1: %\documentclass[12pt,preprint]{aastex}
2: \documentclass{emulateapj}
3: %\documentclass[manuscript]{aastex}
4:
5: %=====================================================================
6: % CUSTOM: PACKAGES, MACROS & SETTINGS
7: %=====================================================================
8:
9: % citation style
10: \usepackage{natbib}
11: % packages for figures
12: \usepackage{graphicx}
13: % packages for symbols
14: \usepackage{latexsym,amssymb}
15: % AMS-LaTeX package for e.g. subequations
16: \usepackage{amsmath}
17:
18: % units
19: \newcommand{\masyr}{mas\,yr$^{-1}$}
20: \newcommand{\kms}{km\,s$^{-1}$}
21: \newcommand{\dgr}{$^\circ$}
22: \newcommand{\Msun}{M$_\odot$}
23: \newcommand{\Lsun}{L$_\odot$}
24: \newcommand{\Lsunpcsq}{L$_\odot$\,pc$^{-2}$}
25: \newcommand{\Msunpcsq}{M$_\odot$\,pc$^{-2}$}
26: \newcommand{\Lsunpccube}{L$_\odot$\,pc$^{-3}$}
27: \newcommand{\Msunpccube}{M$_\odot$\,pc$^{-3}$}
28: \newcommand{\MLsun}{$\Upsilon_\odot$}
29: \newcommand{\MLsunI}{$\Upsilon_{\odot,I}$}
30: \newcommand{\kmsM}{km\,s$^{-1}$\,Mpc$^{-1}$}
31: \newcommand{\Ls}{L$\star$}
32:
33: % abbrevations
34: \newcommand{\Reff}{\ensuremath{R_e}}
35: \newcommand{\Rein}{\ensuremath{R_E}}
36: \newcommand{\Mein}{\ensuremath{M_E}}
37: \newcommand{\Sigc}{\ensuremath{\Sigma_c}}
38: \newcommand{\Rc}{\ensuremath{R_c}}
39: \newcommand{\Mc}{\ensuremath{M_c}}
40: \newcommand{\ML}{\ensuremath{\Upsilon}}
41: \newcommand{\MLein}{\ensuremath{\Upsilon_E}}
42: \newcommand{\MLdyn}{\ensuremath{\Upsilon_\mathrm{dyn}}}
43: \newcommand{\MLstar}{\ensuremath{\Upsilon_\star}}
44: \newcommand{\MLtot}{\ensuremath{\Upsilon_\mathrm{tot}}}
45: \newcommand{\MLj}{\ensuremath{\Upsilon_j}}
46: \newcommand{\Mbh}{\ensuremath{M_\mathrm{BH}}}
47: \newcommand{\du}{\mathrm{d}} % upright d in integrals
48: \newcommand{\plm}{\ensuremath{\,\pm\,}} % +/- in textmode
49: \newcommand{\HI}{\ion{H}{1}}
50: \newcommand{\gmos}{\texttt{GMOS}} % IFS GMOS
51: \newcommand{\sauron}{\texttt{SAURON}} % IFS SAURON
52: \newcommand{\castles}{\texttt{CASTLES}} % Project CASTLES
53: \newcommand{\hst}{HST} % HST
54: \newcommand{\wmap}{WMAP} % WMAP
55: \newcommand{\gemini}{GEMINI} % GEMINI
56: \newcommand{\feh}{\mathrm{[Fe/H]}} % metallicity
57:
58: %=====================================================================
59: % FRONT MATTER
60: %=====================================================================
61:
62: \slugcomment{Submitted to ApJ}
63: \shorttitle{Stellar dynamics and gravitational lensing}
64: \shortauthors{van de Ven et al.}
65:
66: %=====================================================================
67: % BEGIN DOCUMENT
68: %=====================================================================
69:
70: \begin{document}
71:
72: \title{The Einstein Cross: constraint on dark matter from stellar
73: dynamics and gravitational lensing}
74:
75: \author{
76: Glenn van de Ven\altaffilmark{1,2,3},
77: Jes\'us Falc\'on--Barroso\altaffilmark{4,2},
78: Richard M.\ McDermid\altaffilmark{5,2},
79: Michele Cappellari\altaffilmark{6,2},
80: \\
81: Bryan W.\ Miller\altaffilmark{7},
82: P.\ Tim de Zeeuw\altaffilmark{8,2}
83: }
84:
85: \altaffiltext{1}{Institute for Advanced Study, Einstein Drive,
86: Princeton, NJ 08540, USA: glenn@ias.edu}
87:
88: \altaffiltext{2}{Sterrewacht Leiden, Leiden University, Niels Bohrweg
89: 2, 2333 CA Leiden, The Netherlands}
90:
91: \altaffiltext{3}{Hubble Fellow}
92:
93: \altaffiltext{4}{European Space Agency / ESTEC, Keplerlaan 1, 2200 AG
94: Noordwijk, The Netherlands}
95:
96: \altaffiltext{5}{Gemini Observatory, 670 N.\ A'ohoku Place Hilo,
97: Hawaii, 96720, USA}
98:
99: \altaffiltext{6}{Sub-Department of Astrophysics, University of Oxford,
100: Denys Wilkinson Building, Keble Road, Oxford OX1 3RH, United
101: Kingdom}
102:
103: \altaffiltext{7}{Gemini Observatory, Casilla 603, La Serena, Chile}
104:
105: \altaffiltext{8}{European Southern Observatory, Karl-Schwarzschild
106: Strasse 2, 85748 Garching bei M\"unchen, Germany}
107:
108: \begin{abstract}
109: We present two-dimensional line-of-sight stellar kinematics of the
110: lens galaxy in the Einstein Cross, obtained with the \gemini\ 8m
111: telescope, using the \gmos\ integral-field spectrograph. The stellar
112: kinematics extent to a radius of $4$\arcsec\ (with $0.2$\arcsec\
113: spaxels), covering about two-thirds of the effective (or half-light)
114: radius $\Reff \simeq 6$\arcsec\ of this early-type spiral galaxy at
115: redshift $z_l \simeq 0.04$, of which the bulge is lensing a
116: background quasar at redshift $z_s \simeq 1.7$. The velocity map
117: shows regular rotation up to $\sim 100$\,\kms\ around the minor axis
118: of the bulge, consistent with axisymmetry. The velocity dispersion
119: map shows a weak gradient increasing towards a central value of
120: $\sigma_0 = 166 \pm 2$\,\kms.
121: %
122: We deproject the observed surface brightness from \hst\ imaging to
123: obtain a realistic luminosity density of the lens galaxy, which in
124: turn is used to build axisymmetric dynamical models that fit the
125: observed kinematic maps. We also construct a gravitational lens
126: model that accurately fits the positions and relative fluxes of the
127: four quasar images. We combine these independent constraints from
128: stellar dynamics and gravitational lensing to study the total mass
129: distribution in the inner parts of the lens galaxy.
130:
131: We find that the resulting luminous and total mass distribution are
132: nearly identical around the Einstein radius $\Rein = 0.89$\arcsec,
133: with a slope that is close to isothermal, but which becomes
134: shallower towards the center if indeed mass follows light. The
135: dynamical model fits to the observed kinematic maps result in a
136: total mass-to-light ratio $\MLdyn = 3.7 \pm 0.5$\,\MLsunI\ (in the
137: $I$-band). This is consistent with the Einstein mass $\Mein = 1.54
138: \times 10^{10}$\,\Msun\ divided by the (projected) luminosity within
139: $\Rein$, which yields a total mass-to-light ratio of $\MLein =
140: 3.4$\,\MLsunI, with an error of at most a few per cent. We estimate
141: from stellar populations model fits to colors of the lens galaxy a
142: stellar mass-to-light ratio $\MLstar$ from $2.8$ to $4.1$\,\MLsunI.
143: Although a \emph{constant} dark matter fraction of 20 per cent is
144: not excluded, dark matter may play no significant role in the bulge
145: of this $\sim$\Ls\ early-type spiral galaxy.
146: %% This is consistent with indications that less-luminous galaxies
147: %% have no significant dark matter contribution in their inner
148: %% parts.
149: \end{abstract}
150:
151: \keywords{ gravitational lensing --- stellar dynamics --- galaxies:
152: photometry --- galaxies: kinematics and dynamics --- galaxies:
153: structure}
154:
155: %============================= section 1 =============================
156: \section{Introduction}
157: \label{sec:intro}
158: %=====================================================================
159:
160: In the cold dark matter (CDM) paradigm for galaxy formation
161: \citep[e.g.][]{2002SciAm.286..36K}, galaxies are embedded in extended
162: dark matter distributions with a specific and (nearly) universal
163: radial profile \citep[e.g.][]{1997ApJ...490..493N,
164: 2005ApJ...624L..85M}. Measurements of rotation curves from neutral
165: hydrogen (\HI) observations in the outer parts of late-type galaxies
166: have provided evidence for the presence of dark matter in these
167: systems already more than two decades ago
168: \citep[e.g.][]{1985ApJ...295..305V}. In the outer parts of early-type
169: galaxies, however, cold gas is scarce \citep[but see
170: e.g.][]{1994ApJ...436..642F, 1997AJ....113..937M,
171: 2008MNRAS.383.1343W}, so evidence for a dark matter halo has to come
172: from other tracers, such as kinematics of stars, planetary nebulae and
173: globular clusters \citep[e.g.][]{1995ApJ...441L..25C,
174: 2001AJ....121.1936G, 2003Sci...301.1696R, 2003ApJ...591..850C} or
175: hot X-ray gas \citep[e.g.][]{2006ApJ...636..698F}. However, these
176: tracers are not always (sufficiently) available, the observations are
177: often challenging, and the interpretation modeling dependent. Also in
178: the inner parts of galaxies the amount, shape and profile of dark
179: matter is still poorly known \citep[e.g.][]{2004IAUS..220...53P}, even
180: though stellar kinematics are readily available.
181:
182: A fundamental problem in using stellar kinematics (as well as other
183: collisionless kinematic tracers) for this purpose is the
184: mass-anisotropy degeneracy: a change in the measured line-of-sight
185: velocity dispersion can be due to a change in total mass, but also due
186: to a change in velocity anisotropy. Both effects can be disentangled
187: by measuring also the higher-order velocity moments
188: \citep{1987MNRAS.224...13D, 1993ApJ...407..525V, 1993MNRAS.265..213G},
189: but only the inner parts of nearby galaxies are bright enough to
190: obtain the required high-quality kinematic measurements
191: \citep[e.g.][]{1991MNRAS.253..710V, 2001AJ....121.1936G,
192: 2006MNRAS.366.1126C}. A unique alternative to break the
193: mass-anisotropy degeneracy is provided by gravitational lensing. In
194: case of strong lensing, the mass of a foreground galaxy bends the
195: light of a distant bright object behind it, resulting in multiple
196: images. From the separation and fluxes of the images the total mass
197: distribution of the lens galaxy can be inferred directly. The velocity
198: anisotropy then can be determined from the observed velocity
199: dispersion, without the need for higher-order velocity moments.
