is available on deviations of the LOS velocity distributions (LOSVDs) from a Gaussian, as a function of projected position on the sky (van der Marel & Franx 1993; Gerhard 1993). This information is fairly indirect, and to exploit it one must construct complicated numerical models that retrieve the full phase-space distribution function of the stars (van der Marel et al. 1998; Cretton et al. 1999; Gebhardt et al. 2000; Valluri et al. 2004; Thomas et al. 2004; van den Bosch et al. 2008). This has been explored with considerable success in the context of studies of early-type galaxies (e.g., Cappellari et al. 2007).

In the present context, high-quality proper motion data are available. It is then considerably more straightforward to constrain β(r). This was demonstrated by Leonard & Merritt (1989), who were the first to use proper motion data to examine in detail the mass distribution in a star cluster. The function $[{\bar{\Sigma }}_{\rm pmt}/{\bar{\Sigma }}_{\rm pmr}] (R)$ is a direct measure of anisotropy as seen in the plane of the sky. This ratio is independent of either the distance or the mass-to-light ratio scaling of the model. There is only a weak dependence on the presence of a dark mass (and this only affects the very central region). Therefore, the one-dimensional observed function $[{\bar{\Sigma }}_{\rm pmt}/{\bar{\Sigma }}_{\rm pmr}] (R)$ tightly constrains the one-dimensional intrinsic function β(r). In fact, for a spherical system, there is a unique relation between the two (Leonard & Merritt 1989). This is the primary reason why in the present context it is possible to obtain robust results based on something as simple as the spherical Jeans equation, while much more sophisticated techniques have become standard practice for galaxies.

To build some intuitive understanding of the meaning of $[{\bar{\Sigma }}_{\rm pmt}/{\bar{\Sigma }}_{\rm pmr}] (R)$, consider the case in which β(r) is independent of radius and $j \overline{v_r^2}(r) \propto r^{-\xi }$. The relevant integrals in Equation (19) can then be expressed analytically in terms of beta functions. Their ratio is independent of radius R and simplifies to

Equation (27)

For small β (i.e., a system not too far from isotropy), we can use a first-order Taylor approximation, which yields

Equation (28)

where we have defined the function

Equation (29)

in analogy with the definition of β(r). This implies that the deviation of $[{\bar{\Sigma }}_{\rm pmt}/{\bar{\Sigma }}_{\rm pmr}]$ from unity is a factor $(1 - {1\over \xi })$ of the deviation of $[\overline{v_{\rm tan}^2}/\overline{v_r^2}]^{1/2}$ from unity. (This uses the Taylor approximation (1 − β)1/2 ≈ 1 − (β/2) and similarly for βpm.) Hence, $[{\bar{\Sigma }}_{\rm pmt}/{\bar{\Sigma }}_{\rm pmr}(R)]$ is a "diluted" measure of the anisotropy that is present intrinsically, with the dilution factor $(1 - {1\over \xi })$.

The approximations made above cannot be used in the core of a cluster, where ξ = 0. The assumption that the system remains scale free all along the LOS is then not accurate. However, in the outer parts of a cluster the approximation is a reasonably accurate guide, with the value of ξ depending on the exact cluster structure. For example, an isothermal sphere has ξ = 2 at large radii, while a Plummer model has ξ = 6. The corresponding dilution factors $(1 - {1\over \xi })$ are 0.50 and 0.83, respectively. Real clusters typically fall between these extremes. More extensive analytical calculations of βpm(R) are presented in, e.g., Leonard & Merritt (1989) and Wybo & Dejonghe (1995). Either way, analytical calculations are mostly useful here for illustration. In practice, we calculate the integrals in Equation (19) numerically.

The extent to which the velocity distribution of the stars (in either the los, pmr, or pmt direction) differs from a Gaussian provides secondary information on the amount of velocity anisotropy in the system. The modeling technique used here is insufficient to exploit this information, since we are not calculating actual phase-space distribution functions for anisotropic models (although this would in principle be possible, e.g., Dejonghe 1987; Gerhard 1993; van der Marel et al. 2000). Nonetheless, we can observationally characterize the Gauss–Hermite moments as a function of projected distance from the cluster center. Also, we can calculate the predicted velocity distributions ${\cal L}_{\rm iso}(v,R)$ for an isotropic model, and from this the corresponding Gauss–Hermite moments. Comparison of the observed and predicted quantities provides an independent assessment of possible deviations from isotropy.

2.8. Numerical Implementation

To apply the formalism described above, we augmented a software implementation used previously in, e.g., van der Marel (1994a). The code uses a logarithmically spaced grid, from very small to very large radii (so that a negligible fraction of the cluster mass resides outside the grid). Relevant quantities are tabulated on this grid. Integrals are evaluated through Romberg quadrature (e.g., Press et al. 1992) to high numerical precision. Tabulated quantities are interpolated using cubic splines for use in subsequent integrations. This avoids the need for calculation of higher-dimensional integrals. The derivatives dI/dR and dj/dΨ in Equations (3) and (21), respectively, are calculated using finite differences. The derivative dΦtot/dr in Equation (16) can be manipulated analytically so that numerical differentiation is not needed. The accuracy of the code was verified by reproducing analytically known results for special potential-density pairs (e.g., Plummer models).

3. PROJECTED DENSITY PROFILE DATA

The primary input for our models is the projected radial density profile. To characterize this profile over the full range of radii spanned by ω Centauri, we have combined the ground-based measurements at intermediate and large radii collected in Trager et al. (1995), with the new HST measurements at small radii presented in Paper I. Much of the measurements come in the form of projected star count profiles νproj(R). Our equations of Section 2 are mostly expressed in terms of the observed (or implied) surface brightness profile μobs(R). The profiles νproj(R) and μobs(R) are uniquely related to each other (see Equation (30) below), given our assumption to neglect mass segregation among the luminous stars.

3.1. Ground-based Data

Trager et al. compiled literature data from various sources. We rejected from their compilation the data points to which they did not assign the maximum weight (unity in their notation). This leaves only the most reliable literature sources for ω Centauri, namely the integrated-light photo-electric measurements of Gascoigne & Burr (1956) and Da Costa (1979), and the number density measurements of King et al. (1968). All these authors presented averages for concentric circular annuli. Da Costa also presented separate measurements using a drift scan technique, which were subsequently calibrated to estimate the profile along concentric circular annuli. King et al. presented results for two separate plates from Boyden Observatory. We follow Trager et al. by referring to the different data sets with the abbreviations GBANN, DANN, DSCAN, ADH-1792, and ADH-1846; the latter two abbreviations refer to the IDs of the photographic plates analyzed by King et al. For our analysis, we used the values provided by Trager et al., who calibrated the published measurements to a V-band surface brightness scale with a common zero point. We assigned to each data point an error bar of 0.142 mag, based on the procedure and analysis of McLaughlin & van der Marel (2005, see their Table 6). This corresponds to the scatter in the data with respect to a smooth curve.

3.2. HST Data

Paper I presented number density measurements for concentric circular annuli based on HST data. Considerable effort was spent to accurately determine the cluster center, and to correct the results for the effects of incompleteness. Error bars were determined from the Poisson statistics of the counted stars. Paper I discussed number density profiles for various magnitude bins and showed that there is no strong dependence on magnitude. For the analysis presented here we have used the profile for all stars in the HST images with B-band instrumental magnitude <−10 (this is defined in Paper I and corresponds to stars measured with signal-to-noise ratio (S/N) ≳100). The choice of magnitude range was motivated by the desire to maximize the number of stars (i.e., to minimize the error bars on the density profile) while maintaining a similar mean magnitude (∼−12) as for the sample of stars for which high-quality proper motions are available from Paper I.

To extend the profile from the Trager et al. compilation to the smaller radii accessible with the HST data, it is necessary to calibrate the projected number density νproj(R) to a projected surface brightness μobs,V(R). It follows from Equations (1), (4), (12), and (13) that

Equation (30)

where the zero point Z is equal to

Equation (31)

where M☉,V = 4.83 (Binney & Merrifield 1998). Although (f/λ[L]) can in principle be estimated, we have found it more reliable to calibrate Z so that the surface brightness profiles from Trager et al. and Paper I agree in their region of overlap. We have done this by including Z as a free parameter in the fits described below, which yields Z = 18.21. For a canonical distance D = 4.8 kpc (vdV06) this implies that (f/λ[L]) = 1.0. With this Z there is excellent agreement between the different data sets, as shown in Figure 1 (top panel).

Figure 1.

Figure 1. Top panel: V-band surface brightness profile of ω Centauri in mag arcsec−2 vs. radius in arcsec. Data points are from various sources as described in the text. The curve is our best-fit model of Equation (32), which has a central cusp slope γ = 0.05. Bottom panel: residuals of the fit in the top panel. The small solid dots (black) are the HST star count data from Paper I. The bigger open symbols are from the compilation of ground-based data in Trager et al. (1995), as described in the text: GBANN, hexagons (green); DANN, circles (red); DSCAN, squares (magenta); ADH-1792, triangles (cyan); ADH-1846, crosses (blue).

Standard image High-resolution image

NGB08 studied the surface brightness profile of unresolved light in ω Centauri using HST images. In principle, the advantage of using unresolved light over star counts is that one can trace the cumulative effect of all the unresolved faint stars, rather than counting only the less numerous resolved ones. However, we showed in Paper I that even at HST resolution, the unresolved light is dominated by scattered light from the PSF wings of only the brightest stars. So in reality, the inherent statistics of unresolved light measurements are poorer than those of star count measurements. Moreover, we showed in Paper I that the combined photometric and kinematic center of ω Centauri is actually 12'' away from the position where NGB08 identified it to be. There is no simple way to transform their measurements over concentric annuli to a different center position. For these reasons, we have not included the NGB08 surface brightness data points in the data set used for the dynamical modeling.

4. PROJECTED DENSITY PROFILE MODELS

It is important for the modeling to have a smooth representation of the projected intensity profile I(R), since the derivative dI/dR is needed in Equation (3) for the calculation of j(r). For the dynamical modeling, we have used a parametric approach, although this can in principle be done non-parametrically (e.g., Gebhardt & Fischer 1995). We have fit the data using a smooth function of the form

Equation (32)

This functional form has no particular physical significance, but it is useful because it can fit a wide range of observed profiles. The profile has a power-law cusp IR−γ at radii Rb. It then breaks to a logarithmic slope η at Rb, and further to a logarithmic slope epsilon at Rc > b. The parameters α and δ measure the sharpnesses of the breaks. The intensity Ib, corresponding to a magnitude μb through Equation (1), sets the overall intensity scale. The profile is similar to a so-called "nuker" profile (Lauer et al. 1995), but with two breaks instead of just one. We will refer to it as a generalized nuker profile.

The best fit to the data set presented in Section 3 was determined using a Levenberg–Marquardt iteration scheme (Press et al. 1992). It has parameters μb = 17.74, b = 183farcs1, c = 2158'', γ = 0.05, η = 2.37, epsilon = 10.00 (the maximum value allowed during the iteration), α = 2.29, and δ = 1.70. The fit and its residuals are shown in Figure 1. The quality of the fit is excellent. For the subset of the data from Paper I, the χ2 = 130.5 for 125 data points with an rms residual of 0.04 mag. The total χ2min = 169.3 for all 159 data points, corresponding to 151 degrees of freedom. This is acceptable at the ∼1σ level for a χ2 distribution. The residuals appear generally consistent with random scatter in line with the error bars, with little evidence for systematic trends.

The best fit has a central cusp slope of γ = 0.05. To determine the statistical error on this value, we also performed a range of fits in which γ was fixed a priori, while all other parameters were varied to optimize the fit. The central structure of our best fit for γ = 0 is shown in Figure 2 for comparison to our overall best fit. The curve of Δχ2 ≡ χ2 − χ2min as a function of γ was used to derive the error bar on γ, which yields γ = 0.05 ± 0.02.

Figure 2.

Figure 2. V-band surface brightness profile of ω Centauri in mag arcsec−2 vs. radius in arcsec. The data and the symbols are the same as in Figure 1(top panel), but this figure focuses more on the central region. The solid curve (black in the online version) is the overall best model fit of Equation (32), which has a central cusp slope γ = 0.05. The long-dashed curve (blue) is the best fit when the cusp slope is fixed a priori to γ = 0. The dashed (cyan) and dotted (magenta) curves are the Wilson and King models, respectively, presented in McLaughlin & van der Marel (2005). These were tailored to fit the compilation of ground-based data in Trager et al. (1995).

Standard image High-resolution image

When using a generalized nuker fit, a constant surface brightness core is inconsistent with the data at ∼2.5σ confidence. However, this parameterized approach may not yield an unbiased estimate of the asymptotic slope γ at the smallest radii. The data at all radii are given equal weight in the χ2 of the fit, but it is obviously only the data near the very center that contain most of the relevant information on γ. Also, the two breaks at radii b and c in the generalized nuker profile fit correspond to the breaks associated with the traditional core and tidal radii of ω Centauri. The parameterization does not have sufficient flexibility to reproduce any possible change in profile slope well within the core radius (r0 = 141+7−13 arcsec; McLaughlin & van der Marel 2005).

An alternative and simpler approach to determine γ is to fit a straight line to the (log R, log I) data at RRmax. For Rmax = 15'', this yields γ = 0.00 ± 0.07. The values of γ inferred for other values of Rmax in the range 0''–20'' are in statistical agreement with this. This analysis indicates that the central data points are consistent with a flat core.

An additional assessment of models with a flat core can be obtained from comparison of the new data with previously published models. McLaughlin & van der Marel (2005) presented detailed surface brightness fits for a large sample of galactic and extragalactic globular clusters. They considered three types of models with decreasing amounts of diffuse outer structure, the power-law model, the Wilson (1975) model, and the King (1966) model. The power-law models are similar to the parameterizations that we have used here, but with a more restricted subset of the parameters (namely, γ = 0, α = 2, and c = were kept fixed). The models from Wilson and King are both based on a parameterized isotropic phase-space distribution function. The Wilson prescription differs from the better known King models in that it uses a slightly different lowering of the exponential energy distribution, so as to produce a model that has a more extended but still finite outer structure. McLaughlin and van der Marel discussed ω Centauri as a special example (their Section 4.2.1) and found that it is well fit by a Wilson model, but less so by a King model. The central structure of both models is shown in Figure 2. The Wilson model is very similar to our best-fit model with fixed γ = 0. The King model is also similar at radii R ≲ 100'', but it underpredicts the brightness at intermediate radii (before improving again at larger radii, not shown in Figure 2).

The Wilson and King models shown in Figure 2 are the ones that were previously fit by McLaughlin and van der Marel to the Trager et al. data. At small radii, these models represent an extrapolation of the ground-based data under the assumption of a constant surface brightness core. Although the models were not fit to the HST data, they do provide a perfect fit to the two innermost HST star count data points. By contrast, our own best generalized nuker model fit with γ = 0.05 overpredicts both of the innermost data points by about 1σ. The fact that γ = 0.05 is preferred in our fits over γ = 0 is therefore because it fits better the data at radii R ≈ 10'', and not because it provides the best fit closest to the center.

NGB08 inferred from their HST study of integrated light in ω Centauri that it has a central cusp slope γ = 0.08 ± 0.03. At first glance, this result appears consistent with that from our generalized nuker fit. However, since they did not measure the brightness profile around the same cluster center, the agreement between our inferred cusp slopes is likely coincidental.

In summary, we conclude that the central cusp slope of ω Centauri is γ ≲ 0.07 at ∼1σ confidence. That is, we interpret the γ = 0.05 ± 0.02 from our generalized nuker fit as an upper limit. We do this because it overpredicts the central two data points, and because other statistics do not show evidence for a non-zero slope. In the dynamical analysis that follows, we specifically consider two generalized nuker fits that span the range of possibilities: models with a shallow cusp (γ = 0.05) and models with a constant density core (γ = 0). We will refer to these as cusp models and core models, respectively. In all calculations, we have transformed the observed intensity Iobs,V(R) to an extinction-corrected intensity IV(R) using Equation (2) with AV = 0.372 for ω Centauri (McLaughlin & van der Marel 2005).

5. KINEMATICAL DATA

To constrain the parameters of the dynamical models, we need to use both LOS and proper motion data at a range of radii. For this, we have combined ground-based LOS and proper motion data at intermediate and large radii with the new HST proper motion data at small radii presented in Paper I.

We construct spherical models to interpret the data. These models yield only a single radial profile for each projected quantity, with no azimuthal variations from the observed major axis of ω Centauri to its observed minor axis. To get accurate results for the model parameters, it is important that these models be used to reproduce the average radial variations seen in the data. This is achieved by using for each projected quantity that is fit the average of the measurements along concentric annuli. This is indeed what we extracted for the projected density in Section 3, and we therefore do the same here for the projected velocity distributions. When averages are taken over annuli, all the odd moments of the velocity distribution are zero. Of primary interest here are the second moments, which yield the rms velocity ${\bar{\sigma }}_{\rm los}$ and the rms proper motions ${\bar{\Sigma }}_{\rm pmr}$ and ${\bar{\Sigma }}_{\rm pmt}$.

In reality, variations do exist between observed quantities along the major and minor axes. vdV06 discussed this in detail for the existing ground-based data. In Appendix B, we address this for the new HST proper motion data. In particular, the rms proper motions are not the same along the major and minor axes, and the RMS proper motions in the major and minor axes directions on the sky are not the same. Also, the mean proper motion in the transverse (pmt) direction is not zero (vdV06; Section 5.2.3; Appendix C). By modeling these issues, it is possible to constrain, e.g., the cluster inclination and its rotation rate. However, it is not essential to model these issues to achieve a good understanding of the central mass distribution, which is the main goal here.

5.1. Ground-based Data

5.1.1. Line-of-sight Kinematics

vdV06 compiled LOS velocity data of individual stars in ω Centauri. The original data were obtained by Suntzeff & Kraft (1996), Mayor et al. (1997), Reijns et al. (2005), and K. Gebhardt et al. (2010, in preparation). vdV06 performed various operations and cuts to homogenize and correct the data as necessary for dynamical modeling. This included: improved removal of cluster non-members; removal of stars with low enough brightness or large enough velocity error bars to make their reliability suspect; application of additive velocity shifts to correct for offsets in velocity zero points between data sets; multiplicative rescaling of the velocity error bars where necessary, as dictated by actual measurements of scatter between repeat measurements; subtraction of the systemic LOS velocity; and subtraction of the apparent solid-body rotation introduced into the velocity field through the combination of systemic transverse motion and large angular extent ("perspective rotation"). This produced a final homogenized sample of LOS velocities for 2163 stars with uncertainties below 2.0 km s−1.

Glenn van de Ven kindly made the homogenized sample from vdV06 available to us for use in our data–model comparisons. To determine the profile of ${\bar{\sigma }}_{\rm los}(R)$, we binned the stars in radius and then used the statistical methodology described in Appendix A. Radii were calculated with respect to the center determined in Paper I, which differs by 15farcs5 from the center used by vdV06 (which was determined by van Leeuwen et al. 2000). We chose a central bin of 20'' radius (with 25 stars) for maximum resolution near the center, and then chose the remaining bins to contain 200 stars each.

Sollima et al. (2009) obtained LOS velocities of 318 cluster members at large radii, which they merged with observations closer to the center by Pancino et al. (2007). This yielded a homogeneous sample of 946 cluster members, from which they derived the profile ${\bar{\sigma }}_{\rm los}(R)$. Seitzer (1983) presented the ${\bar{\sigma }}_{\rm los}(R)$ profile derived from 118 cluster members with known velocities at a range of radii. We included the published profiles from both studies in our data–model comparisons. For each study, we averaged the first two bins in the profile together, since the first bin had a large error bar and did not provide spatial resolution that competes with the vdV06 compilation.

Scarpa et al. (2003) obtained LOS velocities of 75 cluster members at large radii. From this, they determined the velocity dispersion σlos in four radial bins. The stars were limited to the west side of the cluster, so the results do not represent averages over a complete annulus on the sky. Nonetheless, we included these observations in our data–model comparisons. We subtracted in quadrature from the published σlos values the average measurement error of 1 km s−1. To obtain ${\bar{\sigma }}_{\rm los}$, we added in quadrature V2/2, where V is the rotation velocity amplitude quoted in Scarpa et al. (2003); the factor 1/2 is the average of a squared sinusoidal variation over a full range of angles.

The profiles obtained from the Sollima et al. (2009), Seitzer (1983), and Scarpa et al. (2003) observations do not pertain to the same center as that inferred in Paper I. However, the smallest radial bin that we use from these studies is the 4 arcmin diameter combined central bin of Sollima et al. A center offset of 15farcs5 (as applicable to the Sollima et al. and Scarpa et al. results, who used the van Leeuwen et al. 2000 center) should not significantly affect statistics calculated over bins that extend to such large distances from the center.

At small radii we used as additional data points the integrated-light measurements from NGB08. They inferred velocity dispersions over square fields of 5'' × 5''. Because of the significant size of these fields, we did not include the impact of PSF convolution in our model predictions for these data. We placed the measurements at the distance from the cluster center inferred in Paper I, which is 12'' for both fields. We cannot use the data to calculate a true average over an annulus. However, we do have two measurements at the same distance from the center at a 60° angle. Including both measurements in our model fits is in low-order approximation similar to fitting the average over an annulus.

NGB08 did not measure the mean velocity over their fields as compared to the systemic velocity. Hence, their measurements are dispersions σlos rather than RMS velocities ${\bar{\sigma }}_{\rm los}$. However, it is known from previous observations (e.g., vdV06) that Vloslos in ω Centauri decreases toward the center. This ratio is small enough at the position of the NGB08 fields that ${\bar{\sigma }}_{\rm los}\approx \sigma _{\rm los}$ to high accuracy (see Section 5.2.3 for a quantitative estimate).

Figure 3 (top panel) shows the collected ${\bar{\sigma }}_{\rm los}$ data from all the sources, as a function of the projected distance R from the cluster center. The different data sets match well in their regions of overlap.

Figure 3.

Figure 3. Observed kinematical quantities for ω Centauri as a function of the projected radius R from the cluster center. Data points are based on the analysis of various sources described in the text: circles (magenta) are based on vdV06, which is itself a compilation of LOS data from Suntzeff & Kraft (1996), Mayor et al. (1997), Reijns et al. (2005), and K. Gebhardt et al. (2010, in preparation), and proper motion data from van Leeuwen et al. (2000); triangles (cyan) are based on NGB08; stars (blue) are based on Sollima et al. (2009); squares (green) are based on Seitzer (1983); crosses (red) are based on Scarpa et al. (2003); and small dots (black) are based on the HST data from Paper I. Some error bars are of similar size as the plot symbols. Data from all ground-based sources except NGB08 were multiplied by a factor 1.023 to correct to same the characteristic main-sequence stellar mass as for the HST sample (see Section 5.1.3). Top panel: RMS LOS velocity ${\bar{\sigma }}_{\rm los}$ in km s−1 (data indicated with open symbols). Middle panel: RMS proper motion ${\bar{\Sigma }}_{\rm pm}$ in mas yr−1, averaged over all directions in the plane of the sky, as defined by Equation (24) (data indicated with closed symbols). Bottom panel: ratio ${\bar{\Sigma }}_{\rm pmt}/{\bar{\Sigma }}_{\rm pmr}$ of the RMS motions in the tangential and radial proper motion directions (data indicated with closed symbols). The RMS quantities in each data set are averages over adjacent circular annuli. The extent of the annulus for data point i can be assessed by taking half the distance between adjacent data points, ΔRi ≈ (Ri+1Ri−1)/2.

Standard image High-resolution image

5.1.2. Proper Motion Kinematics

Ground-based proper motion data are available from van Leeuwen et al. (2000). vdV06 performed various operations and cuts to optimize and correct the data as necessary for dynamical modeling. This included: improved removal of cluster non-members; removal of stars with low enough brightness, large enough velocity error bars, or neighbors that are nearby enough to make their reliability suspect; and subtraction of the apparent solid-body rotation introduced by perspective rotation, and other spurious solid body rotation components. This produced a final sample of proper motions for 2295 stars with uncertainties below 0.2 mas yr−1 (i.e., 4.6 km s−1 at the canonical distance D = 4.8 kpc).

Glenn van de Ven kindly made the homogenized sample from vdV06 available to us for use in our data–model comparisons. Proper motions were given along the major and minor axes directions. For each star, we transformed this into proper motions in the radial (pmr) and tangential (pmt) directions. We then determined the profiles of ${\bar{\Sigma }}_{\rm pmr}(R)$ and ${\bar{\Sigma }}_{\rm pmt}(R)$ by binning the stars in radius and using the statistical methodology described in Appendix A. Radii and proper motion coordinate transformations were calculated with respect to the center determined in Paper I. We chose a central bin of 90'' radius (with 30 stars) for maximum resolution near the center, and then chose the remaining bins to contain 200 stars each. We followed vdV06 by excluding the 65 outermost stars with radii R ∼ 20–30 arcmin.

For fitting models to the data, we always use the quantities ${\bar{\Sigma }}_{\rm pmr}$ and ${\bar{\Sigma }}_{\rm pmt}$. However, for visualizing the results, we have found it more convenient to use the quantities ${{\bar{\Sigma }}_{\rm pm}}$ and ${\bar{\Sigma }}_{\rm pmt}/{\bar{\Sigma }}_{\rm pmr}$. The RMS proper motion ${{\bar{\Sigma }}_{\rm pm}}$ averaged over all directions in the plane of the sky can be calculated directly through its definition in Equations (24) and (25). In practice, we found it more convenient to calculate it by applying the statistical methodology described in Appendix A on the combined array of the pmr and pmt data points. For visualization of the ratio ${\bar{\Sigma }}_{\rm pmt}/{\bar{\Sigma }}_{\rm pmr}$, we adopted a radial binning scheme with 400 stars in each bin to obtain smaller and more useful error bars.

Figure 3 (middle panel) shows the ${\bar{\Sigma }}_{\rm pm}$ data as a function of the projected distance R from the cluster center, and Figure 3 (bottom panel) shows the ${\bar{\Sigma }}_{\rm pmt}/{\bar{\Sigma }}_{\rm pmr}$ data. The velocity distribution is close to isotropic near the center, but with some radial anisotropy. There is a significant increase in tangential anisotropy at a few core radii. The latter is mostly due to the contribution of rotation to the second moment ${\bar{\Sigma }}_{\rm pmt}$. The anisotropy in the proper motion dispersions is much less (King & Anderson 2002).