200:
201: \citeauthor{2004ApJ...611..739T} (\citeyear{2004ApJ...611..739T}, and
202: references therein) have applied this approach to several strong
203: lensing systems, of which 0047-281 \citep{2003ApJ...583..606K} is the
204: best constrained case, with three radially separated velocity
205: dispersion measurements extending to about the effective radius of the
206: lens galaxy. They measure the mass within the Einstein radius by
207: fitting a singular isothermal ellipsoid to the positions of the quasar
208: images. This Einstein mass is used to set the amplitude of the total
209: (stellar and dark) matter distribution, which they assume to be
210: spherical. The constant stellar mass-to-light ratio $\MLstar$
211: determines the contribution of the stars with a fixed radial profile,
212: with the remainder due to dark matter with a single power-law profile
213: with slope $\gamma$. They then compare the dispersion profile
214: predicted by the spherical Jeans equations, for an ad-hoc assumption
215: of the velocity anisotropy $\beta$, with the observed dispersion
216: measurements. Based on a reasonably constrained $\MLstar$ and an upper
217: limit on $\gamma$, they conclude that a significant amount of dark
218: matter is present in the inner parts of the lens galaxy, with a slope
219: shallower than the nearly isothermal total mass distribution. An
220: additional (external) constraint on $\MLstar$ (or $\gamma$) is needed
221: to go beyond this limit on the dark matter distribution. Even so,
222: their results are limited by too few kinematic constraints (which
223: leaves the anisotropy degenerate), and by the use of a simple
224: spherical dynamical model.
225:
226: Clearly, most lens galaxies are significantly flattened and so cannot
227: be well-described by spherical models. Non-spherical models provide a
228: more realistic description of the lens galaxy, but the increase in
229: freedom requires also (significantly) more spatially resolved
230: kinematic measurements to constrain them. Only a few of the known
231: strong gravitational lens systems are close enough to obtain such
232: kinematic data. Even then, one can make (ad-hoc) assumptions on the
233: velocity distribution, such as velocity isotropy in the meridional
234: plane of an axisymmetric model. The latter restriction to a
235: distribution function of two (instead of three) integrals of motion
236: allows for the recovery of the intrinsic shape and mass distribution
237: in the presence of two-dimensional kinematic data
238: \citep{2007ApJ...666..726B, 2008MNRAS.384..987C}, possibly even
239: without the additional constraint provided by gravitational lensing
240: \citep[e.g.][]{2006MNRAS.366.1126C}.
241:
242: In this paper, we relax the two-integral assumption and find that one
243: can still derive independent determinations of the intrinsic mass
244: distribution from both gravitational lensing and stellar dynamics with
245: two-dimensional kinematic data. We have observed the gravitational
246: lens system QSO\,2237+0305, well-known as the Einstein Cross, with the
247: integral-field spectrograph \gmos\ on the \gemini-North Telescope. We
248: fit the extracted stellar velocity and velocity dispersion maps of the
249: inner parts of the lens galaxy at a redshift $z_l \simeq 0.04$ with
250: axisymmetric dynamical models. The resulting intrinsic mass
251: distribution agrees well with that inferred from the lens model that
252: fits the four quasar image positions and relative fluxes. Furthermore,
253: we compare this total mass distribution with the stellar mass
254: distribution, based on an independent estimate of the stellar
255: mass-to-light ratio $\MLstar$, to place constraints on the dark matter
256: distribution in the inner parts of the lens galaxy.
257:
258: In Section~\ref{sec:observations}, we briefly describe the Einstein
259: Cross, and we present the photometric and spectroscopic observations.
260: In Section~\ref{sec:analysis}, we extract the two-dimensional stellar
261: kinematics, describe the dynamical modeling method, and construct
262: lens models. From the latter we infer the total mass distribution,
263: which we compare in Section~\ref{sec:results} with the luminous mass
264: distribution inferred from the observed surface brightness. We then
265: build axisymmetric dynamical models, and compare the resulting
266: dynamical mass-to-light ratio with that inferred from the lens models,
267: and in turn with an estimate of the stellar mass-to-light ratio. In
268: Section~\ref{sec:discconcl} we discuss our findings and we summarize
269: our conclusions.
270:
271: Throughout we adopt the \wmap\ cosmological parameters for the Hubble
272: constant, the matter density and the cosmological constant, of
273: respectively $H_0=73$\,\kmsM, $\Omega_M=0.24$ and $\Omega_L=0.76$
274: \citep{2007ApJS..170..377S}.
275: %% although these parameters only have a small effect on the physical
276: %% scales of the lens galaxy due to its proximity.
277:
278: %============================= section 2 =============================
279: \section{Observations and data reduction}
280: \label{sec:observations}
281: %=====================================================================
282:
283: %---------------------------------------------------------------------
284: \subsection{The Einstein Cross}
285: \label{sec:einsteincross}
286: %---------------------------------------------------------------------
287:
288: The Einstein Cross is the well-known gravitational lens system
289: QSO\,2237+0305 (RA: 22h\,40m\,30.3s, Dec:
290: $+03^\circ$\,21\arcmin\,31\arcsec, J2000). In this system, a distant
291: quasar at redshift $z_s = 1.695$ is lensed by the bulge of the
292: early-type spiral PGC\,069457 at $z_l = 0.0394$ (angular diameter
293: distance $D_l = 155$\,Mpc, $1\arcsec = 0.75$\,kpc), resulting in a
294: cross of four bright images separated by about 1.8\arcsec.
295:
296: The Einstein Cross has long been the closest strong gravitational lens
297: system known, and has been very well studied since its discovery by
298: \cite{1985AJ.....90..691H}. There is a wealth of ground- and
299: space-based imaging data at all wavelengths
300: \citep[e.g.][]{1996AJ....112..897F, 1998MNRAS.298.1223B,
301: 2000ApJ...545..657A, 2003ApJ...589..100D}. The resulting precise
302: measurements of the positions and relative fluxes of the quasar images
303: can be used to construct a detailed lens model.
304:
305: In contrast, kinematic data of the lens galaxy is very scarce, with
306: only one measured central stellar velocity dispersion
307: \citep{1992ApJ...386L..43F} and two \HI\ rotation curve measurements
308: in the very outer parts \citep{1999MNRAS.309..641B}. There are several
309: previous integral-field studies of the Einstein Cross: \texttt{TIGER}:
310: \cite{1994A&A...282...11F}; \texttt{INTEGRAL}:
311: \cite{1998ApJ...503L..27M}; \texttt{CIRPASS}:
312: \cite{2004ApJ...607...43M}. However, none of these studies were
313: concerned with the stellar kinematics of the lens galaxy, but instead
314: investigated the quasar spectra.
315:
316: %---------------------------------------------------------------------
317: \subsection{Imaging}
318: \label{sec:imaging}
319: %---------------------------------------------------------------------
320:
321: %%%FIG
322: \begin{figure*}
323: \includegraphics[width=0.330\textwidth]{f1a.eps}
324: \hfill
325: \includegraphics[width=0.325\textwidth]{f1b.eps}
326: \hfill \includegraphics[width=0.330\textwidth]{f1c.ps}
327: \caption{Surface brightness as observed with \hst\ (left and middle
328: panel) and surface mass density of the overall best-fit lens model
329: (right panel) of the lens galaxy in the Einstein Cross. \emph{Left
330: panel}: the contours of the WFPC2/F555W $V$-band image reveal
331: clearly the bulge, spiral arms and bar embedded in the large-scale
332: disk of this early-type spiral galaxy. The ellipticity measured
333: from the MGE fit (contours) is used to estimate the inclination.
334: \emph{Middle panel}: the central 8\arcsec$\times$8\arcsec\ of the
335: WFPC2/F814W $I$-band image, of which the MGE fit is used to
336: construct the (stellar) luminosity density model of the lens
337: galaxy. We use the $I$-band image instead of the longer exposed
338: $V$-band image as it tracers better the old stellar population and
339: is less sensitive to extinction and reddening. The four quasar
340: images are masked out during the MGE fit. In both WFPC2 images the
341: contours are in steps of $0.5$\,mag/arcsec$^2$. The images are
342: rotated such that North is up and East is to the left.
343: \emph{Right panel}: The scale-free lens model with slope
344: $\alpha=1.0$ (dashed contours) fits the positions and relative
345: fluxes of the quasar images, indicated by the filled circles.
346: Superposed is the MGE fit (solid contours). }
347: \label{fig:mgesurf}
348: \end{figure*}
349: %%%FIG
350:
351: We use two WFPC2 images retrieved from the \hst-archive (F555W
352: $V$-band image, 1600 seconds, PI: Westphal, and F814W $I$-band image,
353: 120 seconds, PI: Kochanek; see left and middle panels of
354: Fig.~\ref{fig:mgesurf}) to determine the luminosity density of the
355: lens galaxy. We correct the $I$-band image for extinction following
356: \cite{1998ApJ...500..525S}, and we convert to solar units using the
357: WFPC2 calibration of \cite{2000PASP..112.1397D}, while assuming an
358: absolute $I$-band magnitude for the Sun of 4.08 mag \citep[Table~2 of
359: ][]{1998gaas.book.....B}. From a de Vaucouleurs $R^{1/4}$ profile fit
360: to the $I$-band photometry in the inner bulge-dominated region, we
361: obtain an effective radius $\Reff \simeq 6$\arcsec, which is consistent
362: with previous measurements \citep[e.g.][]{1991AJ....102..454R}.
363:
364: For the construction of the lens model, we use the accurate positions
365: of the quasar images from the website of the \castles\
366: survey\footnote{http://cfa-www.harvard.edu/castles/} based on
367: \textit{Hubble Space Telescope} (\hst) imaging. Although also optical
368: flux ratios are given on this website, we use the radio fluxes
369: provided by \cite{1996AJ....112..897F}, because they are in general
370: (much) less affected by differential extinction or microlensing.
371:
372: %---------------------------------------------------------------------
373: \subsection{Integral-field spectroscopy}
374: \label{sec:integralfield}
375: %---------------------------------------------------------------------
376:
377: Observations of the Einstein Cross lens system were carried out using
378: the integral-field unit of the \gmos-North spectrograph
379: \citep{2003SPIE.4841.1750M, 2004PASP..116..425H} on July
380: 17$^\mathrm{th}$ and August 1$^\mathrm{st}$ 2005 as part of the
381: program GN-2005A-DD-7. The data were obtained using the IFU two-slit
382: mode that provides a field-of-view of 5\arcsec$\times$7\arcsec. An
383: array of 1500 hexagonal lenslets, of which 500 are located 1\arcmin\
384: away from the main field to be used for sky subtraction, sets the
385: 0\farcs2 spatial sampling. Eight individual science exposures of 1895
386: seconds each were obtained during the two nights, resulting in a total
387: on-source integration time of $\simeq 4$ hours. An offset of 0\farcs3
388: was introduced between exposures to avoid bad CCD regions or lost
389: fibers. The R400-G5305 grating in combination with the CaT-G0309
390: filter was used to cover a wavelength range of 7800--9200\,\AA\ with a
391: FWHM spectral resolution of 2.8\,\AA.
392:
393: For the data reduction we used an updated version of the officially
394: distributed Gemini IRAF\footnote{IRAF is distributed by the National
395: Optical Astronomy Observatories, which are operated by the
396: Association of Universities for Research in Astronomy, Inc., under
397: cooperative agreement with the National Science Foundation.}
398: package. We applied bias subtraction, flat-fielding and cosmic ray
399: rejection \citep[using the \emph{L.A.\ Cosmic} algorithm
400: by][]{2001PASP..113.1420V} to each science exposure, and wavelength
401: calibration after the extraction of the data. A careful flat-fielding
402: procedure, crucial for proper removal of fringes in the spectral
403: direction, was carried out using the Quartz halogen (QH) lamp.