5.1.3. Dependence of Kinematics on Stellar Mass

Luminous stars in a cluster undergo two distinct effects as a result of two-body relaxation: mass segregation and energy equipartition. These effects are intimately connected through basic dynamical theory, but from an observational perspective it is useful to think of them as separate effects. Mass segregation causes stars of different masses to have slightly different spatial distributions. In Section 2, we assumed that this could be neglected when parameterizing the run of the mass-to-light ratio with radius, and when connecting star count data at small radii to surface brightness data at large radii. This assumption was motivated by the absence of high-quality data to constrain the amount of mass segregation over the full scale of the cluster, and by the desire to minimize the number of a priori model assumptions. Energy equipartition causes stars of low mass to move faster than those of high mass. We measured in Paper I that low-mass stars in ω Centauri do indeed move faster than high-mass stars, albeit by less than would be predicted for complete energy equipartition (vm−0.5). Since we have a direct measurement of this, there is no reason to ignore this effect in the modeling. We therefore correct for it when comparing the observed kinematics from different data sets.

The available discrete ground-based LOS and proper motion data all pertain to relatively bright stars. The apparent magnitudes of these stars place them on the giant or sub-giant branch. Since stellar evolution proceeds relatively rapidly after the main-sequence turn-off, these stars all have essentially the same mass as the main-sequence turn-off. By contrast, the stars in our new HST proper motion sample from Paper I (discussed in detail in Section 5.2 below) extend from about the main-sequence turn-off to ∼3 mag below that. From the analysis of Paper I, it follows that the RMS proper motion of stars at the average magnitude of the HST sample is ∼2.3% higher than that at the main-sequence turn-off.

For a fair comparison to the new HST proper motion data, we took the kinematics from the discrete velocity studies in Sections 5.1.1 and 5.1.2 and multiplied them by a factor 1.023. This implies that we use the observed (sub-)giant star kinematics to estimate the kinematics of lower mass stars in the same general population. We assume implicitly that the variation of RMS kinematics with mass is the same everywhere in the cluster (although it was only measured near the cluster center), and that it applies also in the LOS direction (although it was only measured in the proper motion direction). In reality, one would expect the relation between mass and velocity dispersion to change with radius, since relaxation proceeds faster near the cluster center than further out. This was neglected in the present context by applying the same small correction throughout the cluster. This was motivated in part by the fact that our study focuses mostly on the central part of the cluster anyway, but also by the fact that we did not actually detect a difference in equipartition in Paper I between the different (central and major axis) fields for which we obtained proper motions. We did not apply the multiplicative correction to the NGB08 data. Their measurements are based on integrated light, excluding the brightest regions in their fields. It is therefore not obvious that their measurements would be dominated by (sub-)giant stars. Either way, a 2.3% correction would be much less than the error bars on their data points.

In summary, we are using our Jeans models to fit the kinematics of stars with masses typical of our HST proper motion sample. We extend our actual HST kinematics to larger radii using ground-based kinematics that have been corrected to the same characteristic mass. We model the cluster density profile using HST star counts for a sample that is also centered on the same characteristic mass (see Section 3.2). We make the simplifying assumption that there is no mass segregation among the luminous stars only so that we can extend the HST density profile to larger radii, where it can be tied to ground-based star count data and integrated photometry that are based largely on giant stars (see Section 3.1). Since this approximation does not affect the very central region of the cluster, it should have little effect on our results for the central mass distribution. In fact, the multiplicative correction of 1.023 discussed above also has little effect on this. We verified explicitly that none of the main conclusions of our paper change at a level that exceeds the formal uncertainties when the multiplicative correction is omitted.

5.2. HST Proper Motion Data

5.2.1. Sample Selection

In Paper I, we derived proper motions for stars observed in two fields observed by HST, a "central field" on the cluster center, and a "major axis field" positioned adjacent to it in a non-overlapping fashion roughly along the cluster major axis. Here we use the "high-quality" sample presented in Paper I, which contains the non-saturated stars that are isolated enough and bright enough (B-band instrumental magnitude <−11) for a particularly reliable proper motion measurement. This sample has 53,382 stars in the central field and 19,593 stars in the major axis field. In this sample, 95.1% of the stars have proper motion errors below 0.2 mas yr−1. Our cut in proper motion accuracy is therefore not very different from that applied by vdV06. However, the median proper motion error per coordinate in the final sample of vdV06 is ∼0.14 mas yr−1, whereas it is a factor of 2 smaller, 0.07 mas yr−1, for our sample. We note that our use of only the high-quality sample is rather conservative. Even the remainder of the sample of Paper I is quite accurate, and could well have been used to constrain the cluster dynamics.

The central HST field covers radii R ⩽ 147farcs4, while the major axis field covers radii 99farcs7 ⩽ R ⩽ 347farcs1. The central field has complete position angle coverage out to R = 71farcs7 (the radius of the largest enclosed circle). For 71farcs7 ⩽ R ⩽ 147farcs4, the coverage over the range of position angles becomes increasingly sparse. The major axis field is restricted at all radii to position angles within 48fdg5 of the major axis. For the analysis of the present paper, it is important to have coverage of all position angles at a given radius, so that average kinematical quantities over circular annuli can be calculated. For this reason, we have used only the 25,194 stars at R ⩽ 71farcs7 to constrain the dynamical models.

In our analysis, we ignore the 47,781 stars with high-quality HST proper motions at radii 71farcs7 ⩽ R ⩽ 347farcs1. We do discuss relevant quantities derived from these data in Appendix C. More sophisticated axisymmetric modeling techniques (such as those in vdV06; van den Bosch et al. 2006; and Chaname et al. 2008) that might be applied in the future probably would not want to exclude these data in their analysis. By contrast, to include these data in our spherical models would require estimates of the ratios ${\cal R}$ of the average of either ${\bar{\Sigma }}_{\rm pmr}$ or ${\bar{\Sigma }}_{\rm pmt}$ over a restricted range of position angles, to its average over a circle. We have experimented with such estimates, but concluded that the approximations that need to be made introduce sufficient uncertainties that these data do not really help to constrain our models. Conversely, we have found no evidence that inclusion of the HST data at 71farcs7 ⩽ R ⩽ 347farcs1, combined with reasonable estimates for the ratios ${\cal R}$, would alter in any way the conclusions draw below (see Appendix C).

To reject cluster non-members, we created scatter plots of total two-dimensional proper motion $|{\vec{p}}|$ versus radius R, where ${\vec{p}}$ is the proper motion vector of an individual star. Such a plot reveals a sparse population of field stars at large values of $|{\vec{p}}|$, in addition to the numerous cluster stars at low $|{\vec{p}}|$. We rejected those stars from the sample that reside at R > 10'' and for which $|{\vec{p}}| > 2.8 - (r/500^{\prime \prime }) \;{\rm mas}\;{\rm yr}^{-1}$. At the radii for which we have HST proper motions, this equation is a good approximation to the escape velocity curve for a spherical model with a canonical distance and mass-to-light ratio (D = 4.8 kpc, ϒV = 2.5; vdV06). With this criterion, only 27 non-member stars were identified at R ⩽ 71farcs7, the closest of which resides at R = 21farcs4. Since the presence of an IMBH may result in the presence of rapidly moving stars at small radii, we did not reject any stars at R ⩽ 10''. The wings of the observed velocity distribution at small radii are discussed in detail in Section 6.7 below.

Our final HST proper motion sample for the dynamical modeling consists of 25,167 stars. This is a factor ∼11 larger than the number of stars with proper motions in the vdV06 sample. Most of their stars reside at larger radii than those sampled here. At the radii R ⩽ 71farcs7 of our sample, vdV06 had only some two dozen stars with proper motions, and ∼270 stars with LOS velocities. The HST data therefore provide a major advance for constraining dynamical models, especially as it pertains to the gravitational influence of a possible IMBH at small radii.

5.2.2. rms Proper Motions

For each star in the HST sample, we decomposed the observed proper motion vector ${\vec{p}}$ into components in the pmr and pmt directions, respectively. We then binned the stars in radius and derived the profiles ${\bar{\Sigma }}_{\rm pmr}(R)$, ${\bar{\Sigma }}_{\rm pmt}(R)$, and ${\bar{\Sigma }}_{\rm pm}(R)$. We did this in a similar fashion as for the ground-based proper motion data discussed in Section 5.1.2, using the statistical methodology described in Appendix A.

Figure 3 (middle panel) shows the ${\bar{\Sigma }}_{\rm pm}$ data as a function of the projected distance R from the cluster center. The results match well with those from vdV06 in the region where the data sets overlap. For the individual data points of ${\bar{\Sigma }}_{\rm pmt}$, ${\bar{\Sigma }}_{\rm pmr}$, and ${\bar{\Sigma }}_{\rm pm}$, we used binning in annuli that are 1'' wide. For the central aperture, we used a circle of radius 1farcs55, which gives an average radius R = 1farcs1 for the data points enclosed by it. This central aperture is larger than the ∼1'' uncertainty in the position of the cluster center derived in Paper I. This uncertainty should therefore not significantly affect the inferred kinematics. The adopted binning gives acceptably small error bars in all apertures, without losing too much spatial resolution near the center.

One immediate result from the inferred ${\bar{\Sigma }}_{\rm pm}(R)$ profile is that there is no strong increase toward the center, consistent with arguments presented in Paper I. A straight line fit to the $(R,{\bar{\Sigma }}_{\rm pm})$ data for RRmax, with Rmax = 15'', yields a slope of −0.06 ± 0.08 km s−1 arcsec−1. The slopes inferred for other values of Rmax in the range 0''–20'' are statistically consistent with this. This analysis indicates that the central data points are consistent with a flat profile of RMS velocity as a function of radius.

The error bars on the observed RMS velocities can be reduced by increasing the amount of binning. For the ratio of ${\bar{\Sigma }}_{\rm pmt}/{\bar{\Sigma }}_{\rm pmr}$ shown in Figure 3 (bottom panel), we used binning in annuli of 8'' wide, so that even small deviations from isotropy can be measured. As for ${\bar{\Sigma }}_{\rm pm}$, the results match well with those from vdV06 in the region where the data sets overlap. The average over all HST data points is ${\bar{\Sigma }}_{\rm pmt}/{\bar{\Sigma }}_{\rm pmr}= 0.983 \pm 0.006$. Hence, there is a very small, but significant, amount of radial anisotropy in the central arcminute of ω Centauri.

5.2.3. Proper Motion Rotation

Our data calibration of Paper I (see discussion in Section 3.6.4 of that paper) removed by necessity any possible solid-body rotation component from the data. This is because we did not have access to stationary background sources in the HST fields, as would be required to measure absolute motions. This calibration has the advantage that it automatically removes any perspective rotation present in the actual proper motions (however, this is negligible for the small radii R ⩽ 71farcs7 sampled by our data anyway; ⩽0.0035 mas yr−1 = 0.08 km s−1 at the canonical D = 4.8 kpc). But of course, it has the disadvantage that it also removes a solid-body rotation fit to the actual rotation field of the cluster. This is a fundamental limitation of the data. However, we discuss below that this has only negligible importance for our models.

In Paper I, we also used a "local correction" that removed all other mean systematic motions (as opposed to random motions) from the proper motion data. However, that was by choice, and not by necessity. The motivation for this is to calibrate out small time variations in the higher order geometric distortion terms. This improves the measurement of the random motions in the cluster. However, one would of course not want this to remove an important signal that is present in the data, and in particular, any possible differential rotation in the plane of the sky (i.e., the part of the rotation field that remains after subtraction of a solid-body fit). We tested for such differential rotation in Paper I, and found that it is negligible in the central HST field. Proper motion rotation curves (the mean tangential motion as function of radius) derived from data with and without the local correction agree to within 1 km s−1 at all radii. Therefore, solid-body rotation is the only unrepresented rotation component that could affect the modeling.

In principle, the loss of information on solid-body rotation can be partly recovered as in vdV06. They use the fact that for an equilibrium axisymmetric system the equation

Equation (33)

must hold everywhere on the projected plane of the sky. Here i is the inclination, y is the projected minor axis direction, and the symbol ${\cal V}$ is used to distinguish a mean proper motion from a mean velocity V. However, to use this equation with accuracy for our HST proper motions would require many more stars with LOS velocities than are actually available at R ⩽ 71farcs7.

The pmr direction is perpendicular to the direction of absolute rotation, and our estimates of ${\bar{\Sigma }}_{\rm pmr}$ are therefore not affected by our inability to measure absolute rotation. Also, in an equilibrium system there cannot be mean streaming in the pmr direction, when averaged along a circle. However, our calculated values of ${\bar{\Sigma }}_{\rm pmt}= ({\cal V}_{\rm pmt}^2 + \Sigma _{\rm pmt}^2)^{1/2}$ will be systematically low by a multiplicative factor g ⩽ 1, because we are unable to include any actual solid-body contribution from ${\cal V}_{\rm pmt}^2$. The proper motion data presented by vdV06 did include measurements of absolute rotations in the plane of the sky. Figure 4 shows for their data the ratios ${\cal V}_{\rm pmt}/\Sigma _{\rm pmt}$ and $g \equiv \Sigma _{\rm pmt} / {\bar{\Sigma }}_{\rm pmt}$, where the quantities in the nominator and denominator are all averages over an annulus. These measurements allow us to estimate the value of g for the HST data. The central data point shows explicitly that there is very little rotation in ω Centauri at the radii of our central HST field. More generally, in the central 10 arcmin, the rotation rate is approximately linear with radius, ${\cal V}_{\rm pmt}/\Sigma _{\rm pmt} \approx R/750^{\prime \prime }$. Therefore, g ≈ [1 + (R/750'')2]−1/2. These approximations are shown as solid lines in the figure. This approximation implies that 0.995 ⩽ g ⩽ 1 for the radii R ⩽ 71farcs7 pertaining to the HST sample.