404: Wavelength calibration was performed with CuAr lamp exposures taken
405: before each science exposure. At the observed wavelengths, a complex
406: spectrum of H$_2$O absorption features overlaps with the CaT lines we
407: are interested in. We constructed a correction spectrum from
408: observations of the white dwarf star Wolf1346, taken with the same
409: instrumental setup as our science frames. We checked the range of
410: fiber-to-fiber variations of the spectral resolution of the instrument
411: by measuring the width of the sky lines in each fiber, resulting in
412: the nominal FWHM value of $2.8 \pm 0.2$\,\AA\ for most of the data
413: cubes. In order to combine all the data, we homogenized the science
414: frames by convolving all spectra to an instrumental FWHM resolution of
415: 3.1\,\AA\ (or $\sigma_\mathrm{instr} \simeq 44.2$\,\kms) and
416: resampling the spectra to the same range and sampling in wavelength.
417: After interpolating each science frame to a common spatial grid,
418: taking into account the small spatial offsets applied during the
419: observations, we sum the spectra sharing the same position in the sky
420: to produce the final merged data cube.
421:
422: The significant contribution from the sky lines to the overall
423: spectrum of the lens galaxy made the sky subtraction the most
424: challenging step in the data reduction process. Although the scatter
425: in the instrumental resolution was small within the individual science
426: frames, the combined effect of other steps in the data reduction (e.g.
427: fringing, telluric absorption, resampling of data in wavelength),
428: introduced alterations in the shape of the sky lines (from the sky
429: fibers) that were not equally reproduced in the science fibers. As a
430: consequence any attempt to use the sky from the sky fibers resulted in
431: serious systematic effects (i.e.\ P-Cygni like residuals) in the galaxy
432: spectra underneath. In order to minimize these effects, we chose to
433: extract the sky spectra from the outer most regions of our fully
434: merged science data cube, where no galaxy light was appreciable. The
435: sky was then subtracted optimally as described below.
436:
437: %============================= section 3 =============================
438: \section{Analysis}
439: \label{sec:analysis}
440: %=====================================================================
441:
442: %---------------------------------------------------------------------
443: \subsection{Two-dimensional stellar kinematics}
444: \label{sec:stellarkin}
445: %---------------------------------------------------------------------
446:
447: To measure reliable stellar kinematics we first co-add spectra using
448: the adaptive spatial two dimensional binning scheme of
449: \cite{2003MNRAS.342..345C} to obtain in each resulting (Voronoi) bin a
450: minimum signal-to-noise ratio (S/N). The resulting spectra are
451: spectrally rebinned to steps of constant $\ln\lambda$, equivalent to a
452: sampling of 24\,\kms\ per pixel. We adopt the single stellar
453: population (SSP) models of \cite{2003MNRAS.340.1317V} as spectral
454: templates, which are convolved to have the same effective instrumental
455: broadening as the \gmos\ data. Next, using an updated version (v4.2)
456: of the penalized pixel-fitting (pPXF) algorithm of
457: \cite{2004PASP..116..138C}, a non-negative linear combination of these
458: templates is convolved with a Gaussian line-of-sight velocity
459: distribution (LOSVD) and fit to each observed spectrum, to derive the
460: mean line-of-sight velocity $V$ and velocity dispersion $\sigma$ per
461: bin on the plane of the sky. This version of the code allows for the
462: inclusion of a sky spectrum into the set of stellar templates, and
463: determines the optimal scaling of the sky to the overall spectrum. We
464: used this procedure to {\it clean} the individual spectra in our
465: merged data cube, before we derived the final stellar kinematics.
466:
467: %%%FIG
468: \begin{figure}
469: \begin{center}
470: \includegraphics[width=1.0\columnwidth]{f2.ps}
471: \end{center}
472: \caption{
473: Spectrum from the centerer of the lens galaxy (black curve), showing
474: the Ca II triplet region fitted by a composite of stellar
475: population models (smooth gray curve). The peaks are the sky
476: spectrum.}
477: \label{fig:spectrum}
478: \end{figure}
479: %%%FIG
480:
481: As an example of how this routine performed, we show in
482: Fig.~\ref{fig:spectrum} the spectrum from the center of the lens
483: galaxy, with the best-fit SSP template overplotted. We also plot the
484: removed sky spectrum on top, to show that the sky emission-lines can
485: severely affect the Ca~II triplet absorption lines that we use to
486: extract the stellar kinematics, especially at larger radii where the
487: the relative contribution from the sky is more significant. In
488: particular, in our experiments, a poor sky subtraction caused
489: significant systematic effects in the map of $\sigma$ at the receding
490: side of the galaxy, where one of the absorption lines is redshifted
491: onto (the residual of) a sky emission line. The map of $V$, however,
492: was (nearly) unaffected, showing regular rotation with a straight
493: zero-velocity curve that coincides with the minor axis of the (bulge)
494: photometry. These tests already show that, at least in the inner parts
495: of the lens galaxy, both the photometry and kinematics are consistent
496: with axisymmetry.
497:
498: If we now \emph{assume} this symmetry from the start and apply it
499: during the extracting of the stellar kinematics, we can not only
500: increase the S/N, but at the same time also suppress or even remove
501: systematic effects. To do this, we make use of a built-in
502: functionality in the pPXF algorithm that allows diametrically opposed
503: spectra to be fitted simultaneously. The symmetry imposed by this
504: technique can be either mirror-symmetric or point-symmetric,
505: corresponding to the behavior of the LOSVD in axisymmetric or
506: triaxial stellar systems, respectively. In practice, this is done
507: following the method outlined by \cite{1992MNRAS.254..389R}, whereby
508: the two spectra from opposed positions in the galaxy are laid
509: end-to-end to create a single vector. The template library is also
510: doubled in length, where the second half is convolved by the symmetric
511: counterpart of the broadening function used for the first half (simply
512: reflecting the broadening kernel around the systemic velocity). The
513: two galaxy spectra are therefore fitted simultaneously, with the
514: implied assumption of symmetry.
515:
516: Performing a symmetric extraction of the kinematics presents a number
517: of advantages, and has been used by numerous authors not just as a
518: quality-check \citep[e.g.][]{1992MNRAS.254..389R}, but also when the
519: assumption of symmetry is an integral part of the analysis
520: \citep[e.g.][]{2003ApJ...583...92G}. The two main advantages for the
521: case presented here are in reducing the impact of systematic effects
522: in the data (including sky subtraction) and increasing the effective
523: spatial resolution. The latter is possible since under the assumption
524: of axisymmetry, all moments of the LOSVD are symmetric about the
525: galaxy major axis. We can therefore `fold' the data cube about the
526: major axis, combining spectra from the two hemispheres, which
527: increases the effective S/N within a given spatial element, and
528: therefore reduces the required spatial bin size.
529:
530: We interpolate the merged \gmos\ data cube onto a new spatial grid
531: with $x'$ and $y'$ coordinates parallel to the major and minor
532: photometric axes respectively. We then fold the data-cube about the
533: major axis by summing together spatial pixels (`spaxels') with equal
534: $(x,|y|)$ positions. Before spatially binning the spectra to a minimum
535: S/N, the folded data cube is folded again, into a single quadrant by
536: combining spaxels with equal $(|x|,y)$ positions. The Voronoi bins are
537: derived in this quadrant, and the resulting bin centers are mirrored
538: around $x'=0$. The spectra are combined according to these Voronoi
539: bins to ensure symmetry around the minor axis. At the field perimeter,
540: it is generally not the case that both sides of the galaxy are sampled
541: at the exact same positions, due to the alignment of the \gmos\ field
542: with respect to the galaxy's principle axes. This results in some bins
543: where the S/N of the opposed spectra is far from equal, or in some
544: cases where observations are only present on one side of the galaxy.
545: The former is taken into account by consideration of the error
546: spectrum of each bin. For the latter, the bin is fitted as a single
547: spectrum, giving kinematic values only at that position, without a
548: mirrored counterpart.
549:
550: %%%FIG
551: \begin{figure*}
552: \begin{center}
553: \includegraphics[width=1.0\textwidth]{f3.ps}
554: \end{center}
555: \caption{Mean line-of-sight velocity $V$ and velocity dispersion
556: $\sigma$ maps of the lens galaxy in the Einstein Cross as measured
557: from observations with the integral-field spectrograph \gmos\ on
558: Gemini-North. At the right side of each map, the (linear) scale in
559: \kms\ is indicated by the color bar, and the limits are given
560: below. The third and fourth panel show the $V$ and $\sigma$ maps
561: after symmetrization assuming oblate axisymmetry (see
562: \S~\ref{sec:stellarkin}). The overplotted (black) contours show
563: the isophotes of the image reconstruction from collapsing the
564: unsymmetrized data cube, highlighting the regions excluded in the
565: symmetrized kinematics due to contamination from the quasar
566: images. The $V$ map shows clear and regular rotation around the
567: (vertically aligned) short axis of the bulge. The $\sigma$ map
568: shows a weak gradient decreasing towards the edge of the field.}
569: \label{fig:eckin}
570: \end{figure*}
571: %%%FIG
572:
573: From Monte-Carlo simulations for data of this spectral range and
574: resolution, a minimum S/N of 20 is deemed optimal. This results in an
575: average error in both $V$ and $\sigma$ of about $9$\,\kms, while still
576: preserving the spatial resolution of the data. However, towards the
577: edge of the \gmos\ field, the errors increase to $\sim 15$\,\kms\ in
578: $V$ and $\sim 20$\,\kms\ in $\sigma$. The final symmetrically binned
579: data cube consists of $4 \times 71 = 284$ bins, with the largest
580: containing $19$ individual 0.2\arcsec\ spaxels. The resulting
581: (original and symmetrized) maps of $V$ and $\sigma$ are presented in
582: Fig.~\ref{fig:eckin}, with superposed contours of the integrated flux
583: from the data cube. The velocity map shows regular rotation with
584: amplitude of $V$ up to $\sim 100$\,\kms. The dispersion map shows
585: somewhat higher values of $\sigma$ at the position of the quasar
586: images. This is caused by imperfect subtraction of the quasar
587: continuum, which acts to strongly dilute the absorption lines and make
588: them appear broader.
589:
590: In our stellar template fit to the observed spectra, we include
591: additive polynomials to account for remaining systematic effects in
592: the data, e.g. due to imperfect flat fielding. At the same time, we
593: use these additive components to mimic the contaminating contribution
594: of the quasar images, which impart a strong featureless continuum on
595: top of the galaxy's integrated stellar spectrum. This method nicely
596: removes nearly all contribution from the quasar, except where the
597: quasar images are the brightest, in which case $\sigma$ cannot be
598: reliable measured. We exclude these regions when we fit dynamical
599: models. However, between the quasar images, the spectra are dominated
600: by the stellar light of the galaxy, giving a reliable central
601: dispersion $\sigma_0 \simeq 166 \pm 2$\,\kms. Likewise, at the edges
602: of the field, the quasar contamination is small, providing a reliable
603: gradient in the velocity dispersion.
604:
605: Comparing in Fig.~\ref{fig:eckin} the symmetrized stellar
606: kinematics with what is obtained without imposing axisymmetry, there
607: is a marked reduction of noise and systematic effects, especially in
608: the velocity dispersion map. However, the main features of regular
609: rotation, decreasing $\sigma$ at larger radii, as well as
610: corresponding velocity and dispersion amplitudes, are present in both
611: approaches, albeit with larger uncertainty and systematics in the
612: non-symmetrized case. Given the various advantages, we consider
613: hereafter the symmetrically extracted kinematics.
614:
615: %---------------------------------------------------------------------
616: \subsection{Dynamical models}
617: \label{sec:dynamicalmodel}
618: %---------------------------------------------------------------------
619:
620: As mentioned above, both the photometry and (unsymmetrized) kinematics
621: in the inner parts of the lens galaxy are consistent with axisymmetry.
622: Hence, in constructing dynamical models we consider an axisymmetric
623: stellar system in which both the potential $\Phi(R,z)$ and
624: distribution function (DF) are independent of azimuth $\phi$ and time.
625: By Jeans' (\citeyear{1915MNRAS..76...70J}) theorem the DF only depends
626: on the isolating integrals of motion: $f(E,L_z,I_3)$, with energy $E =
627: (v_R^2+v_\phi^2+v_z^2)/2 + \Phi(R,z)$, angular momentum $L_z =
628: Rv_\phi$ parallel to the symmetry $z$-axis, and a third integral $I_3$
629: for which in general no explicit expression is known. However,
630: usually\footnote{If resonances are present, $I_3$ may loose this
631: symmetry.} $I_3$ is invariant under the change $(v_R,v_z) \to
632: (-v_R,-v_z)$. This implies that the mean velocity is in the azimuthal
633: direction ($\overline{v_R}=\overline{v_z}=0$) and the velocity
634: ellipsoid is aligned with the rotation direction ($\overline{v_R
635: v_\phi} = \overline{v_\phi v_z} = 0$).
636:
637: \cite{1979ApJ...232..236S} introduced a method that sidesteps our
638: ignorance about the non-classical integrals of motion. It finds the
639: set of weights of orbits computed in an arbitrary gravitational
640: potential that best reproduces all available photometric and kinematic
641: data at the same time. The method has proved to be powerful in
642: building detailed spherical and axisymmetric models of nearby galaxies
643: \citep[e.g.][]{1997ApJ...488..702R, 1998ApJ...493..613V,
644: 2003ApJ...583...92G, 2004ApJ...602...66V, 2006MNRAS.366.1126C,
645: 2007MNRAS.382..657T} as well as globular clusters
646: \citep{2006A&A...445..513V, 2006ApJ...641..852V}, and since recently
647: also triaxial models \citep{2008MNRAS.385..614V, 2008MNRAS.385..647V}.
648: However, in all cases the stellar systems are significant closer than
649: the lens galaxy in the Einstein Cross, allowing for (even) more and
650: higher-order stellar kinematic measurements necessary to constrain the
651: large freedom in this general modeling method. Instead we construct
652: simpler, but still realistic dynamical models based on the solution of
653: the axisymmetric Jeans equations.
654:
655: When we multiply the collisionless Boltzmann equation in cylindrical
656: coordinates by respectively $v_R$ and $v_z$ and integrate over all
657: velocities, we obtain the two Jeans equations \citep[see
658: also][eq.~4-29]{BT87}
659: %
660: \begin{eqnarray}
661: \label{eq:cylcbevR}
662: \frac{\partial(R\nu\overline{v_R^2})}{\partial R}
663: + R\frac{\partial(\nu\overline{v_R v_z})}{\partial z}
664: - \nu\overline{v_\phi^2}
665: + R\nu\frac{\partial \Phi}{\partial R}
666: & = & 0,
667: \\
668: \label{eq:cylcbevz}
669: \frac{\partial(R\nu\overline{v_R v_z})}{\partial R}
670: + R\frac{\partial(\nu\overline{v_z^2})}{\partial z}
671: + R \nu \frac{\partial \Phi}{\partial z}
672: & = & 0,
673: \end{eqnarray}
674: %
675: where $\nu(R,z)$ is the intrinsic luminosity density. Due to the
676: assumed axisymmetry, all terms in the third Jeans equation, that
677: follows from multiplying by $v_\phi$, vanish.
678:
679: We are thus left with four unknown second order velocity moments
680: $\overline{v_R^2}$, $\overline{v_z^2}$, $\overline{v_\phi^2}$ and
681: $\overline{v_R v_z}$ and only two equations. This means we have to
682: make assumptions about the velocity anisotropy, or in other words the
683: shape and alignment of the velocity ellipsoid. In case the velocity
684: ellipsoid is aligned with the cylindrical $(R,\phi,z)$ coordinate
685: system $\overline{v_R v_z} = 0$, so that we can readily solve
686: equation~\eqref{eq:cylcbevz} for $\overline{v_z^2}$. If we next assume
687: a constant flattening of the velocity ellipsoid in the meridional
688: plane, we can write $\overline{v_R^2} = \overline{v_z^2}/(1-\beta)$
689: and solve equation~\eqref{eq:cylcbevR} for $\overline{v_\phi^2}$. This
690: assumption provides in general a good description for the kinematics
691: of real disk galaxies \citep{2008arXiv0806.0042C}, such as the lens
692: galaxy in the Einstein Cross, which is an early-type spiral galaxy.
693: When $\beta = 0$, the velocity distribution is isotropic in the
694: meridional plane, corresponding to the well-known case of a
695: two-integral DF $f(E,L_z)$ \citep[e.g.][]{1962MNRAS.123..447L,
696: 1977AJ.....82..271H}.
697:
698: Knowing the intrinsic second-order velocity moments, the line-of-sight
699: second-order velocity moment for a stellar system viewed at an
700: inclination $i > 0$ away from the $z$-axis follows as
701: %
702: \begin{eqnarray}
703: \label{eq:cylvlosmge}
704: \overline{v_\mathrm{los}^2} & = & \frac{1}{I(x',y')}
705: \int_{-\infty}^{+\infty} \nu \biggl[
706: \left(\overline{v_R^2}\sin^2\phi +
707: \overline{v_\phi^2}\cos^2\phi\right) \sin^2i
708: \biggr.
709: \nonumber \\
710: &&
711: \biggl.
712: + \overline{v_z^2} \cos^2i
713: - \overline{v_R v_z}\sin\phi\sin(2i)
714: \biggr] \, \du z',
715: \end{eqnarray}
716: %
717: where $I(x',y')$ is the (observed) surface brightness with the
718: $x'$-axis along the projected major axis. For each position $(x',y')$
719: on the sky-plane, $\overline{v_\mathrm{los}^2}$ yields a prediction of
720: the combination $V^2 + \sigma^2$ of the (observed) mean line-of-sight
721: velocity $V$ and dispersion $\sigma$.
722:
723: Under the above assumptions, besides the anisotropy parameter $\beta$
724: (and possibly the inclination $i$), the only unknown quantity is the
725: gravitational potential, which via Poisson's equation is related to
726: the total mass density $\rho(R,z)$. We may estimate the latter from
727: the intrinsic luminosity density $\nu(R,z)$, derived from deprojecting
728: the observed surface brightness $I(x',y')$, once we know the total
729: mass-to-light ratio $\MLtot$. It is common in dynamical studies of the
730: inner parts of galaxies to consider $\MLtot$ as an additional
731: parameter and to assume its value to be constant, i.e., mass follows
732: light \citep[e.g.][]{2006MNRAS.366.1126C}. Since $\MLtot$ may be
733: larger than the stellar mass-to-light ratio $\MLstar$, this still
734: allows for possible dark matter contribution, but with constant
735: fraction. As an alternative to these 'standard' constant-$\MLtot$
736: models, we also construct dynamical models in which we use the strong
737: gravitational lensing to constrain the gravitational potential. In
738: \S~\ref{sec:lensmodel} below, we use the positions and relative fluxes
739: of the quasar images to estimate the surface mass density
740: $\Sigma(x',y')$, which we deproject to obtain $\rho(R,z)$, so that
741: $\MLtot$ is not anymore a free parameter. For both type of dynamical
742: models, the best-fits to the stellar kinematic data are shown in
743: Fig.~\ref{fig:jeansmodel}, and discussed further in
744: \S~\ref{sec:axijeansmodels} below.
745:
746: We use the Multi-Gaussian Expansion method
747: \citep[MGE;][]{1992A&A...253..366M,1994A&A...285..723E} to
748: parameterize both the observed surface brightness $I(x',y')$ as well
749: as the estimated surface mass density $\Sigma(x',y')$ as a set of
750: two-dimensional Gaussians. Even though Gaussians do not form a
751: complete set of functions, in general surface density distributions
752: are accurately reproduced, including deviations from an elliptical
753: distribution and ellipticity variations with radius. Representing also
754: the point-spread function (PSF) by a sum of Gaussians, the convolution
755: with the PSF becomes straightforward. Moreover, the
756: MGE-parameterization has the advantage that the deprojection can be
757: performed analytically once the viewing angle(s) are given. Also many
758: intrinsic quantities such as the potential can be calculated by means
759: of simple one-dimensional integrals. Similarly, the calculation of
760: $\overline{v_\mathrm{los}^2}$ in equation~\eqref{eq:cylvlosmge}
761: reduces from the (numerical) evaluation of in general a triple
762: integral to a straightforward single integral \citep[see][for further
763: details and a comparison with more general dynamical
764: models]{2008arXiv0806.0042C}. The latter integral is given in
765: Appendix~\ref{sec:appmge}, together with expressions for $\nu(R,z)$
766: and $\Phi(R,z)$ for the MGE-parametrization of $I(x',y')$ and
767: $\Sigma(x',y')$ as given in Table~\ref{tab:mgepar}, and further
768: discussed in \S~\ref{sec:massdistr} below.
769:
770: %---------------------------------------------------------------------
771: \subsection{Lens models}
772: \label{sec:lensmodel}
773: %---------------------------------------------------------------------
774:
775: Under the thin-lens approximation, the gravitational lensing
776: properties of a galaxy are characterized by its potential projected
777: along the line-of-sight, also known as the deflection potential
778: $\phi(x',y')$. In case of a MGE parametrization of the surface mass
779: density of the galaxy, the projection along the line-of-sight of the
780: corresponding gravitational potential, or, alternatively, solving the
781: two-dimensional Poisson equation becomes straightforward. The
782: calculation of $\phi(x',y')$ and its derivatives reduce to the
783: (numerical) evaluation of a single integral, as shown in
784: Appendix~\ref{sec:appmgeaxilens} for an axisymmetric system. This MGE
785: approach thus provides a powerful technique to build general lens
786: models. However, since in this case the quasar images only provide a
787: few constraints within a limited radial range, we adopt instead a
788: simpler and less general approach.
789:
790: We use the algorithm of \cite{2003MNRAS.345.1351E} to
791: construct a lens model that accurately fits the (optical) positions
792: and relative (radio) fluxes of the four quasar images in the Einstein
793: Cross. The deflection potential is assumed to be a scale-free function
794: $\phi(R,\theta) \propto R^\alpha F(\theta)$ of the polar
795: coordinates $R$ and $\theta$ in the lens sky-plane, with $0<\alpha<2$
796: for realistic models. The angular part $F(\theta)$ is expanded as a
797: Fourier series
798: %
799: \begin{equation}
800: \label{eq:eqFtheta}
801: F(\theta) = \frac12\,c_0 +
802: \sum_{m=1}^n [c_m\cos(m\theta) + s_m\sin(m\theta)].