Figure 4.

Figure 4. Observed ratios ${\cal V}_{\rm pmt}/\Sigma _{\rm pmt}$ (bottom; big dots; magenta in the online version) and $g \equiv \Sigma _{\rm pmt} / {\bar{\Sigma }}_{\rm pmt}$ (top; small dots; magenta), as a function of the projected radius R from the cluster center. The quantities in the nominator and denominator are all averages over an annulus, and are based on the data of vdV06. The solid lines indicate simple approximations to these quantities as described in the text. The dashed vertical line indicates the maximum radius for which we use HST proper motions from Paper I. The results shown here indicate that our inability to determine ${\cal V}_{\rm pmt}$ from the HST data does not prevent us from accurately determining the RMS motion ${\bar{\Sigma }}_{\rm pmt}$ in the tangential proper motion direction.

Standard image High-resolution image

Given what is already known about proper motion rotation from the vdV06 data, our estimates of ${\bar{\Sigma }}_{\rm pmt}$ should be quite accurate. This is despite the fact that any solid-body part of the ${\cal V}_{\rm pmt}$ field cannot be measured directly in the HST data. In particular, the value of g is insufficient to explain the observed anisotropy ${\bar{\Sigma }}_{\rm pmt}/{\bar{\Sigma }}_{\rm pmr}= 0.983 \pm 0.006$. Also, the data of vdV06 show that Vloslos is about a factor of 2 lower than ${\cal V}_{\rm pmt}/\Sigma _{\rm pmt}$. Hence, the influence of rotation is even more negligible for the LOS measurements at R = 12'' of NGB08 (see the discussion in Section 5.1.1).

5.2.4. Gauss–Hermite Moments

To analyze the shapes of the HST proper motion distributions, we binned the data in concentric circular annuli as in Section 5.2.2. For each bin, we created separate histograms of the proper motions in the pmt and pmr directions. We used bins that correspond to 1 km s−1 for the canonical distance of ω Centauri (D = 4.8 kpc; vdV06). For each histogram, we calculated the best-fitting Gaussian and the corresponding Gauss–Hermite moments, as in van der Marel et al. (2000). The odd moments are zero for distributions that are averaged over circles. Figure 5 shows the lowest order non-trivial even Gauss–Hermite moments h4 and h6. These were calculated for annuli that are 8'' wide. The average values over all radii R ⩽ 71farcs7 are h4,pmr = −0.022 ± 0.006, h4,pmt = −0.024 ± 0.006, h6,pmr = −0.003 ± 0.006, and h6,pmt = 0.005 ± 0.006. The 8th- and 10th-order moments were also calculated, but are not shown here. They were found to be consistent with zero, like h6. For all moments, the values for the pmr and pmt directions were found to be consistent within the errors. The only moment that is non-zero is h4, with an average value over the two proper motion directions of h4 = −0.023 ± 0.004. This is indicative of proper motion distributions that are slightly more flat-topped than a Gaussian, and have slightly less extended wings. We will compare this to model predictions in Section 6.7.

Figure 5.

Figure 5. Lowest order even Gauss–Hermite moments h4 (top) and h6 (bottom) of the observed HST proper motion distributions as a function of the projected radius R from the cluster center. Different symbols indicate distributions in the pmr (solid dots) and pmt (open circles) directions. Data points in the two directions were slightly offset horizontally for visual clarity. The Gauss–Hermite moments are not far from zero, indicating that the proper motion distributions are close to Gaussian. There are no systematic trends with radius. The curves show predictions of three isotropic models, as discussed in Section 6.7. The long-dashed curves (red in the online version) are for the isotropic core model with no dark mass. The solid curves (blue) are for the isotropic cusp model with its best-fit IMBH mass MBH = 1.8 × 104M. Both of these models match the observed Gauss–Hermite moments to an average |Δhi| ≲ 0.01. The short-dashed curve (green) is for the isotropic cusp model with no dark mass. This model does not fit the observed Gauss–Hermite moments.

Standard image High-resolution image

6. DYNAMICAL DATA–MODEL COMPARISON

6.1. Velocity Anisotropy

The observed profile of ${{\bar{\Sigma }}_{\rm pmt}}/{{\bar{\Sigma }}_{\rm pmr}}$ shows statistically significant deviations from unity, both at small and at large radii (see Figure 3 (bottom panel)). It is therefore important to allow for anisotropy in our models for ω Centauri.

To illustrate the relation between the model anisotropy function β(r) and the observed quantity ${{\bar{\Sigma }}_{\rm pmt}}/{{\bar{\Sigma }}_{\rm pmr}}(R)$, we show in Figure 6 (top) the predicted ${{\bar{\Sigma }}_{\rm pmt}}/{{\bar{\Sigma }}_{\rm pmr}}(R)$ for models with constant β as a function of radius r. Models were calculated with $(\overline{v_{\rm tan}^2}/\overline{v_r^2})^{1/2}$ ranging from 2/3 to 3/2, in equal logarithmic steps. At large radii, the proper motion anisotropy is a good tracer of the intrinsic anisotropy, consistent with the arguments presented in Section 2.7.3. For example, at R = 10 arcmin, $({{\bar{\Sigma }}_{\rm pmt}}/{{\bar{\Sigma }}_{\rm pmr}}) - 1 = \tau [(\overline{v_{\rm tan}^2}/\overline{v_r^2})^{1/2} - 1]$, with τ = 0.7–0.9 for the models calculated here, and τ ≈ 0.8 for β ≈ 0. At small radii, the behavior of ${{\bar{\Sigma }}_{\rm pmt}}/{{\bar{\Sigma }}_{\rm pmr}}$ is a more complicated function of β. Tangentially anisotropic models (β < 0) all converge to ${{\bar{\Sigma }}_{\rm pmt}}/{{\bar{\Sigma }}_{\rm pmr}} = 1$ near the center, while radially anisotropic models (0 < β ⩽ 1) predict ${{\bar{\Sigma }}_{\rm pmt}}/{{\bar{\Sigma }}_{\rm pmr}} < 1$.

Figure 6.

Figure 6. Dynamical data–model comparisons for the ratio ${\bar{\Sigma }}_{\rm pmt}/{\bar{\Sigma }}_{\rm pmr}$ as a function of the projected radius R from the cluster center. The data and the symbols are the same as in Figure 3 (bottom panel), and are the same in both panels. Top panel: anisotropic models with constant β as a function of radius. The models shown have $(\overline{v_{\rm tan}^2}/\overline{v_r^2})^{1/2}$ ranging from 2/3 to 3/2, in equal logarithmic steps. The corresponding β values are indicated for some of the curves. The horizontal line at ${\bar{\Sigma }}_{\rm pmt}/{\bar{\Sigma }}_{\rm pmr}= 1$ is for the isotropic model. Tangentially anisotropic models have higher ${\bar{\Sigma }}_{\rm pmt}/{\bar{\Sigma }}_{\rm pmr}$ than radially anisotropic models. The data are not well fit by models with constant β. Bottom panel: anisotropic model with β(r) in Equation (15) chosen the optimize the fit to the data (blue in the online version). The model has β0 = 0.13, β = −0.53, and ra = 729''. The predictions in this figure were calculated for a core model with no dark mass. However, the predicted ratio ${{\bar{\Sigma }}_{\rm pmt}}/{{\bar{\Sigma }}_{\rm pmr}}(R)$ depends mostly on β(r), and is virtually indistinguishable when considering instead models with a cusp or dark mass.

Standard image High-resolution image

It is evident from Figure 6 (top) that the observed ${{\bar{\Sigma }}_{\rm pmt}}/{{\bar{\Sigma }}_{\rm pmr}}(R)$ must tightly constrain the intrinsic β(r). We fitted the combined kinematical data from Figure 3 with models that spanned a grid in β0, β and ra. The parameter β0 determines the behavior of ${{\bar{\Sigma }}_{\rm pmt}}/{{\bar{\Sigma }}_{\rm pmr}}(R)$ near the center, β determines the behavior at large radii, and ra is the transition radius. The minimum χ2 was obtained for β0 = 0.13 ±  0.02, β = −0.53 ± 0.22 and log ra[arcsec] = 2.86 ± 0.12. The error bars were obtained from the χ2 contours of the fit.

The predicted ${{\bar{\Sigma }}_{\rm pmt}}/{{\bar{\Sigma }}_{\rm pmr}}(R)$ for the best-fit model is shown in Figure 6 (bottom). It provides an excellent fit. The predictions shown in the figure were calculated for core models with no dark mass. However, the predicted ratio ${{\bar{\Sigma }}_{\rm pmt}}/{{\bar{\Sigma }}_{\rm pmr}}(R)$ depends mostly on β(r), and is virtually indistinguishable when considering instead models with a cusp or dark mass.

The error bars of the fit show that velocity anisotropy is detected at high statistical significance in ω Centauri. The parameters β0 and β imply that the ratio $\overline{v_{\rm tan}^2}/\overline{v_r^2}$ transitions from 0.935 near the center to 1.235 at large radii. These are of course not necessarily the values at the very center and at infinity, since our model is constrained by proper motion data only over the range 1–1000 arcsec. Either way, there is a transition from radial anisotropy at small radii to tangential anisotropy at large radii. The parameter ra indicates that the transition occurs at approximately 12 ± 3 arcmin (5 ± 1 core radii). This is broadly consistent with the modeling results of vdV06. They found (their Section 9.2) that ω Centauri is slightly radially anisotropic for r ≲ 10 arcmin and that it becomes increasingly tangentially anisotropic outside this region.

The centers of globular clusters are generally believed to have isotropic velocity distributions because two-body relaxation tends to isotropize the orbits. However, ω Centauri has a long half-mass relaxation time of ∼109.96±0.03 yr (McLaughlin & van der Marel 2005). We showed in Paper I that mass segregation in the central region has not yet progressed to the point of energy equipartition. Since full energy equipartition has not yet been achieved, it is not surprising that complete velocity isotropy has not yet been achieved either.

A sufficient number of stars are necessary to measure the proper motion anisotropy with accuracy. Due to the finite number of stars at small radii, the anisotropy is not as well constrained at radii R ≲ 10'' as it is further out. However, because Omega Cen has a very large core, most of the stars that are observed near the projected center are not actually close to the center in three dimensions (see, e.g., Section 6.1 of Paper I). One consequence of this is that the projected kinematics predicted near the center are influenced primarily by the anisotropy at larger radii. Changing the model anisotropy in the central 10'' even quite substantially (e.g., ±30% in σrt) does not change the model predictions for the HST proper motion kinematics by more than a fraction of the error bars, even for the innermost data point. Therefore, uncertainties in anisotropy near the center do not significantly impact our predictions or conclusions.

6.2. Cusp versus Core Models

Figures 7 (top) and (middle) show the predicted and observed RMS projected velocities for the best-fit β(r). Each panel of this figure shows the data for ${\bar{\sigma }}_{\rm los}(R)$ and ${\bar{\sigma }}_{\rm pm}(R)$ together in the same panel, with the observed proper motions ${\bar{\Sigma }}_{\rm pm}(R)$ transformed to km s−1 using the best-fit distance implied by the model. For each model, there are separate curves predicted for ${\bar{\sigma }}_{\rm los}(R)$ and ${\bar{\sigma }}_{\rm pm}(R)$ (dashed and solid curves, respectively), since these quantities are not the same in an anisotropic model. However, the difference is generally smaller than the observational error bars over the radial range where both types of data are available. There is no LOS data available for R ≲ 10'', so for visual clarity we do not show the predicted ${\bar{\sigma }}_{\rm los}(R)$ at those radii. At small radii ${\bar{\sigma }}_{\rm los}(R)$ exceeds ${\bar{\sigma }}_{\rm pm}(R)$ because of the model anisotropy, with ${\bar{\sigma }}_{\rm los}-{\bar{\sigma }}_{\rm pm}\approx 0.9 \;{\rm km}\;{\rm s}^{-1}$ at R = 10'' and ${\bar{\sigma }}_{\rm los}-{\bar{\sigma }}_{\rm pm}\approx 1.1 \;{\rm km}\;{\rm s}^{-1}$ at R = 1''.

Figure 7.