803: \end{equation}
804: %
805: The positions $(x_i,y_i) = (R_i\cos\theta_i,R_i\sin\theta_i)$ of the
806: images are related to the position $(\xi,\eta)$ of the source by the
807: lens equation \citep[e.g.][]{1992grle.book.....S}
808: %
809: \begin{eqnarray}
810: \xi &=& x_i \left[ 1 - \left( \alpha F_i - \tan\theta_i \, F'_i
811: \right) R_i^{\alpha-2} \right],
812: \nonumber \\ \label{eq:lenseq}
813: \eta &=& y_i \left[ 1 - \left( \alpha F_i - \cot\theta_i \, F'_i
814: \right) R_i^{\alpha-2} \right],
815: \end{eqnarray}
816: %
817: where $F_i \equiv F(\theta_i)$ and $F'_i$ denotes the derivative to
818: $\theta_i$. The flux ratios of the images follow from their
819: magnifications $\mu_i$, which are given by
820: %
821: \begin{multline}
822: \label{eq:fluxeq}
823: \left( p_i \mu_i \right)^{-1} =
824: 1 - \left( F_i \alpha^2 + F''_i \right) R_i^{\alpha-2} \\
825: + (\alpha-1) \left[ \alpha^2 F_i^2 -(\alpha-1) {F'_i}^2 + \alpha F_i
826: F''_i \right] R_i^{2(\alpha-2)},
827: \end{multline}
828: %
829: where $p_i$ is the parity of image $i$.
830:
831: Since the Fourier coefficients $c_1$ and $s_1$ correspond to a
832: displacement of the source position $(\xi,\eta)$, we set $c_1=s_1=0$
833: to remove this degeneracy. For a given slope $0<\alpha<2$, we are left
834: with $2n+1$ free parameters $(\xi,\eta,c_2,s_2,\dots,c_n,s_n)$ and
835: eleven constraints (four image positions and three flux ratios). For
836: each image, we obtain a prediction of the source position
837: $(\tilde{\xi}_i,\tilde{\eta}_i)$ and magnification $\tilde{\mu}_i$
838: through equations~\eqref{eq:lenseq} and~\eqref{eq:fluxeq}, respectively. We
839: find the best-fit parameters by minimizing
840: %
841: \begin{equation}
842: \label{eq:minChi2}
843: \chi^2 =
844: \sum_{i=1}^4 \left[
845: \frac{(\xi-\tilde{\xi}_i)^2}{\Delta\xi_i^2} +
846: \frac{(\eta-\tilde{\eta}_i)^2}{\Delta\eta_i^2}
847: \right]
848: + \sum_{j=1}^3
849: \frac{(f_j-\tilde{f_j})^2}{\Delta f_j^2},
850: \end{equation}
851: %
852: with $\Delta\xi_i$ and $\Delta\eta_i$ errors in the observed image
853: positions, and $f_j \equiv \mu_1/\mu_{1+j}$ and $\Delta f_j$ the
854: observed flux ratios and corresponding errors\footnote{For $\alpha=1$
855: (flat rotation curve) equations~\eqref{eq:lenseq}
856: and~\eqref{eq:fluxeq} are linear in the free parameters, so that the
857: solution follows by straightforward matrix inversion
858: \citep{2003MNRAS.345.1351E}.}. During the minimization we guide the
859: solution towards a smooth, realistic lens model by adding to $\chi^2$
860: a constant times $\sum_{m \ge 3}^n (c_m^2+s_m^2)$, which suppresses
861: (strong) deviations from an elliptic shape.
862:
863: %%%FIG
864: \begin{figure*}
865: \begin{center}
866: \includegraphics[width=1.0\textwidth]{f4.ps}
867: \end{center}
868: \caption{Lens model with $n=4$ Fourier terms (see
869: \S~\ref{sec:lensmodel} for details) fitted to the position of the
870: quasar source (first two panels, in milliarcseconds) and to the
871: observed radio flux ratios of the quasar images (third panel).
872: Over the full range of slopes $0.5<\alpha<1.5$, the predicted
873: source positions from each quasar image A through D (as indicated
874: by the different symbols) fall within the observed error (dotted
875: lines). However, the recovery of the observed flux ratios (solid
876: lines) within the observed errors (dotted lines) places
877: constraints on the slope. The minimum reduced $\chi^2$ value of
878: about unity in the fourth panel shows that for $\alpha \simeq 1.0$
879: a good overall best-fit model is obtained, with corresponding
880: $99$\,\%-confidence interval (dotted line) of $0.9 \lesssim \alpha
881: \lesssim 1.1$.}
882: \label{fig:ecfit}
883: \end{figure*}
884: %%%FIG
885:
886: For a range of slopes $0.5<\alpha<1.5$, the fits of the above lens
887: model with $n=4$ Fourier terms are shown in Fig.~\ref{fig:ecfit}.
888: Whereas the predictions of the source position remain within the
889: observed errors over the full range of slopes, the recovery of the
890: observed radio fluxes places constraints on the slope. Dividing the
891: corresponding $\chi^2$ values by the ten degrees of freedom (eleven
892: constraints minus the slope parameter) provides a quality of fit,
893: shown in the lower-right panel of Fig.~\ref{fig:ecfit}. The minimum
894: value $\chi^2/10 \simeq 1$ shows that for $\alpha \simeq 1.0$ a good
895: overall best-fit model is obtained. The corresponding
896: $99$\,\%-confidence interval (dotted line) is $0.9 \lesssim \alpha
897: \lesssim 1.1$.
898:
899: With $n<4$ Fourier terms the fits are significantly worse ($\chi^2/10
900: > 2.5$), and with $n=5$ Fourier terms the fit does not improve (for
901: $n>5$ the fitting problem becomes underdetermined). If we replace the
902: radio fluxes by the optical fluxes from the \castles\ website, no
903: acceptable lens model fit is possible ($\chi^2/10 > 30$). On the
904: other hand, the mid-infrared fluxes from \cite{2000ApJ...545..657A}
905: yield similar good fits as the radio fluxes. The best-fit slope
906: $\alpha=0.9$ is lower because the fluxes of image A and C are
907: respectively about 10\,\% and 30\,\% lower than in the radio. Image C
908: is the faintest in both mid-infrared and radio, while image D is the
909: faintest in the optical. Since the radio is expected to be (much) less
910: affected by differential extinction and microlensing, we adopt the
911: corresponding lens models.
912:
913: %%%FIG
914: \begin{figure*}
915: \begin{center}
916: \includegraphics[width=1.0\textwidth]{f5.ps}
917: \end{center}
918: \caption{
919: Best-fit lens models for five different values of the slope
920: $\alpha$ indicated at the top of each panel (see
921: Fig.~\ref{fig:ecfit} for corresponding quality of the fit). The
922: filled circles indicate the positions of the four quasar images,
923: the cross represents the center of the lens galaxy, and the
924: diamond shows the best-fit position of the quasar source. The
925: smooth and pinched solid curves show respectively the radial
926: critical curve and radial caustic. The dashed curves are the
927: tangential counterparts when they exist (for $\alpha > 1.0$).}
928: \label{fig:ecmod}
929: \end{figure*}
930: %%%FIG
931:
932: %%%TAB
933: \begin{table*}
934: \begin{center}
935: \caption{Lens model parameters\label{tab:ecmod}}
936: \begin{tabular}{c*{13}{c}}
937: \hline
938: \hline
939: $\alpha$ & $\xi$ & $\eta$
940: & $c_0$ & $c_2$ & $c_3$ & $c_4$ & $s_2$ & $s_3$ & $s_4$ & $\Rein$ & $\Mein$ & $\Mc$\\
941: \hline
942: 0.8 & 0.0824 & -0.0151 & 2.1616 & -0.0590 & 0.0001 & 0.0017 & 0.0593 & 0.0022 & 0.0030 & 0.8858 & 1.5414 & 1.6004\\
943: 0.9 & 0.0736 & -0.0143 & 1.9456 & -0.0494 & 0.0001 & 0.0010 & 0.0500 & 0.0005 & 0.0019 & 0.8862 & 1.5425 & 1.5846\\
944: 1.0 & 0.0651 & -0.0133 & 1.7733 & -0.0415 & 0.0002 & 0.0006 & 0.0423 & -0.0006 & 0.0012 & 0.8866 & 1.5441 & 1.5756\\
945: 1.1 & 0.0570 & -0.0121 & 1.6325 & -0.0349 & 0.0004 & 0.0003 & 0.0357 & -0.0014 & 0.0009 & 0.8872 & 1.5461 & 1.5708\\
946: 1.2 & 0.0492 & -0.0107 & 1.5154 & -0.0293 & 0.0005 & 0.0002 & 0.0299 & -0.0018 & 0.0007 & 0.8879 & 1.5485 & 1.5691\\
947: \hline
948: \end{tabular}
949: \tablecomments{Parameters of the lens models shown in
950: Fig.~\ref{fig:ecmod}. For each slope $\alpha$, the best-fit quasar
951: source position $(\xi,\eta)$ and Fourier coefficients $c_m$ and
952: $s_m$ ($c_1=s_0=s_1=0$) are given, all in arcseconds. The last
953: three columns provide the Einstein radius (in arcsec), the projected
954: mass within this radius and within the critical curve (both in
955: $10^{10}$\,\Msun).}
956: \end{center}
957: \end{table*}
958: %%%TAB
959:
960: In the investigation below, we use the lens models with slopes
961: $\alpha=\{0.8,0.9,1.0,1.1,1.2\}$, shown in Fig.~\ref{fig:ecmod}, and
962: with best-fit parameters given in Table~\ref{tab:ecmod}. The critical
963: curves $\Rc(\theta)$ are obtained by solving the
964: (quadratic) equation~\eqref{eq:fluxeq} for infinite magnification,
965: i.e., for vanishing left-hand-side. The caustics then follow upon
966: substitution of $\Rc(\theta)$ in the lens
967: equation~\eqref{eq:lenseq}. We see in Fig.~\ref{fig:ecmod} that even
968: the two models outside the $99$\,\%-confidence interval for $\alpha$
969: are relatively smooth as expected for a realistic galaxy model.
970:
971: The surface mass density of the lens models follows from Poisson's
972: equation as
973: %
974: \begin{equation}
975: \label{eq:surfdens}
976: \Sigma(R,\theta) = \Sigc \,
977: \frac12 \left[ \alpha^2 F(\theta) + F''(\theta) \right] R^{\alpha-2},
978: \end{equation}
979: %
980: with the critical surface mass density defined as
981: %
982: \begin{equation}
983: \label{eq:critsurfdens}
984: \Sigc = \frac{c^2\,D_s}{4\pi G\,D_l\,D_{ls}},
985: \end{equation}
986: %
987: where $c$ is the speed of light and $D_l$, $D_s$ and $D_{ls}$ are the
988: (angular diameter) distance to the lens galaxy, the quasar source and
989: the distance from lens to source, respectively.
990:
991: The (projected) mass within the critical curve
992: %
993: \begin{equation}
994: \label{eq:masscrit}
995: \Mc =
996: %\frac{\Sigc}{2\pi\alpha}
997: \Sigc
998: \int_0^{2\pi} \frac{1}{2\alpha}
999: \left[ \alpha^2 F(\theta) + F''(\theta) \right]
1000: \Rc(\theta)^\alpha \du\,\theta,
1001: \end{equation}
1002: %
1003: is in general close to the projected mass within the Einstein radius.
1004: The latter is the radius $\Rein$ for which the projected mass is equal
1005: to the Einstein mass defined as $\Mein = \Sigc \pi \Rein^2$. Because
1006: $\Rein$ is independent of $\theta$, all higher order Fourier terms
1007: average out, and we are left with $\Rein^{2-\alpha} = \alpha c_0/2$,
1008: so that the Einstein mass becomes
1009: %
1010: \begin{equation}
1011: \label{eq:masseinstein}
1012: \Mein = \Sigc \, \pi \left( \frac{\alpha c_0}{2} \right)^{2/(2-\alpha)},
1013: \end{equation}
1014: %
1015: Since we express all (projected) coordinates on the plane of the sky
1016: in arcseconds, including the radius $R$, we multiply
1017: expressions~\eqref{eq:masscrit} and~\eqref{eq:masseinstein} by $(D_l
1018: \, \pi/0.648)^2$ to convert from arcseconds to pc for a given (angular
1019: diameter) distance $D_l$ to the lens galaxy in Mpc.
1020:
1021: From the last two columns in Table~\ref{tab:ecmod}, we see indeed that
1022: $\Mc \approx \Mein = 1.54 \times 10^{10}$\,\Msun, nearly independent
1023: of the slope $\alpha$. Taking into account an inverse scaling with the
1024: Hubble constant of $H_0 = 73$\,\kmsM, our values are within a few per
1025: cent of previous measurements
1026: \citep[e.g.][]{1992AJ....104..959R,1994AJ....108.1156W,
1027: 1998ApJ...495..609C, 1998MNRAS.295..488S, 2002MNRAS.334..621T}.
1028:
1029: %============================= section 4 =============================
1030: \section{Results}
1031: \label{sec:results}
1032: %=====================================================================
1033:
1034: %% We compare the luminous density inferred from the observed surface
1035: %% brightness with the total mass density derived from the best-fit lens
1036: %% models. We use the corresponding gravitational potentials to built
1037: %% axisymmetric Jeans models which we match to the observed kinematics.
1038: %% The best-fit total mass-to-light ratio we compare with an estimate of
1039: %% the stellar mass-to-light ratio.
1040:
1041: %---------------------------------------------------------------------
1042: \subsection{Luminous versus total mass distribution}
1043: \label{sec:massdistr}
1044: %---------------------------------------------------------------------
1045:
1046: %%%TAB
1047: \begin{table}
1048: \begin{center}
1049: \caption{Multi-Gaussian Expansion parameters\label{tab:mgepar}}
1050: \begin{tabular}{c*{3}{r}c*{3}{r}}
1051: \hline
1052: \hline
1053: & \multicolumn{3}{c}{$I$-band surface brightness} &
1054: & \multicolumn{3}{c}{$\alpha=1.0$ lens model} \\
1055: \cline{2-4} \cline{6-8}
1056: $i$ & $\log\mathrm{I}_0$ & $\log\sigma'$ & $q'$ &
1057: & $\log\Sigma_{0}$ & $\log\sigma'$ & $q'$ \\
1058: % & (\Lsunpcsq) & (\arcsec) &
1059: % & (\Msunpcsq) & (\arcsec) & \\
1060: \hline
1061: 1 & 4.329 & -1.564 & 0.700 & & 5.114 & -1.398 & 0.645\\
1062: 2 & 3.935 & -0.942 & 0.700 & & 4.556 & -1.037 & 0.660\\
1063: 3 & 3.606 & -0.627 & 0.700 & & 4.238 & -0.779 & 0.662\\
1064: 4 & 3.293 & -0.239 & 0.700 & & 3.928 & -0.544 & 0.678\\
1065: 5 & 3.005 & -0.043 & 0.700 & & 3.628 & -0.350 & 0.673\\
1066: 6 & 2.845 & 0.230 & 0.700 & & 3.459 & -0.186 & 0.675\\
1067: 7 & 2.261 & 0.587 & 0.700 & & 3.354 & -0.001 & 0.675\\
1068: 8 & 2.160 & 0.917 & 0.414 & & 3.208 & 0.220 & 0.675\\
1069: 9 & 1.334 & 1.135 & 0.700 & & 3.069 & 0.496 & 0.675\\
1070: 10 & - & - & - & & 2.991 & 1.000 & 0.675\\
1071: \hline
1072: \end{tabular}
1073: \tablecomments{The parameters of the Gaussians in the MGE fit to the
1074: \hst/WFPC2/F814W $I$-band image of the surface brightness (columns
1075: 2--4, ``deconvolved'' with the WFPC2 PSF), and to the surface mass
1076: density of the overall best-fit lens model with slope $\alpha = 1.0$
1077: (columns 5--7) of the lens galaxy in the Einstein Cross (see also
1078: Fig.~\ref{fig:mgesurf}). Columns two and five give for each
1079: Gaussian component respectively the central surface mass density (in
1080: \Msunpcsq) and the central surface brightness (in \Lsunpcsq),
1081: columns three and six the dispersion (in arcsec) along the major
1082: axis, and columns four and seven the observed flattening.}
1083: \end{center}
1084: \end{table}
1085: %%%TAB
1086:
1087: In the left and middle panel of Fig.~\ref{fig:mgesurf} we show MGE
1088: fits (contours) to respectively the $V$-band and the $I$-band \hst\
1089: image, obtained with the software of \cite{2002MNRAS.333..400C}, while
1090: masking the quasar images. For each fit we assume the same position
1091: angle (PA) for the Gaussians, consistent with the adopted axisymmetry
1092: in the dynamical models. Although the bar and spirals cannot be
1093: reproduced, it provides a good description of the disk in the outer
1094: parts (left panel) and reproduces well the bulge in the inner parts
1095: (middle panel). Similarly, we derive MGE fits to the surface mass
1096: density in equation~\eqref{eq:surfdens} of the best-fit lens models
1097: with slopes $\alpha$ from 0.8 to 1.2. In the right panel of
1098: Fig.~\ref{fig:mgesurf}, we show the MGE fit (solid contours) to the
1099: surface mass density (dashed contours) of the overall best-fit lens
1100: model with slope $\alpha = 1.0$. The corresponding parameters are
1101: given in Table~\ref{tab:mgepar} (columns 5--7), together with the
1102: parameters of the MGE fit to the $I$-band surface brightness (columns
1103: 2--4). The latter parameters are ``deconvolved'' using a MGE model
1104: fitted to the WFPC2 PSF as determined with Tiny Tim \citep{TinyTim}.
1105:
1106: The Gaussians in the MGE fit to the $I$-band surface brightness have a
1107: PA of $\sim 70$\dgr, which is similar to the PA of $\simeq 67$\dgr\ in
1108: the MGE fit to the surface mass density of lens models. Both these
1109: values are consistent with the measurements by
1110: \cite{1988AJ.....95.1331Y}, who found a PA of $\simeq 67$\dgr\ for the
1111: axis through quasar images C and D, bracketed by a PA of $\simeq
1112: 77$\dgr\ for the outer disk and a PA of $\simeq 39$\dgr\ for the bar
1113: \citep[see also Fig.~1 of][]{2002MNRAS.334..621T}.
1114:
1115: %%%FIG
1116: \begin{figure*}
1117: \begin{center}
1118: \includegraphics[width=1.0\textwidth]{f6.ps}
1119: \end{center}
1120: \caption{
1121: Radial profiles of the mass distribution of the lens galaxy in the
1122: Einstein Cross as function of radius $R$, in arcsec at the bottom
1123: and in kpc in at the top. In each panel the dashed and solid
1124: vertical lines indicate respectively the Einstein radius $\Rein$ and
1125: the effective radius $\Rein$. The \emph{top panels} shows from left
1126: to right the surface mass density profile $\Sigma$, the
1127: corresponding projected ellipticity $\epsilon = 1 - q'$, and the
1128: mass $M$ enclosed within the projected radius $R$ along the
1129: major-axis on the plane of the sky. The dotted, dash-dotted and
1130: dashed curves are from the best-fit lens models with (fixed)
1131: slopes $\alpha$ of respectively $0.9$, $1.0$ and $1.1$ (see
1132: \S~\ref{sec:lensmodel}). The thick solid curves are from the
1133: observed ($I$-band) surface brightness, multiplied with a constant
1134: mass-to-light ratio $\MLein = 3.4$\,\MLsunI. The latter is
1135: obtained from dividing the Einstein mass (indicated by the
1136: horizontal dashed line in the top-right panel) by the (projected)
1137: luminosity within the Einstein radius. The \emph{bottom panels}
1138: shows from left to right the mass density profile $\rho$, the
1139: corresponding intrinsic ellipticity $\epsilon_\mathrm{intr} = 1 -
1140: q$, and the circular velocity $v_\mathrm{circ}$ as function of the
1141: radius $R$ in the meridional plane. These intrinsic quantities are
1142: obtained from the MGE fits to the projected density, under the
1143: assumption of oblate axisymmetry and for an inclination
1144: $i=68$\dgr\ (see also Appendix~\ref{sec:appmge}). }
1145: \label{fig:ecmassdistr}
1146: \end{figure*}
1147: %%%FIG
1148:
1149: As shown in Fig.~\ref{fig:ecmassdistr}, not only the orientation, but
1150: also the shape of the luminous density distribution as inferred from
1151: the surface brightness, is very similar to that of the total mass
1152: density distribution as derived from the lens models. The left and
1153: middle panel in the top row show respectively the profile and the
1154: projected ellipticity $\epsilon = 1 - q'$ of the surface mass density
1155: as function of the major axis on the sky-plane. The thin curves are
1156: for the lens models with fixed slope $\alpha$ (as indicated in the
1157: bottom-right panel). The thick solid curve is for the $I$-band surface
1158: brightness, multiplied with a constant mass-to-light ratio $\MLein =
1159: 3.4$\,\MLsunI. The latter is derived from dividing the projected mass
1160: from the lens models (shown in the top-right panel of
1161: Fig.~\ref{fig:ecmassdistr}) within the Einstein radius, i.e., the
1162: Einstein mass $\Mein$, by the projected luminosity within the Einstein
1163: radius $\Rein$.
1164:
1165: We see that around $\Rein$ (indicated by the dashed vertical line),
1166: where the lens models are best constrained by the observed quasar
1167: image positions and relative flux ratios, both the slope and the
1168: ellipticity of the \emph{projected} luminous and total mass density
1169: are nearly the same. The same holds true for the corresponding
1170: \emph{intrinsic} mass densities, of which the profiles and
1171: ellipticities $\epsilon_\mathrm{intr} = 1 - q$ are shown in the left
1172: and middle panel in the bottom row of Fig.~\ref{fig:ecmassdistr}.
1173: These (analytic) deprojections (see Appendix~\ref{sec:appmge}) are
1174: under the assumption of oblate axisymmetry and for a given inclination
1175: which we derive from the flattening of the disk in the outer parts. We
1176: find from the MGE fit to the $V$-band surface brightness (left panel
1177: of Fig.~\ref{fig:mgesurf}), that Gaussian components with a measured
1178: flattening as small as $q' \simeq 0.4$ are required for an acceptable
1179: fit. This sets a lower limit to the inclination of $i \gtrsim 66$\dgr
1180: \citep[significantly larger than $i \simeq 60$\dgr\ found
1181: by][]{1989AJ.....98.1989I}. Adopting a lower limit for the
1182: \textit{intrinsic} flattening of the disk of $q = 0.15$
1183: \citep[e.g.][]{1992MNRAS.258..404L}, we then obtain an inclination of
1184: $i \simeq 68$\dgr.
1185:
1186: Around $\Rein$ the luminosity density is close to isothermal ($\rho
1187: \propto R^{-2}$ intrinsic or $\Sigma \propto R^{-1}$ in projection)
1188: like the mass density of the overall best-fit lens model. Towards the
1189: center the slope of the luminosity density becomes shallower. The
1190: (fixed) slope of the lens model is not anymore (well) constrained by
1191: the quasar images towards the center, neither for radii a few times
1192: $\Rein$. At these larger radii also the luminosity density is not only
1193: due to the bulge, but the bar and disk start contributing, as can also
1194: be seen from the increase in the ellipticity. Nevertheless, within a
1195: radius $R \lesssim 4$\arcsec, the intrinsic luminous and total mass
1196: distribution are very similar, i.e., mass follows light. This means we
1197: can infer the gravitational potential either from the observed surface
1198: brightness adopting a (constant) total mass-to-light ratio $\MLtot$,
1199: or directly from the best-fit lens model (see also
1200: Appendix~\ref{sec:appmge}). The corresponding circular velocity
1201: curves, defined as $v_\mathrm{circ}^2 = R \du\Phi/\du R$ in the
1202: equatorial plane, are shown in the bottom-right panel of
1203: Fig.~\ref{fig:ecmassdistr}, using $\MLein = 3.4$\,\MLsunI\ as above.