Figure 7. Dynamical data–model comparisons for the RMS projected velocity as a function of the projected radius R from the cluster center. The data and the symbols are the same as in Figure 3, and are the same in all panels. For plotting purposes, the proper motions ${{\bar{\Sigma }}_{\rm pm}}$ in mas yr−1 were transformed to ${\bar{\sigma }}_{\rm pm}$ in km s−1 (solid symbols) using Equation (25) with the best-fitting distance implied by the models. The arrow in the bottom panel indicates how one of the two data points from NGB08 moved due to our improved determination of the cluster center in Paper I. Curves are the predictions from various models. Solid curves are for ${\bar{\sigma }}_{\rm pm}$ and short-dashed curves are for ${\bar{\sigma }}_{\rm los}$. For visual clarity, ${\bar{\sigma }}_{\rm los}$ is not shown for R ⩽ 10'', where no ${\bar{\sigma }}_{\rm los}$ data are available. At larger radii, ${\bar{\sigma }}_{\rm pm}$ and ${\bar{\sigma }}_{\rm los}$ are quite similar, with differences becoming visual only at R ≳ 1000''. The curves that predict the lowest ${\bar{\sigma }}$ at small radii (blue) are models with no dark mass. The remaining curves (red) have some kind of dark mass in the center. Top panel: anisotropic models in which β(r) was optimized to fit the data (see Figure 6), based on the core model for the projected intensity (see Figure 2). The top curve has the best-fit IMBH of mass MBH = 4.1 × 103M. Middle panel: Anisotropic models similar to the top panel, but now based on the cusp model for the projected intensity (see Figure 2). The top curve has the best-fit IMBH of mass MBH = 8.7 × 103M. Bottom panel: models similar to the middle panel, but now for an isotropic velocity distribution (which does not fit the data for ${\bar{\Sigma }}_{\rm pmt}/{\bar{\Sigma }}_{\rm pmr}(R)$ in Figure 6). The quantities ${\bar{\sigma }}_{\rm los}$ and ${\bar{\sigma }}_{\rm pm}$ are the same for these isotropic models. The top three curves have different dark masses. The solid curve is for the best-fitting IMBH mass MBH = 1.8 × 104M. The dotted curve is for the IMBH mass MBH = 4.0 × 104M advocated by NGB08. The long-dashed curve is for the best-fitting dark cluster, which has Mdark = 2.0 × 104M and adark = 3farcs0. It follows from the analysis of the predictions in this figure, and in particular the top panel, that the presence of an IMBH in ω Centauri is not required, although IMBHs with masses MBH ≲ 1.2 × 104M cannot be ruled out at ∼1σ confidence.

Standard image High-resolution image

Figure 7 (top) shows the predictions for the core model fit to the projected intensity and Figure 7 (middle) for the cusp model fit. The curves that predict the lowest ${\bar{\sigma }}$ at small radii (blue) are models with no dark mass. The core and cusp models generally predict the same dynamics at large radii, but differ slightly at small radii.

It is true in many analytical models with a cusped density profile and no central dark mass that the projected RMS velocity decreases toward the center (e.g., Tremaine et al. 1994). The same can be seen in Figure 7 (middle) for our model with γ = 0.05. Even though the cusp is quite shallow in projection, this corresponds to a much steeper three-dimensional luminosity density jr−1−γ at asymptotically small radii. Similarly, the decrease in RMS velocities toward the center is much stronger intrinsically in three dimensions than it is in projection. Either way, the observed RMS velocities do not show a significant dip near the center. In the absence of a dark mass, the core models therefore provide a better fit to the data than the cusp models.

The best fits without a dark mass have χ2 = 235.3 and 249.2 for the core and cusp models, respectively. There are 202 data points, yielding NDF = 197 degrees of freedom (there are five free parameters in the models when there is no dark mass). One expects $\chi \approx N_{\rm DF} \pm \sqrt{2 N_{\rm DF}}$. The anisotropic core model is therefore consistent with the data at the 1.9σ level, while the cusp model is consistent at the 2.6σ level. These confidence levels assume that all data points in the fit are statistically independent, which is somewhat of a simplification. It is likely that samples from different LOS velocity studies have some stars in common. However, this cannot be fully explored here since Scarpa et al. (2003) and Sollima et al. (2009) did not list individually which stars they observed.

Much of the excess E2) ≡ χ2NDF > 0 for the fits is due to just two data points for ${\bar{\sigma }}_{\rm los}$ that appear anomalously high. The first "discrepant" data point is the measurement ${\bar{\sigma }}_{\rm los}= 23.0 \pm 2.0 \;{\rm km}\;{\rm s}^{-1}$ at R = 12farcs4 by NGB08. The second discrepant data point is the average ${\bar{\sigma }}_{\rm los}= 19.9 \pm 1.0 \;{\rm km}\;{\rm s}^{-1}$ for an annulus centered at R = 46farcs0, obtained from the compiled data of vdV06. The former contributes 6.8 to the χ2 of the fit for the core model, while the latter contributes 11.3. This adds up to 18.1, which is almost half of E2) = 38.3.

The discrepant NGB08 data point pertains to similar radius as the other field that they observed, which is at R = 12farcs1 from the cluster center (see Paper I). The measurement ${\bar{\sigma }}_{\rm los}= 18.6 \pm 1.6 \;{\rm km}\;{\rm s}^{-1}$ for that latter field is more consistent with the other measurements in our combined data set from Section 5. The two NGB08 measurements are inconsistent with each other at the 1.7σ level, despite the fact that that they pertain to the same radius, and despite the fact that the proper motions in these fields are consistent with each other (Paper I). We believe that this may be due to underestimates in the NGB08 error bars, since they did not explicitly include the contribution of shot noise (from the finite number of stars) to their error bars. Similarly, the measurement in the vdV06 annulus at R = 46farcs0 appears inconsistent with other measurements in that same data set. The weighted average of the two adjacent annuli is lower and inconsistent with it at the 2.1σ level. This suggests that both discrepant measurements that contribute disproportionately to χ2 may be spuriously high for unknown reasons.

Evidently, there are some internal discrepancies in the data sets themselves. Hence, the observed values of E2) should not be taken as a sign of shortcomings in the models. In fact, the values of E2) are remarkably low, given that we are fitting kinematical data compiled from 10 different original sources. Potential uncorrected systematics between different studies could easily have induced much larger data–model discrepancies.

One clear result from Figures 7 (top) and (middle) is that the data are well fit by the models out to the largest available radii. This contradicts arguments put forward by Scarpa et al. (2003) that the large radii kinematics of ω Centauri may be inconsistent with traditional theories of gravity. Our conclusion on this issue agrees with that of McLaughlin & Meylan (2003) and Sollima et al. (2009).

6.3. IMBH

To test for the presence of an IMBH, we have also constructed models with MBH > 0 (and the same β(r) profile). Models with no IMBH have one free parameter less than models with an IMBH. One would therefore on average expect them to fit worse by Δχ2 = 1. Moreover, random variations can add another Δχ2 = 1 (at 1σ confidence), even when the no-IMBH model is correct. Therefore, an IMBH is statistically significant in the model only if it improves the fit over a no-IMBH model by Δχ2 ≳ 2.

Figure 8 shows the curves of χ2(MBH) that quantify the quality of the fit. For the core model, the best-fit IMBH has mass MBH = 4.1 × 103M. This improves the fit over the no-IMBH model by only Δχ2 = 1.8. This is not statistically significant. For the core model, we can therefore determine only an upper limit to the mass of a possible IMBH, namely MBH ⩽ 7.4 × 103M at 1σ confidence. Figure 7 (top) compares the best-fit core models with and without an IMBH. The inclusion of the IMBH makes very little difference to the predictions. The difference for the innermost data point at R = 1farcs1 is only 0.8 km s−1, which is less than a quarter of the observational error bar at that radius.

Figure 8.

Figure 8. χ2 of the model fits to the kinematic data in Figure 3, as a function of IMBH mass MBH. The resulting parabolae are, from bottom left to top right, respectively, obtained for the following models: anisotropic models with a core (black in the online version; see Figure 7 (top panel) for actual model predictions), anisotropic models with a cusp (red; see Figure 7 (middle panel)), and isotropic models with a cusp (blue; see Figure 7 (bottom)). Centered above the minimum of each parabola is a triangle with an associated error bar that indicates the best-fit IMBH mass and its formal 1σ uncertainty. The anisotropic core models provide the overall best fits to the kinematics. The IMBH that optimizes the χ2 for those models is not statistically significant. The small difference in χ2 (vertical height) between the parabolae for the anisotropic core and cusp models is not a reason to prefer to core models over cusp models, in particular because the latter actually provide a somewhat better fit to the photometry (see the discussion in the text).

Standard image High-resolution image

For the cusp model, the inclusion of an IMBH improves the fit by Δχ2 = 9.8, which is significant at the 3.1σ level. The best-fit IMBH mass is MBH = (8.7 ± 2.9) × 103M (with the error bar determined by the requirement Δχ2 ⩽ 1). Figure 7 (middle) compares the best-fit cusp models with and without an IMBH. The inclusion of the IMBH provides a small but visible improvement in the fit near the center. The reason that the IMBH is more massive and significant in the cusp model is due to the dip in RMS velocity predicted by these models when there is no dark mass.

In assessing the meaning of formal error bars on MBH, it is important to keep in mind what one may call "black hole bias." Statistical error bar estimates take into account only the random Gaussian uncertainties in the data. They do not take into account the systematic residuals that are inevitably present in any astronomical data–model comparison. In the present context, these may include oversimplifications in the model assumptions (e.g., sphericity, equilibrium, the adopted density parameterization, no mass segregation among the luminous stars), the data reduction process (e.g., inability to measure absolute proper motion rotation), or the meaning of the data themselves (e.g., inability to separate orbital motion from potential binary motion). By adding MBH as an extra free parameter to the fit, the model gains in ability to fit these systematic residuals that are not actually due to an IMBH. Because masses are positive definite (MBH ⩾ 0), this generally tends to bias in the direction of invoking more mass then there actually is, as compared to the opposite. Hence, MBH tends to be biased high. Marginally significant non-zero MBH values with no telltale signs of an IMBH (e.g., fast-moving stars or an R−1/2 increase in ${\bar{\sigma }}$) may simply be indications of small systematic issues in the data–model comparison unrelated to an IMBH.

The core model with no IMBH fits the kinematical data a little better than the cusp model with an IMBH (see Figure 8). The difference is Δχ2 = 4.0, which is nominally significant at the 2.0σ level. However, as discussed in Section 4, the cusp model nominally provides a better fit to the photometric data. The difference for a generalized nuker fit is Δχ2 = 6.8, which is nominally significant at the 2.6σ level. One could in principle construct a grand total Δχ2 that sums the photometric and kinematic values. However, we have not done this here. The arguments about whether a core or cusp model better fits the photometry are complex (see Section 4). They are not easily captured in a single number without having to make arbitrary choices about, e.g., what radii to include in the comparison. Either way, when comparing both the photometric and kinematic information for the core and cusp models we feel that there is no sound statistical basis to prefer one model over the other. In view of this, we conclude that the existing data do not imply that an IMBH is present in ω Centauri. However, IMBHs with masses MBH ≲ 1.2 × 104M cannot be ruled out at 1σ confidence (or ≲1.8 × 104M at 3σ confidence).

6.4. Distance and Mass-to-Light Ratio

The best-fit distance and mass-to-light ratio depend only slightly on the particular model we adopt. The core model without a dark mass has D = 4.70 ± 0.06 kpc and ϒV = 2.64 ± 0.03 (here and henceforth, mass-to-light ratios are given in V-band solar units). The cusp model with an IMBH has D = 4.75 ± 0.06 kpc and ϒV = 2.59 ± 0.03.

Our results for D and ϒV are entirely consistent with those inferred by vdV06 from fitting only ground-based projected intensity and kinematical data. They found that D = 4.8 ± 0.3 kpc and ϒV = 2.5 ± 0.1. Their modeling technique was considerably more sophisticated than ours, allowing for axisymmetry, inclination, ellipticity variations with radius, and arbitrary profiles of rotation and velocity-dispersion anisotropy. The reason that our error bars are smaller than theirs is likely a combination of two effects. First, we have considerably more data at our disposal than did vdV06, and in particular we have a factor ∼11 more stars with measured proper motions. This produces a significant decrease in the random uncertainties. But second, in our simpler models we cannot explore the full variety of phase-space distributions that might fit the data with non-standard distances or mass-to-light ratios. Therefore, we are not quantifying systematic errors as rigorously as vdV06, and as a result our errors are artificially lowered compared to theirs.

The fact that our D and ϒV agree with those of vdV06 indicates that there is no reason to mistrust that our spherical models can address the central mass distribution of ω Centauri with credible accuracy. This is further supported by the facts that both the elongation (Geyer et al. 1983) and the rotation rate (Section 5.2.3) decrease toward the cluster center. To get good results, we did have to make sure to model RMS velocities (which sum rotation velocities and velocity dispersions in quadrature) and not merely velocity dispersions. Also, one should fit only quantities that are properly averaged over annuli on the projected plane of the sky. If these things are not done, then biased estimates are obtained for D and ϒV, as demonstrated in Appendix C of vdV06.

6.5. Isotropic Models

Isotropic models predict that ${{\bar{\Sigma }}_{\rm pmt}}/{{\bar{\Sigma }}_{\rm pmr}} = 1$ at all radii, so they are not generally appropriate for ω Centauri. However, isotropic models have the advantage that the full velocity distributions ${\cal L}_{\rm iso}(v,R)$ can be calculated as described in Section 2.5. Also, isotropic models are often used for globular clusters, so it is useful to understand how the predictions of such models deviate from more general anisotropic models.

Figure 8 shows the curve of χ2(MBH) for isotropic models with an IMBH. The models fit significantly worse than the anisotropic models, as shown by the higher χ2. Within the realm of isotropic models, the best fit is provided by MBH = (1.8 ± 0.3) × 104M. This model has D = 4.81 ± 0.06 kpc and ϒV = 2.61 ± 0.03. Figure 7 (bottom) shows the data–model comparison for the isotropic model with this IMBH. We also show the predictions for MBH = 0, as well as for the value MBH = 4.0 × 104M advocated by NGB08.