1204:
1205: %---------------------------------------------------------------------
1206: \subsection{Axisymmetric Jeans models}
1207: \label{sec:axijeansmodels}
1208: %---------------------------------------------------------------------
1209:
1210: %%%FIG
1211: \begin{figure*}
1212: \begin{center}
1213: \includegraphics[width=1.0\textwidth]{f7.ps}
1214: \end{center}
1215: \caption{Axisymmetric Jeans model of the lens galaxy in the
1216: Einstein Cross. The second panel in the bottom row shows the
1217: square root of the combination $V^2+\sigma^2$ of the observed mean
1218: line-of-sight velocity $V$ and dispersion $\sigma$ maps
1219: (Fig.~\ref{fig:eckin}). The (linear) scale in \kms\ is indicated
1220: by the color bar and limits at the top, and is the same for all
1221: other panels, which show the line-of-sight second-order velocity
1222: moment as predicted by the axisymmetric Jeans models
1223: (\S~\ref{sec:dynamicalmodel} and
1224: Appendix~\ref{sec:appmgeaxijeans}). In the \emph{top row}, the
1225: gravitational potential is inferred from the $I$-band surface
1226: brightness, assuming that mass follows light. The corresponding
1227: (constant) dynamical mass-to-light ratio $\MLdyn$ that provides
1228: the best-fit to the observations is indicated at the top, together
1229: with the assumed value for the anisotropy parameter $\beta$,
1230: increasing from left to right. The latter parameter represents the
1231: (constant) flattening of the velocity ellipsoid in the
1232: meridional plane. The \emph{middle row} follows the same
1233: sequence of increasing $\beta$ values, but the gravitational
1234: potential is now computed directly, i.e., without mass-to-light
1235: ratio conversion, from the overall best-fit lens model with slope
1236: $\alpha = 1.0$. In the \emph{bottom row}, besides the
1237: observations in the second panel, the predictions for different
1238: slopes $\alpha$ are shown, with $\beta=0$, corresponding to an
1239: isotropic velocity distribution in the meridional plane.}
1240: \label{fig:jeansmodel}
1241: \end{figure*}
1242: %%%FIG
1243:
1244: In Fig.~\ref{fig:jeansmodel}, the second panel in the bottom row shows
1245: the square root of the combination $V^2+\sigma^2$ of the observed (and
1246: symmetrized) mean line-of-sight velocity $V$ and dispersion $\sigma$
1247: maps of the lens galaxy (Fig.~\ref{fig:eckin}, two panels on the
1248: right). The other panels in Fig.~\ref{fig:jeansmodel} show the
1249: line-of-sight second-order velocity moment as predicted by the
1250: axisymmetric Jeans models, using the solution given in
1251: equation~\eqref{eq:cylvlos2mge} of Appendix~\ref{sec:appmge}. With a
1252: central velocity dispersion of $\sigma_0 \simeq 166$\,\kms, the
1253: $\Mbh-\sigma$ relation as given by \cite{2002ApJ...574..740T} predicts
1254: a central black hole of mass $\Mbh \simeq 6.4 \times 10^7$\,\Msun\ for
1255: the lens galaxy. Since the corresponding sphere-of-influence
1256: $\Mbh/\sigma^2 \simeq 0.01$\arcsec\ is not resolved by our kinematic
1257: data, we cannot fit for it, but do include the predicted BH in the
1258: Jeans models (as an additional Gaussian). Finally, we convolve the
1259: predicted second-order velocity moment with the Gaussian PSF of the
1260: kinematic data with FWHM$=1$\arcsec. The (linear) scale in \kms, as
1261: indicated by the color bar and limits at the top of the panel with
1262: the observations, is the same in all panels of
1263: Fig.~\ref{fig:jeansmodel}.
1264:
1265: The axisymmetric Jeans models in the top row use the gravitational
1266: potential inferred from the $I$-band surface brightness, assuming that
1267: mass follows light. The corresponding (constant) dynamical
1268: mass-to-light ratio $\MLdyn$ that provides the best-fit to the
1269: observations is indicated at the top, together with the assumed value
1270: for the anisotropy parameter $\beta$, increasing from left to right.
1271: The corresponding increase in the (constant) flattening of the
1272: velocity ellipsoid in the meridional plane (or decrease in
1273: $\overline{v_z^2}/\overline{v_R^2}$), has two main effects: a less
1274: pinched ``butterfly'' shape and a weaker gradient parallel to the
1275: (vertically aligned) short axis. Comparing with the observations, we
1276: see that at lower $\beta$ values the pinching is too strong, while at
1277: higher $\beta$ the gradient along the short axis is too shallow.
1278: Therefore, it is not unexpected that the ``best-fit'' is when $\beta
1279: \simeq 0$ and $\MLdyn \simeq 4.0$\,\MLsunI\ (in the $I$-band), but
1280: with reduced $\chi^2 \simeq 3$ significantly above unity.
1281:
1282: Part of the latter mismatch might be due to (systematic) discrepancies
1283: in the data, in particular due to the challenging measurement of the
1284: velocity dispersion. The average uncertainty in the observed
1285: second-order velocity moment is about $10$\,\kms, which translates
1286: into an error of around $12$ per cent in $\MLdyn$, or about
1287: $0.5$\,\MLsunI. However, the uncertainties in the observed
1288: second-order velocity moment vary quite a lot, from only $\sim
1289: 2$\,\kms\ in the center to $\sim 20$\,\kms\ a the edge of the \gmos\
1290: field. And although we were very careful in disentangling the
1291: contribution from the quasar and excluded the regions at the quasar
1292: images, the measured velocity dispersion around these regions might
1293: still be systematically affected by (broadening due to) residual light
1294: from the quasar source. This possibly explains the observed excess in
1295: the observed second-order velocity moment around the quasar images
1296: with respect to the accurately measured central value. At the same
1297: time, such an excess might mimic a steeper gradient in the observed
1298: second-order velocity moment. Indeed, in all Jeans models in the top
1299: row of Fig.~\ref{fig:jeansmodel} the prediction at the center is
1300: significantly higher than the observed value. As a result, we expect
1301: that the dynamical mass-to-light ratio is overestimated.
1302:
1303: If, alternatively, we infer the gravitational potential directly,
1304: i.e., without the need for a mass-to-light ratio, from the
1305: (deprojected) surface mass density of the best-fit lens models, the
1306: axisymmetric Jeans models predict maps of the second-order velocity
1307: moments that are similar in shape but with lower values near the
1308: center. This is shown in the middle row of Fig.~\ref{fig:jeansmodel}
1309: for the overall best-fit lens model with slope $\alpha=1$, and for the
1310: same range in anisotropy parameters $\beta$. The formal best-fit is
1311: obtained for $\beta \simeq 0$, but the observed central value is more
1312: closely matched in case of mild velocity anisotropy in the meridional
1313: plane with $\beta \gtrsim 0.1$. For $\alpha=0.9$. i.e., density
1314: slopes steeper then isothermal, the observed central value is best
1315: matched when $\beta \simeq 0$, with at the same time a steeper
1316: gradient along the short-axis, as shown in the first panel in the
1317: bottom row of Fig.~\ref{fig:jeansmodel}. The last two panels show that
1318: for $\alpha>1$, velocity isotropy in the meridional plane ($\beta=0$)
1319: results in a central value much lower than the observed second-order
1320: velocity moment.
1321:
1322: We conclude that for the lens models with slopes $\alpha = 1.0 \pm
1323: 0.1$, as constrained by the lensing geometry, and for a mild
1324: anisotropic velocity distribution in the meridional plane with $\beta
1325: = 0.1 \pm 0.1$, the predictions of the second-order velocity moments
1326: match the accurately measured central value, and follow the gradient
1327: within the observational uncertainties outside the region that might
1328: be affected by (residual) quasar light. The latter results in
1329: broadening of the velocity dispersion, so that the best-fit dynamical
1330: mass-to-light ratios we derive when we infer the gravitational
1331: potential from the surface brightness, are likely overestimated.
1332: Fitting instead only the unaffected and accurately measured
1333: second-order velocity moment in the center, we find for $\beta=0.1$ a
1334: best-fit value of $\MLdyn \simeq 3.7$\,\MLsunI, with an error around
1335: $0.5$\,\MLsunI. On the other hand, both the Einstein mass and
1336: luminosity within the Einstein radius are well constrained, resulting
1337: in at most a few per cent error in $\MLein = 3.4$\,\MLsunI.
1338: Therefore, in what follows, we adopt the latter as the total
1339: mass-to-light ratio in the inner bulge-dominated region of the lens
1340: galaxy.
1341:
1342: In the above axisymmetric Jeans models we fix the inclination to $i =
1343: 68$\dgr\ determined from the observed ellipticity of the outer disk
1344: (see \S~\ref{sec:massdistr}). When we vary the inclination over a
1345: reasonable range above the lower limit $i \gtrsim 66$\dgr, there is no
1346: significant change in the above results, which is expected because the
1347: inclination only has a weak effect on the mass-to-light ratio
1348: \citep[see also Fig.~4 of][]{2006MNRAS.366.1126C}.
1349:
1350: %---------------------------------------------------------------------
1351: \subsection{Stellar mass-to-light ratio}
1352: \label{sec:stellarml}
1353: %---------------------------------------------------------------------
1354:
1355: We have shown that in the bulge-dominated inner region of the lens
1356: galaxy, the total mass distribution closely follows the light
1357: distribution. This does not mean there is no dark matter present,
1358: since if dark matter follows (nearly) the same distribution it might
1359: contribute to part of the above estimated total mass-to-light ratio
1360: $\MLtot \simeq 3.4$\,\MLsunI. To determine this possible
1361: \emph{constant} dark matter fraction, we need to measure the stellar
1362: mass-to-light ratio $\MLstar$. To this end we use the single stellar
1363: population (SSP) models of \cite{1996ApJS..106..307V}, adopting as
1364: initial reference the \cite{2001MNRAS.322..231K} initial mass function
1365: (IMF) with lower mass cut-off $0.01$\,\Msun (faintest star
1366: $0.09$\,\Msun) and upper mass cut-off of $120$\,\Msun.
1367:
1368: We derive a high-S/N spectrum from our data cube by collapsing the
1369: central spectra that are unaffected by the quasar images. Even so, the
1370: Ca~II triplet is equally well fitted by SSP models of nearly all ages
1371: and metallicities of about $\feh = \pm 0.2$ around solar, which leaves
1372: the stellar mass-to-light ratio nearly unconstrained. Therefore, we
1373: also derived colors from archive HST images (PI: Kochanek).
1374: % Proposal ID = 8252
1375: After matching all images, using both the quasar positions and
1376: isophotes of the lens galaxy, we compute the magnitudes in a circular
1377: aperture of radius $R = 0.2$\arcsec\ (well within the quasar images at
1378: $R \simeq 0.9$\arcsec). Next, we convert the resulting magnitudes in
1379: the F555W, F675W, F814W (WFPC2), F160W and F205W (NICMOS2) filters to
1380: (rest frame) magnitudes in respectively the Johnson $V$, $I$, $R$, $H$
1381: and $K$ filters. We use the method described in \S~3 of
1382: \cite{2003MNRAS.344..924V} to take simultaneously into account the
1383: filter responses and the redshift of the lens galaxy.