The NGB08 IMBH mass was based on their measurement of ${\bar{\sigma }}_{\rm los}= 23.0 \pm 2.0 \;{\rm km}\;{\rm s}^{-1}$ at a position they believed to be the cluster center. This high IMBH mass is now clearly ruled out, even for an isotropic model, for two reasons. First, Paper I showed that this measurement actually applies to a larger distance R = 12farcs4 from the cluster center. The arrow in Figure 7 (bottom) demonstrates how this moved the corresponding LOS velocity measurement. And second, the high measured RMS velocity was not confirmed by proper motion measurements, either at the same position or on the actual cluster center. Note in Figure 7 (bottom) that the isotropic model with the NGB08 IMBH does appear to match well the data at R = 5''–7''. However, it overpredicts the data at both R ⩽ 4'' and R = 8''–10''. Since all data points were derived in the same manner and are equally valid, there is no reason attribute more weight to the data at R = 5''–7''. In particular, the data at R ⩽ 4'' contain most of the information on the very central mass distribution, and they clearly rule out an IMBH mass as high as advocated by NGB08 (independent of the model anisotropy near the center).

It is evident from Figure 8 that isotropic models require a higher IMBH mass than the best-fit anisotropic models. This is not generally true for isotropic models, but is a specific consequence of the fact that ω Centauri is radially anisotropic near the center and tangentially anisotropic at large radii. This affects the general gradient in RMS velocity from small to large radii (see e.g., Figures 10 and 11 of van der Marel 1994a). The anisotropic model without a dark mass has a steeper gradient toward the center than the corresponding isotropic model. The anisotropic model therefore requires less dark mass to fit the observed gradient. This indicates that one should be careful in application of isotropic models to studies of IMBHs. Depending on the exact anisotropy of the cluster, the assumption of isotropy can lead to either an overestimated or an underestimated IMBH mass, or it can require an IMBH when none is present.

6.6. Dark Cluster

When a dark mass is clearly required by the data, as in the case of isotropic models with a cusp, one can ask whether an extended dark cluster may provide a better fit than an IMBH. Figure 9 shows a two-dimensional contour plot of Δχ2 in the parameter space of (adark, Mdark) for the isotropic cusp model. The best fit is obtained for Mdark = 2.0 × 104M and adark = 3''. The predictions of this model are shown in Figure 7 (bottom). It has a somewhat shallower increase in RMS velocity toward the center than the IMBH model. The dark-cluster model makes an improvement of only Δχ2 = 1.7 over the IMBH model, while increasing the number of free parameters by one. This is not statistically significant. Therefore, one can interpret Figure 9 to mean that if there is a dark mass in ω Centauri, then its extent must be ≲7'' at 1σ confidence.

Figure 9.

Figure 9. Contours of Δχ2 in the two-dimensional plane of (adark, Mdark) for the fit of isotropic cusp models to the data in Figure 3. The distance D and mass-to-light ratio ϒV were optimized separately at each (adark, Mdark) for which a model was calculated (indicated with small dots). The model for which the minimum χ2 was obtained is indicated by the big dot. The first three contours (solid) around the minimum correspond to 1σ, 2σ, and 3σ (bold) confidence contours, respectively. Each subsequent contour (dashed) increases Δχ2 by a constant factor. Models with a central dark points mass (an IMBH; i.e., adark = 0) are shown on the vertical axis on the left. Models with no central dark mass lie on the horizontal axis on the bottom. If there is a dark mass in ω Centauri (as must be the case in isotropic models, but these do not fit the data for ${\bar{\Sigma }}_{\rm pmt}/{\bar{\Sigma }}_{\rm pmr}(R)$ in Figure 6) then its extent must be ≲7'' at 1σ confidence.

Standard image High-resolution image

The central density of the best-fitting dark cluster is ρ0,darkMdark/(4πa3dark/3), according to Equation (7). Using Equation (4) to transform angular units to physical units, this yields for the core model ρ0,dark = 1.4 × 107M pc−3. The central density of luminous matter in this model is only ρ0,lum = 3.3 × 103M pc−3. This implies a central M/L = ϒV[1 + (ρ0,dark0,lum)] = 1.1 × 104. This applies to an isotropic model with Mdark = 2.0 × 104M. For an anisotropic model with a lower Mdark, the required M/L scales approximately linearly with Mdark. The (maximum) size adark of the dark mass fit is relatively insensitive to the exact anisotropy. Either way, if ω Centauri does indeed have a dark central cluster, this would be indicative of quite an extreme amount of mass segregation or a very top-heavy initial mass function. If a dark mass is invoked to explain the data, then an IMBH may be easier to explain than such a concentrated dark cluster.

6.7. Velocity Distribution Shapes

For our isotropic models, we calculated the projected velocity distributions ${\cal L}_{\rm iso}(v,R)$ as described in Section 2.5, and from this the Gauss–Hermite moments hi,iso(R). We did not take the effect of observational proper motion uncertainties into account in the models (essentially a convolution of the model predictions with the error distribution). This has negligible effect on the Gauss–Hermite moments because the observational errors are much less than the intrinsic cluster dispersion. The curves overplotted in Figure 5 show the model predictions for three models: the isotropic core model with no dark mass, the isotropic cusp model with its best-fit MBH = 1.8 × 104M, and the isotropic cusp model with no dark mass.

The core model with no dark mass and the cusp model with an IMBH both fit the observed Gauss–Hermite moments within the error bars. Over the range 5'' ≲ R ⩽ 71farcs7, the predicted Gauss–Hermite moments for these models are approximately constant. For the core model without dark mass they are h4,iso(R) ≈ −0.013 and h6,iso(R) ≈ −0.002, and for the cusp model with an IMBH they are h4,iso(R) ≈ −0.034 and h6,iso(R) ≈ 0.009. These averages do deviate in statistically significant manner from the observed averages, h4 = −0.023 ± 0.004 and h6 = 0.001 ± 0.004 (see Section 5.2.4). This is not surprising, given that we have already established that ω Centauri does not have an isotropic velocity distribution, not even near its center. However, the differences between the predicted and the average observed Gauss–Hermite moments are quite small |Δhi| ≲ 0.01. For comparison, models that range in anisotropy from fully tangential to fully radial can easily span the range Δhi ≈ ±0.2 (e.g., van der Marel & Franx 1993). The measured Gauss–Hermite moments are therefore consistent with the fact that the velocity distribution is not far from isotropic near the center of ω Centauri.

The cusp model without dark mass does not fit the observed Gauss–Hermite moments. As discussed in Section 6.2, this model has a strong decrease in its intrinsic (unprojected) RMS velocity toward the center. The decrease in the projected RMS velocity is relatively mild, as shown in Figure 7 (bottom). However, there is a strong gradient in the RMS velocity along the LOS. This leads to strongly non-Gaussian velocity distributions. In a central aperture of R = 8'' radius, as used for the innermost observational data point in Figure 5, the predictions are h4,iso = 0.063 and h6,iso = −0.063. This deviates for each order by ∼3σ from the observed values h4 = −0.026 ± 0.029 and h6 = 0.031 ± 0.029 (averaged over the pmr and pmt directions). While this result was derived for isotropic models, the large predicted deviations from Gaussian velocity distributions are likely generic, and would exist in anisotropic models as well. So if ω Centauri has a central cusp in its projected intensity profile, then the observed velocity distributions rule out models without a dark mass. Of course, Figure 7 (middle) and (bottom) show that such models also do not provide good fits to the projected profiles of RMS velocity.

Figure 5 shows that an isotropic model with an IMBH predicts increased h4 and h6 in the central R ≲ 5'', corresponding to profiles with broader wings than a Gaussian. To do a more detailed data–model comparison for the wings, we show in Figure 10 the histograms of the observed proper motions for apertures of radii R = 3'' and R = 10''. These apertures contain N = 43 and N = 585 stars with well-measured proper motions, respectively. The proper motion coordinates in the pmr and pmt direction were both included in each histogram, as appropriate for an isotropic velocity distribution. The histograms were symmetrized to decrease the uncertainties. The predictions for the isotropic cusp model with MBH = 1.8 × 104M are overplotted as curves. The predicted velocity distribution for the R = 3'' aperture has more extended wings than the observed histogram. The wings are much reduced in the predicted velocity distribution for the R = 10'' aperture. This is because none of the added stars at 3'' < R ⩽ 10'' are close to the black hole, so they do not move at unusually high velocities.

Figure 10.

Figure 10. Symmetrized histograms of the observed proper motion velocities for apertures of radii R = 3'' (bold; red) and R = 10'' (thin; blue), respectively. Proper motions in both the pmr and pmt directions are included. They were transformed to physical units of km s−1 using the best-fit model distance. Typical error bars are shown on the left; they were offset horizontally from the histograms for visual clarity. The predictions for the isotropic cusp model with the best-fit IMBH mass MBH = 1.8 × 104M are overplotted as curves (solid for the R = 3'' aperture, and dashed for the R = 10'' aperture). The inset on the right shows a blow-up of the profile wings on an expanded vertical scale. The models predict broader wings than observed, but the difference is not statistically significant (as discussed in the text). This is because the total area under the model curves beyond 60 km s−1 corresponds to less than 1 predicted star.

Standard image High-resolution image

The observed histograms do not extend beyond 60 km s−1. The maximum observed value of either |vpmt| or |vpmr| inside R = 10'' is 59.4 km s−1. No stars were rejected as non cluster members inside this radius because of their fast motion (see Section 5.2.1). To assess whether these observations are statistically consistent with the wings of the predicted distributions, let f be the fraction of the normalized model distribution at |v|>60 km s−1. According to the models, this is f = 3.8 × 10−3 for the R = 3'' aperture and f = 3.9 × 10−4 for the R = 10'' aperture. The expectation value for the number of stars with a velocity component at |v|>60 km s−1 is E = 2Nf. The probability that all of the 2N measured velocities have |v| ⩽ 60 km s−1 is P = (1 − f)2N. This yields for the R = 3'' aperture that E = 0.32 and P = 0.72, while for the R = 10'' aperture E = 0.46 and P = 0.63.

Since less than a single fast-moving star was predicted, the fact none were observed is not statistically inconsistent with the models. This does not change when we use larger apertures, since the models do not predict any fast-moving stars outside R = 10''. It also does not change by using smaller apertures, since in that case the sample of observed stars becomes very small. Therefore, the absence of fast-moving stars in the data cannot be used to independently rule out the presence of an IMBH with mass MBH = 1.8 × 104M. To increase the predicted number of fast-moving stars one would need to increase the number of stars with measured proper motions, i.e., extend the kinematical observations to fainter magnitudes. We have experimented with looser cuts on our proper motion catalog. This allows us to increase the sample size by a factor ∼2 (see Paper I). In this extended catalog there is no star with a proper motion coordinate in excess of 65 km s−1 within R = 10''. This provides somewhat tighter constraints on the presence of an IMBH, with the quoted values of P roughly decreasing to their squared values. Even then, it is still not possible to rule out an IMBH of mass MBH = 1.8 × 104M with reasonable confidence.

Alternatively, E increases and P decreases when the BH mass is increased. For the IMBH of mass MBH = 4.0 × 104M suggested by NGB08, and sticking as before with our high-quality proper motion catalog of Section 5.2.1, we obtain E = 2.9 and P = 0.05 for the aperture with R = 10''. This implies that the absence of fast-moving stars rules out all IMBHs with MBH ≳ 4.0 × 104M at 95% confidence. This result is not very competitive for ω Centauri, since we already know from modeling of the RMS velocities that any IMBH must have a mass at least 3.3 times lower than this anyway. However, this method has the advantage that it does not require any accurate measurement of RMS velocities. It may therefore be useful for other clusters, in cases when the observational errors may be too large for accurate measurement of the RMS velocity profile.

6.8. Brownian Motion

We have assumed in the modeling that any IMBH must reside at the cluster center. In reality, an IMBH would exhibit Brownian motion as a result of its energy equipartition with surrounding stars. Chatterjee et al. (2002) derived for the one-dimensional RMS offset from the cluster center that

Equation (34)

This was derived for an analytical Plummer model with core radius r0, which provides a reasonable description of (non core-collapsed) globular clusters. Chatterjee et al. considered single-mass models with stars of mass ${\tilde{m}}$. However, their results should be valid for multi-mass models as well, provided that the quantity ${\tilde{m}} \equiv \langle m^2 \rangle / \langle m \rangle$ is defined appropriately in terms of averages over the stellar mass function (Merritt et al. 2007).

We take for r0 the King core radius of 141'' (McLaughlin & van der Marel 2005) and for MBH the upper limit of 1.2 × 104M derived in Section 6.3. For a Salpeter mass function from 0.1–10 M, ${\tilde{m}} = 1.27 \;{M_{\odot }}$. We then obtain xRMS = 0farcs69. This is significantly less than the 1farcs55 radius of the smallest aperture used in the kinematical analysis (see Section 5.2.2). Therefore, the expected Brownian motion of an IMBH should not affect significantly the analysis that we have presented here.

IMBHs with larger masses than those considered here, placed at considerable offsets from the cluster center (consistent with their expected Brownian motion), might have gone unnoticed in our analysis. However, there is no independent motivation to consider such a situation. In Paper I, we showed that the kinematical center of stars in ω Centauri agrees with the star count center to within its ∼2'' uncertainty.

7. GLOBULAR CLUSTER IMBH DEMOGRAPHICS

The self-gravity of a cluster gives it a negative heat capacity that makes it vulnerable to the so-called "gravothermal catastrophe." This makes the core collapse on a timescale proportional to the half-mass relaxation time. This is normally halted by formation and/or heating of binaries. However, the most massive stars in a cluster tend to undergo core collapse more or less independently of the other cluster stars on a timescale that is less than the core collapse time for the cluster as a whole. If this happens faster than the main sequence evolution time of these stars (i.e., before they become compact remnants), then they will undergo physical collisions that lead to the formation of a single very massive star (VMS). Such a VMS could plausibly evolve into an IMBH.

Portegies Zwart & McMillan (2002) proposed the aforementioned scenario and performed N-body simulations to show that a VMS might grow to MVMS/Mtot ≈ 0.1%, with Mtot being the total cluster mass. Gurkan et al. (2004) and Freitag et al. (2006) used Monte Carlo simulations with prescriptions and models for stellar collisions and found that the resulting VMS could even grow up to MVMS/Mtot ≈ 1%. However, it is not clear what the resulting IMBH mass might be. The evolution of a VMS may yield significant mass loss before the final collapse. Also, a VMS will only form through this scenario in clusters with very short initial half-mass relaxation times, ≲30 Myr. Some star clusters (e.g., the Arches cluster) are known to have such short relaxation times. However, most globular clusters have much longer current relaxation times, although it is not straightforward to estimate what their relaxation time at birth may have been (Ardi et al. 2008).

Alternative scenarios for IMBH formation in globular clusters have been proposed as well (see e.g., the review in van der Marel 2004). This includes the merging of stellar-mass black holes. O'Leary et al. (2006) found that this may lead to growth up to MIMBH/Mtot ≈ 0.1%. Whether this indeed happens depends on the typical BH merger recoil speed. Significant recoils will kick black holes out of the cluster before they have a chance to grow.

Overall, the predictions for IMBH formation in globular clusters have significant uncertainties. Observations will therefore need to be the primary guide for the study of IMBH demographics in globular clusters. For ω Centauri one may combine the distance and mass-to-light ratio derived in Section 6.4 with the extinction-corrected total V-band magnitude to determine that Mtot = 2.8 × 106M. The IMBH mass upper limit from Section 6.3 therefore corresponds to MBH/Mtot ≲ 0.43%. Table 1 lists the corresponding constraints for the other globular clusters that have been well studied dynamically, as discussed in Section 1. We also include the limit recently reported by Pasquato et al. (2009) based on a study of the observed mass segregation in the cluster NGC 2298. An IMBH is expected to reduce the amount of mass-segregation that is otherwise expected in a well-relaxed cluster. The suggested IMBH in G1 weighs in at MBH/Mtot = (0.32 ±  0.09)%. The four other clusters in the table only provide upper limits in the range 0.1%–1%. Therefore, the data allow us to probe in a mass range that is of interest in view of existing theories.

Table 1. Globular Clusters with IMBH Constraints

Cluster D V0 M/LV Mtot σe MBH[Tre] MBH[obs] data MBH[obs] / Mtot
  (kpc) (mag) (☉) ( M) (km s−1) ( M) ( M)    
(1) (2) (3) (4) (5) (6) (7) (8) (9) (10)
47 Tuc 4.5a 3.91a 2.4a 1.1 × 106 9.8b 7.3 × 102 <1.5 × 103c discrete PM <0.13%
Omega Cen 4.7d 3.14a 2.6d 2.8 × 106 15.7e 4.8 × 103 <1.2 × 104d discrete PM+LOS <0.43%
M15 10.3f 5.89g 1.6h 6.2 × 105 12.1h 1.7 × 103 <4.4 × 103i discrete PM+LOS <0.71%
NGC 2298 10.7a 9.36a 1.9a 3.4 × 104 2.4b 2.6 × 100 <3.4 × 102j mass segregation <1.0 %
G1 770k 13.49k 2.8l 5.7 × 106 25.1k 3.2 × 104 (1.8 ± 0.5) × 104l unresolved LOS (0.32 ± 0.09)%

Notes. Column 1 lists the cluster name. Column 2 lists the cluster distance D. Column 3 lists the extinction-corrected total apparent magnitude V0. Column 4 lists the V-band mass-to-light ratio M/LV in units of  M/ LV. Column 5 lists the total cluster mass Mtot derived from Columns 2–4. Column 6 lists the effective velocity dispersion σe. Column 7 lists the value of MBH predicted by the MBH vs. σe relationship for galaxies of Tremaine et al. (2002), using the value of σe in Column 6. The uncertainty in the predicted MBH due to the uncertainties in the relation itself is a factor ∼1.5. There is also intrinsic scatter about the relationship of a factor ∼2. Column 8 lists the observational constraint on the mass MBH of any IMBH. Column 9 lists the type of data on which the observational constraint in Column 8 is based. This can be either discrete velocities in the proper motion (PM) and/or LOS directions; unresolved observations of LOS kinematics; or measurements of the amount of mass segregation, which can be compared to model predictions. Column 10 lists the observational constraint on the ratio MBH/Mtot. Sources of the numbers in the table are: aMcLaughlin & van der Marel (2005). bThe central model velocity-dispersion value in McLaughlin & van der Marel (2005), multiplied by the same 85% correction derived in footnote j. cMcLaughlin et al. (2006). dThis paper. eLOS velocity dispersion in an aperture of size equal to the effective radius, calculated from the models in this paper; this is 85% of the central value. fvan den Bosch et al. (2006). gHarris (1996). hGerssen et al. (2002). iRefer to the addendum of Gerssen et al. (2002). jPasquato et al. (2009). kMeylan et al. (2001). lGebhardt et al. (2005).

Download table as:  ASCIITypeset image

The IMBHs suspected in some globular clusters may have been similar to the initial seeds that grew to SMBHs in the centers of galaxies. It is therefore of interest to compare the results in Table 1 to what is known about SMBHs. The masses of SMBHs scale with either the galaxy bulge mass Mbul, or alternatively, with the fourth power of the velocity dispersion σ. Haering & Rix (2004) found that MBH/Mbul ≈ 0.14% ± 0.04%. This is below where it has been possible to probe for most globular clusters. Only for 47 Tuc is there a limit MBH/Mtot ≲ 0.13% (McLaughlin et al. 2006). Tremaine et al. (2002) discussed the MBH–σ relation proposed for SMBHs. In Table 1, we list for each cluster the value of MBH predicted by their relation. All of the available MBH upper limits are at least a factor of 2 above these predictions. Therefore, the data are insufficient to confirm or rule out the hypothesis that globular clusters generally have IMBHs that follow the same relations as established for SMBHs. However, for the case of G1 the suggested IMBH mass does seem reasonably consistent with these relations (Gebhardt et al. 2005).

8. DISCUSSION AND CONCLUSIONS

We have presented a detailed dynamical analysis of the star count, surface brightness, and kinematical data available for ω Centauri, with a particular focus on the new HST data presented in Paper I. Based on the observed profile of the projected density, our models use the Jeans equation to yield predictions for the projected profiles of RMS LOS velocity ${\bar{\sigma }}$ or proper motion ${\bar{\Sigma }}$ as a function of the projected distance R from the cluster center, in each of the three orthogonal coordinate directions (LOS, proper motion radial, and proper motion tangential). We do not include the effect of mass segregation on the model predictions at large radii, but we do take into account the observed (partial) energy equipartition between stars of different masses. The predictions are compared in a χ2 sense to observations, to infer the best-fit values of the model parameters and their error bars.

The model parameters are uniquely determined by the data. The profile β(r) of the intrinsic velocity anisotropy is tightly constrained by the observed profile $[{{\bar{\Sigma }}_{\rm pmt}}/{{\bar{\Sigma }}_{\rm pmr}}](R)$. The models are therefore unaffected by the mass-anisotropy degeneracy that often plagues LOS velocity studies. The mass-to-light ratio ϒ of the stellar population is fixed by normalization of the observed ${\bar{\sigma }}_{\rm los}(R)$ and ${\bar{\Sigma }}_{\rm pm}(R)$ at large radii. The distance D is set by comparison of the observed ${\bar{\Sigma }}_{\rm pm}(R)$ in mas yr−1 to the predicted ${\bar{\sigma }}_{\rm pm}(r)$ in km s−1. And the presence and properties of any dark mass in the center, such as an IMBH or dark cluster, are constrained by the behaviors of ${\bar{\sigma }}_{\rm los}(R)$ and ${\bar{\Sigma }}_{\rm pm}(R)$ at small radii.

The new HST proper motion data provide a significant improvement in our ability to constrain the dynamical models of ω Centauri, as compared to previous modeling efforts (e.g., vdV06; NGB08). The HST sample increases the number of stars with one or more accurately measured velocity components by about an order of magnitude. This allows for the first time the measurement of velocity dispersions at R ≲ 10''. We have derived kinematical profiles down to R ≈ 1'', at which point there cease to be enough stars with measured proper motions to accurately measure a dispersion. The profile of ${\bar{\Sigma }}_{\rm pm}(R)$ is consistent with being flat in the central R ≲ 15''. The absence of an increase in RMS stellar velocities toward the center provides strong constraints on the possible presence of any dark mass.

The HST star count data from Paper I provide an upper limit γ ≲ 0.07 to the central logarithmic cusp of slope of the projected density profile. The innermost data points are well matched by previously published King and Wilson models with a flat core. A generalized nuker profile fits best at all cluster radii when a shallow cusp is invoked, but this overpredicts the innermost data points at R ≲ 4''. Dynamical model predictions depend sensitively on whether or not there is a density cusp, so in our analysis we considered separately models with a core (γ = 0) and models with a cusp (γ = 0.05).

The inferred intrinsic anisotropy profile, distance, and stellar mass-to-light ratio are all insensitive to the exact light and mass distributions near the center. We find that the anisotropy changes from slightly radial near the center to significantly tangential in the outer parts. The presence of velocity anisotropy is not surprising, given the long two-body relaxation time of ω Centauri and the existence of other evidence that the cluster is not yet fully relaxed. The inferred distance is 4.73 ± 0.09 kpc, and the mass-to-light ratio is ϒV = 2.62 ± 0.06. The error bars take into account both the uncertainties in the kinematical data and the uncertainty in the cusp slope γ. These results match extremely well with those obtained by vdV06 using more sophisticated modeling tools. This agreement supports our assertion that spherical models based on the Jeans equation are sufficient for the goals of the present paper, and in particular for constraining the mass distribution near the center. The facts that the ellipticity and rotation of ω Centauri are both known to decrease towards the center provide further motivation for this assertion.

Models with a core provide a good fit to the kinematical data without any dark mass. Since a core is also consistent with the observed density profile, this implies that the presence of an IMBH is not required in ω Centauri. By contrast, models with a shallow cusp provide a good fit only when an IMBH is invoked, with MBH = (8.7 ± 2.9) × 103M. Indeed, an IMBH can induce a shallow density cusp (Baumgardt et al. 2005). Without a dark mass, the predicted RMS velocities in the cusp model decrease toward the cluster center and the predicted velocity distributions become highly non-Gaussian. Neither prediction is consistent with the observations.

Upon combination of the predictions for the core and cusp models, we obtain as a final result an upper limit to the mass of any possible IMBH of 1.2 × 104M at 1σ confidence (or 1.8 × 104M at 3σ confidence). The 1σ limit corresponds to MBH/Mtot ≲ 0.43%. This is of similar magnitude as the range that has now been probed for several other clusters as well. However, lower values of MBH/Mtot will need to be probed to definitively assess the most plausible theories for IMBH formation in globular clusters, or to assess whether globular clusters may follow the same black hole demographics correlations as galaxies.

Isotropic models are not accurate for ω Centauri, since they do not reproduce the observed deviations of $[{{\bar{\Sigma }}_{\rm pmt}}/{{\bar{\Sigma }}_{\rm pmr}}](R)$ from unity. Use of isotropic models to fit ${\bar{\sigma }}(R)$ yields spuriously high IMBH masses. Isotropic models do provide a reasonable guide for calculation of predicted velocity distribution shapes (as opposed to RMS values) in the central region. The predicted Gauss–Hermite moments as a function of the radius for such models fit the observations in the central arcmin to within |Δhi| ≲ 0.01. In models with an IMBH, the wings of the velocity distribution for an aperture on the very center are predicted to be more extended than in models without an IMBH. However, for cusp models with plausible IMBH masses, we would not have expected more than a single star in the data set to have a velocity component in excess of 60 km s−1 (∼3.5 times the cluster velocity dispersion). No such stars were observed in our data. Given the small number statistics, this does not provide meaningful new constraints on the IMBH mass for this particular cluster.

The dark mass invoked in cusp models may in principle be extended. The χ2 of the fit yields a limit adark ≲ 7'' = 0.16 pc on the extent of any dark cluster at 1σ confidence. The corresponding central density is ≳500 times larger than the central density of luminous matter. This would be indicative of quite an extreme amount of mass segregation. If ω Centauri does have a cusp and a dark mass, then an IMBH may be easier to explain theoretically than such a concentrated dark cluster.

The IMBH mass suggested by NGB08 is strongly ruled out in all models. They found MBH = 4.0+0.75−1.0 × 104M from isotropic modeling and MBH = (3.0 ± 1.0) × 104M from anisotropic modeling. By contrast, we find that any IMBH in excess of 1.8 × 104M is ruled out at 3σ or higher confidence. This difference is not surprising, given that our observations have not reproduced the arguments on which their IMBH mass was based (namely, their choice of cluster center, their finding of an increase in dispersion towards that center, and their finding of a cusp around that center). If one ignores these fundamental differences, then maybe our results do not appear so different from those NGB08. When comparing similar models between the two studies (i.e., cusp models with and without anisotropy), the differences in the implied IMBH masses are at the ∼2σ confidence level, given the uncertainties quoted by NGB08 (with our masses being lower). However, the more important difference is that we find that core models with no IMBH are perfectly consistent with the data. By contrast, NGB08 concluded that models with a core and models without an IMBH were ruled out with least 3σ confidence. So while NGB08 argued for a significant IMBH detection, we find by contrast that there is no need to invoke an IMBH in Omega Cen at all. Our finding that ω Centauri cannot have an IMBH as massive as suggested by NGB08 is consistent with arguments presented by Maccarone & Servillat (2008). They found that accretion models with the NGB08 IMBH mass predict much more X-ray and radio emission than is observed from the cluster center.

To further strengthen the conclusions from this paper and our understanding of ω Centauri, it will be useful to proceed with the construction of more sophisticated equilibrium models. Methods for numerical creation of axisymmetric models through orbit superposition have been developed by, e.g., vdV06, van den Bosch et al. (2006), and Chaname et al. (2008). These models can fit the observed differences between major and minor axis kinematics (discussed in Appendix B). They can also make use of the 47,781 high-quality proper motions along the major axis at R ≈ 1–6 arcmin, which were excluded from consideration here (discussed in Appendix C). This will yield constraints on the inclination, rotation properties, and the presence of any kinematic subcomponents. It should also yield further improved limits on the possible presence of any dark mass in the center. Construction of non-equilibrium models for ω Centauri would be useful as well. N-body, Monte Carlo, or Fokker–Planck models might shed light on the evolutionary status of ω Centauri, and on how it arrived at its present structure. This would also provide a more self-consistent description of the joint effects of mass segregation and energy equipartition than we have provided here. Moreover, it would be able to explicitly account for the Brownian motion of a possible IMBH.

We are grateful to Ivan King, Dean McLaughlin, Pat Seitzer, and Glenn van de Ven for useful discussions and/or providing data in electronic format. Suggestions from the referees helped us improve the presentation of our results.

APPENDIX A: EXTRACTING KINEMATICS FROM DISCRETE DATA

To extract kinematics from a set of discrete data points, we select the stars that fall in a given area on the sky (e.g., a circular aperture, annulus, polar wedge, or square aperture). For these stars, we then select the measurements wi ± Δwi in a given coordinate direction (e.g., LOS, proper motion radial, proper motion tangential, proper motion major, or proper motion minor). The wi may be either velocities in km s−1 or proper motions in mas yr−1. The goal is to determine the underlying distribution from which the measurements are drawn. We assume that this is a Gaussian with mean W and dispersion σ, and we set out to determine W and σ with their error bars. If the measurement errors Δwi are all identical then the problem has straightforward analytical solutions. However, this is not generally the case, so that more complicated analysis is required (e.g., Appendix A of vdV06).

We start with a trial estimate of σ. Each measurement wi is then drawn from a Gaussian with dispersion di = (Δw2i + σ2)1/2. Hence, the velocity W and its error bar ΔW follow from straightforward weighted averaging

Equation (A1)

We then determine the value of σ which maximizes the likelihood

Equation (A2)

by numerically solving the equation dL/dσ = 0. Having obtained σ, we return to an improved determination of W and we iterate the procedure until convergence. This provides a fixed point iteration scheme for determination of the joint maximum likelihood solutions for W and σ.

As described in Appendix A of vdV06, the maximum likelihood solution for σ yields a subtly biased estimate of the true dispersion. vdV06 used an analytical approximation to correct for this. Here we have used a Monte Carlo procedure instead. After the determination of (W, σ), we draw pseudodata sets from the corresponding Gaussian probability distribution. To this we add random Gaussian errors drawn from the observational uncertainties. We then analyze numerous such pseudo-data sets in the same fashion as the real data. The statistics of the Monte Carlo results provide both an estimate of the bias in σ (which we use to correct our maximum likelihood estimate) and of the uncertainty Δσ.

After the calculation of W and σ has been completed, the RMS value ${\bar{\sigma }}$ and its uncertainty can be obtained from

Equation (A3)

For the HST proper motion data, our analysis procedure removed all mean motions (see Section 5.2.3). Therefore, we know that the mean of the underlying distribution for any data set extracted from it should be W = 0. For this case, we determined ${\bar{\sigma }}$ more efficiently as the solution of the maximum likelihood equation dL/dσ = 0 at fixed W ≡ 0.

APPENDIX B: MAJOR VERSUS MINOR AXIS

In a spherical system, one expects the projected kinematics to be the same along any axis through the center. One also expects that the projected kinematics for the proper motion coordinate along any fixed direction, when averaged over a circle, will be independent of the chosen direction. These expectations are not realized in practice, and this provides a kinematical signature of the axisymmetry of ω Centauri. vdV06 discussed this in detail for the ground-based kinematical data. Here we briefly discuss this issue for the HST proper motion data.

We have extracted major and minor axis subsamples from our HST proper motion sample, consisting of the stars lying within 10° from either axis. When averaged over all radii R ⩽ 71farcs7, we find for the radial proper motion direction that ${\bar{\Sigma }}_{\rm pmr}= 0.758 \pm 0.010 \;{\rm mas}\;{\rm yr}^{-1}$ along the major axis, and ${\bar{\Sigma }}_{\rm pmr}= 0.739 \pm 0.010 \;{\rm mas}\;{\rm yr}^{-1}$ along the minor axis. For the tangential proper motion direction, we find that ${\bar{\Sigma }}_{\rm pmt}= 0.691 \pm 0.009 \;{\rm mas}\;{\rm yr}^{-1}$ along the major axis, and ${\bar{\Sigma }}_{\rm pmt}= 0.764 \pm 0.010 \;{\rm mas}\;{\rm yr}^{-1}$ along the minor axis. In a spherical system, the kinematics would be the same along the two axes. In reality, ${\bar{\Sigma }}_{\rm pmr}$ is (2.5 ± 1.8)% lower along the minor axis than along the major axis, while ${\bar{\Sigma }}_{\rm pmt}$ is (10.6 ± 2.1)% higher along the minor axis than along the major axis. There is no statistically significant variation with radius in these comparisons (for R ⩽ 71farcs7).

An alternative way to look at these results is to consider the fixed major and minor directions on the sky, and to measure the RMS projected proper motion components ${\bar{\Sigma }}_{\rm pmmj}$ and ${\bar{\Sigma }}_{\rm pmmi}$ along these directions, respectively. For the major axis subsample, we find ${\bar{\Sigma }}_{\rm pmmj}/ {\bar{\Sigma }}_{\rm pmmi}\approx {\bar{\Sigma }}_{\rm pmr}/ {\bar{\Sigma }}_{\rm pmt}= 1.097 \pm 0.021$. For the minor axis subsample, we find ${\bar{\Sigma }}_{\rm pmmj}/ {\bar{\Sigma }}_{\rm pmmi}\approx {\bar{\Sigma }}_{\rm pmt}/ {\bar{\Sigma }}_{\rm pmr}= 1.039 \pm 0.020$. When averaged over the whole sample (i.e., over all position angles) we find that ${\bar{\Sigma }}_{\rm pmmj}/ {\bar{\Sigma }}_{\rm pmmi}= 1.048 \pm 0.007$. Therefore, the RMS projected velocity along the major axis is always larger than that along the minor axis. This was also found by vdV06, who obtained for the 718 stars with all three velocity components measured that ${\bar{\Sigma }}_{\rm pmmj}/ {\bar{\Sigma }}_{\rm pmmi}= 1.055 \pm 0.053$ (see their Appendix C and also their Figure C.1).

These results are relatively easily understood heuristically. For an axisymmetric system that is not too far from edge-on, ${\bar{\Sigma }}_{\rm pmmj}$ is primarily related to the pressure along the equatorial plane and ${\bar{\Sigma }}_{\rm pmmi}$ is primarily related to the pressure along the symmetry axis. The former must exceed the latter in an oblate system, as dictated by the tensor virial theorem (e.g., Binney & Tremaine 1987).

APPENDIX C: PROPER MOTION KINEMATICS FOR THE MAJOR AXIS HST FIELD

As discussed in Section 5.2.1, there is also proper motion data available from Paper I for another field that was observed at multiple epochs with HST. This field lies roughly along the major axis, and therefore does not provide coverage of the full range of position angles in ω Centauri. We therefore did not use it for our model fitting. However, we address here the major axis profiles that can be extracted from these data. This allows a comparison with the central field and our model predictions to check for consistency.

We defined both for the central field and the major axis field a major-axis subsample consisting of stars lying within 10° from the major axis. This is the largest wedge for which coverage over the full range of allowed position angles is guaranteed at the majority of radii R. For the central field, we have complete coverage for R ⩽ 115''. For the major axis field, we have complete coverage for 108'' ⩽ R ⩽ 292''. We extracted the profiles of ${\bar{\Sigma }}_{\rm pmr}(R)$ and ${\bar{\Sigma }}_{\rm pmt}(R)$ for both subsamples, following the procedures described in Section 5.2.2. The results are shown in Figure 11. The HST profiles are continuous at the boundary between the central and major axis fields, as they should be.

Figure 11.

Figure 11. Observed major axis profiles of the RMS projected proper motions ${\bar{\Sigma }}_{\rm pmr}$ and ${\bar{\Sigma }}_{\rm pmt}$ in mas yr−1, as a function of the projected radius R from the cluster center. Symbols are the same as in Figure 3(b). For the HST data points, we included stars within 10° from the major axis. For the vdV06 data points (magenta in the online version, with horizontal error bars), we included stars within a slightly larger angle from the major axis (17fdg5), to increase the sample size. The vdV06 motions were multiplied by a factor 1.023 to correct to the same characteristic main-sequence stellar mass as for the HST sample (see Section 5.1.3). HST data points at R < 110'' (closed dots) are based on the data from the central field, and those at R > 110'' (open dots) are based on the major axis field. The figure focuses on the central region and uses a linear scale in R, by contrast to Figure 3(b). The curves are predictions based on the best-fitting anisotropic model with a core and no dark mass, as discussed in the text. The data for the major axis field are consistent with expectation, given the detailed modeling of the central field and the inability of the HST data to measure absolute rotation in the pmt direction.

Standard image High-resolution image

At radii R ⩽ 71farcs7, we have complete coverage in the central field over all position angles. We can therefore calculate the ratios of the projected RMS proper motions measured within 10° from the major axis and averaged over an entire circle, respectively. This yields $({\bar{\Sigma }}_{\rm pmr})_{\rm maj}/({\bar{\Sigma }}_{\rm pmr})_{\rm circ} = 1.011 \pm 0.014$ and $({\bar{\Sigma }}_{\rm pmt})_{\rm maj}/({\bar{\Sigma }}_{\rm pmt})_{\rm circ} = 0.938 \pm 0.014$. If we assume as a simple approximation that these ratios are independent of radius, then this allows us to compare the predictions of our dynamical models to the observed major axis profiles. The curves in Figure 11 are the model predictions for our best-fitting anisotropic model with a core and no dark mass (see Figure 7 (top)), scaled by the previously quoted ratios $({\bar{\Sigma }}_{\rm pmr})_{\rm maj}/({\bar{\Sigma }}_{\rm pmr})_{\rm circ}$ and $({\bar{\Sigma }}_{\rm pmt})_{\rm maj}/({\bar{\Sigma }}_{\rm pmt})_{\rm circ}$.

The model curves thus obtained provide an excellent fit to the ${\bar{\Sigma }}_{\rm pmr}(R)$ data for the major axis field. Also, the ${\bar{\Sigma }}_{\rm pmr}(R)$ data for the major axis field are consistent with the data from vdV06, which are also shown in the figure. By contrast, the fit to the ${\bar{\Sigma }}_{\rm pmt}(R)$ data for the major axis field is good only for R ≲ 200''. At larger radii, the data fall below the model, and they also fall below the data from vdV06. We attribute this to the inability of the HST proper motion data to measure absolute rotation ${\cal V}_{\rm pmt}$. As a result, we expect to progressively underestimate ${\bar{\Sigma }}_{\rm pmt}$ with increasing radius, since ${\cal V}_{\rm pmt}/\Sigma _{\rm pmt}$ increases with radius (see Section 5.2.3). The fact that the data–model comparison is more successful for ${\bar{\Sigma }}_{\rm pmr}(R)$ is explained by the fact that pmr is orthogonal to direction of absolute rotation, and it is therefore insensitive to it.

Taking into consideration the limitations inherent to the data for the major axis field, and the results presented here, we have no reason to believe that detailed modeling of the data for the major axis field would significantly alter the conclusions of our paper.

Footnotes

  • Based on observations with the NASA/ESA Hubble Space Telescope, obtained at the Space Telescope Science Institute, which is operated by AURA, Inc., under NASA contract NAS 5-26555.

Please wait… references are loading.
10.1088/0004-637X/710/2/1063