1384:
1385: The resulting colors are best fitted by a SSP model with age $t \simeq
1386: 8$\,Gyr and metallicity $\feh \simeq 0.2$, corresponding to an
1387: $I$-band stellar mass-to-light ratio $\MLstar \simeq 3.3$\,\MLsunI.
1388: The $1\sigma$ confidence limits yield a range in $t$ from 7 to
1389: 14\,Gyr, and in $\feh$ from $0.0$ to $0.3$, corresponding to a range
1390: in $\MLstar$ from $2.8$ to $4.1$\,\MLsunI. The lower limit implies at
1391: most $\sim 20$ per cent of dark matter within $R \la 4$\arcsec, but
1392: the results are fully consistent with no dark matter at all. The SSP
1393: models of \cite{1998MNRAS.300..872M, 2005MNRAS.362..799M} with a
1394: Kroupa IMF yield similar results. Adopting a
1395: \cite{1955ApJ...121..161S} IMF with the same lower and upper mass
1396: cut-offs, increases $\MLstar$ by about $30$ per cent. This not only
1397: implies no dark matter, but even for the lower limit $\MLstar >
1398: \MLtot$, which is unphysical. This is consistent with evidences for
1399: both late-type galaxies \citep{2001ApJ...550..212B} and early-type
1400: galaxies \citep{2006MNRAS.366.1126C} that, if the IMF is universal, a
1401: Salpeter IMF is excluded, whereas a Kroupa IMF matches the
1402: observations \citep[see][for a review]{2007iuse.book..107D}.
1403:
1404: %============================= section 5 =============================
1405: \section{Discussion and conclusions}
1406: \label{sec:discconcl}
1407: %=====================================================================
1408:
1409: We used the \gmos-North integral-field spectrograph to obtain
1410: two-dimensional stellar kinematics of the lens galaxy in the Einstein
1411: Cross. In addition to the four bright quasar images and the distance of
1412: the lens galaxy ($D_l = 155$\,Mpc), in particular the presence of sky
1413: lines in the observed Ca II triplet region made the extraction of the
1414: absorption line kinematics challenging. Even so, we were able to derive
1415: high-quality line-of-sight velocity $V$ and dispersion $\sigma$ maps
1416: of the bulge-dominated inner region $R \lesssim 4$\arcsec, reaching
1417: about two-thirds of the effective radius $\Reff \simeq 6$\arcsec\ of
1418: this early-type spiral galaxy. The $V$ map shows regular rotation up
1419: to $\sim 100$\,\kms\ around the minor axis of the bulge, consistent
1420: with axisymmetry. The $\sigma$ map shows a weak gradient increasing
1421: towards a central value of $166 \pm 2$\,\kms.
1422:
1423: The only other direct (single) measurement of the velocity dispersion
1424: is $215 \pm 30$\,\kms\ by \cite{1992ApJ...386L..43F}. For a singular
1425: isothermal sphere lens model, we can use the relation
1426: $\Delta\theta=8\pi(\sigma_\mathrm{SIS}/c)^2D_{ls}/D_s$
1427: \citep[e.g.][]{2000ApJ...543..131K} with a separation $\Delta\theta =
1428: 1.8$\arcsec\ of the four quasar images, to obtain a simple estimate
1429: for the dispersion of $\sigma_\mathrm{SIS} \simeq 180$\,\kms. Taking
1430: into account aperture correction and a range in velocity anisotropy,
1431: \cite{2003MNRAS.344..924V} convert this to a central velocity
1432: dispersion of $168 \pm 17$\,\kms. The King and de Vaucouleurs models
1433: of \cite{1988AJ.....96.1570K} predict a similar value of $\sim
1434: 166$\,\kms, and also \cite{1999MNRAS.309..641B} find a value of $165
1435: \pm 23$\,\kms\ based on their two \HI\ rotation curve measurements.
1436: It is likely that the long-slit measurement by
1437: \cite{1992ApJ...386L..43F} is affected by the bright quasar images,
1438: whereas our spatially resolved measurements allow for a clean
1439: separation of the quasar contribution.
1440:
1441: A large variety of different lens models have been constructed for the
1442: Einstein Cross, most of which fit the positions of the quasar images
1443: but not their relative flux ratios. Adopting the scale-free lens model
1444: of \cite{2003MNRAS.345.1351E}, we found that fitting at the same time
1445: also the (radio) flux ratios constrained the slope of the total mass
1446: surface density $\Sigma \propto R^{\alpha-2}$ to be $\alpha = 1.0 \pm
1447: 0.1$. The total mass within the Einstein radius $\Rein = 0.89$\arcsec,
1448: i.e., the Einstein mass, is $\Mein = 1.54 \times 10^{10}$\,\Msun,
1449: nearly independent of the slope $\alpha$, and consistent with previous
1450: measurements in the literature. Dividing by the projected luminosity
1451: within $\Rein$, as measured from the observed $I$-band surface
1452: brightness, we obtained a mass-to-light ratio of $\MLein =
1453: 3.4$\,\MLsunI, with an error of at most a few per cent.
1454:
1455: We determined the mass-to-light ratio in an additional, independent
1456: way by fitting dynamical models to the (combined) observed $V$ and
1457: $\sigma$ maps. We used the solution of the axisymmetric Jeans
1458: equations to predict this second-order velocity moment, with the
1459: gravitational potential inferred from the deprojected $I$-band surface
1460: brightness, assuming that mass follows light. We expect the resulting
1461: best-fit constant mass-to-light ratio to be an overestimation due to
1462: possible residual contribution from the quasar light. Even so,
1463: restricting the fit to the unaffected and accurately measured central
1464: second-order velocity moment, we found $\MLdyn = 3.7 \pm
1465: 0.5$\,\MLsunI. When we used instead the gravitational potential that
1466: follows directly, i.e., without mass-to-light ratio conversion, from
1467: the surface mass density of the best-fit lens models, we arrived at a
1468: similar prediction of the second-order velocity moment. We showed that
1469: the reason is that the luminous and total mass distribution, as
1470: inferred from respectively the surface brightness and lens model, are
1471: very similar. This implies that the inner region of the lens galaxy
1472: mass (closely) follows light, with a total mass-to-light ratio $\MLtot
1473: \simeq 3.4$\,\MLsunI.
1474:
1475: By fitting single stellar population models to measured colors of the
1476: center of the lens galaxy, we estimated an $I$-band stellar
1477: mass-to-light ratio $\MLstar$ from 2.8 to 4.1\,\MLsunI. Although a
1478: \emph{constant} dark matter fraction of 20 per cent is thus not
1479: excluded, it is likely that dark matter does not play a significant
1480: role in the inner region of this early-type spiral galaxy of
1481: luminosity $\sim L\star$. This is consistent with indications that
1482: less-luminous early-type galaxies \citep[e.g.][]{2001AJ....121.1936G,
1483: 2006MNRAS.366.1126C, 2006ApJ...649..599K, 2007MNRAS.382..657T} and
1484: late-type galaxies \citep[e.g.][]{2000AJ....120.2884P,
1485: 2001ApJ...550..212B, 2006ApJ...643..804K} lack the dark matter that
1486: is ubiquitous in dwarf galaxies as well as in (at least the outer
1487: parts of) giant elliptical and spiral galaxies.
1488:
1489: The constraint $\alpha = 1.0 \pm 0.1$ on the slope of the lens model
1490: implies that the intrinsic total mass density is close to isothermal,
1491: consistent with previous studies of lens galaxies
1492: \citep[e.g.][]{2006ApJ...649..599K}. However, one has to be very
1493: careful not to over-interpret this result, not only due to assumptions
1494: in the modeling, but most of all because the slope is only (well)
1495: constrained around the Einstein radius $\Rein$. If indeed mass follows
1496: light in the inner region of this lens galaxy, the (deprojected)
1497: surface brightness indicates deviations from isothermal towards the
1498: center where the slope becomes shallower as well as a possible
1499: steepening at larger radii. Interestingly, the construction of
1500: realistic galaxy density profiles with a stellar and dark matter
1501: component that are both non-isothermal, shows that the combined slope
1502: and corresponding lensing properties are nevertheless consistent with
1503: isothermal around $\Rein$ \citep{vandeVenetal2008,Mandelbaumetal2008}.
1504: With more extended images, e.g. in the case of galaxy-galaxy lensing,
1505: one can place a stronger constraint on the total mass distribution of
1506: the lens galaxy \citep[e.g.][]{2007ApJ...666..726B,
1507: 2008arXiv0804.2827S}. Since the extension in these cases is still
1508: mostly tangential, constraining a large radial range is in particular
1509: possible when lensing occurs in multiple (redshift) planes, resulting
1510: in images at different Einstein radii
1511: \citep[e.g.][]{2008ApJ...677.1046G}.
1512:
1513: Nevertheless, already the quasar images provide an accurate constraint
1514: on the total mass within $\Rein$, nearly independent of the details of
1515: the lens model \citep[e.g.][]{1991ApJ...373..354K,
1516: 2001MNRAS.327.1260E}. Therefore, at least around this radius no
1517: higher-order velocity moments are needed to break the mass-anisotropy
1518: degeneracy. Towards the center the surface brightness increases
1519: steeply, so that high enough S/N to measure velocity moments beyond
1520: $V$ and $\sigma$ might be achievable even at higher redshift. In the
1521: outer parts, the degeneracy might be (partially) broken by combining
1522: the kinematics of stars and/or discrete tracers, such as globular
1523: clusters and planetary nebulae, with total mass estimates from hot
1524: X-ray gas and/or weak lensing measurement
1525: \citep[e.g.][]{2007ApJ...667..176G}. In the current and even more in
1526: the upcoming extensive and deep photometric surveys, numerous (strong)
1527: gravitational lensing systems will be discovered. They provide
1528: important and independent constraints on the total mass distribution
1529: in galaxies, especially in combination with (resolved) stellar
1530: kinematics, as we showed in this paper \citep[see
1531: also][]{2007ApJ...666..726B, 2008MNRAS.384..987C}.
1532:
1533: %=====================================================================
1534: % ACKNOWLEDGMENTS
1535: %=====================================================================
1536:
1537: \acknowledgments
1538:
1539: We are grateful to Jean-Ren\'e Roy and Matt Mountain for granting us
1540: director's discretionary time for this project and for generous
1541: hospitality in Hilo to TdZ. We thank Ed Turner for initial support of
1542: this project and Tracy Beck for efficient and cheerful assistance.
1543: GvdV acknowledges support provided by NASA through Hubble Fellowship
1544: grant HST-HF-01202.01-A awarded by the Space Telescope Science
1545: Institute, which is operated by the Association of Universities for
1546: Research in Astronomy, Inc., for NASA, under contract NAS 5-26555. MC
1547: acknowledges support from a STFC Advanced Fellowship (PP/D005574/1).
1548: Based on observations obtained at the Gemini Observatory, which is
1549: operated by the Association of Universities for Research in Astronomy,
1550: Inc., under a cooperative agreement with the NSF on behalf of the
1551: Gemini partnership: the National Science Foundation (United States),
1552: the Particle Physics and Astronomy Research Council (United Kingdom),
1553: the National Research Council (Canada), CONICYT (Chile), the
1554: Australian Research Council (Australia), CNPq (Brazil), and CONICET
1555: (Argentina).
1556:
1557: %=====================================================================
1558: % APPENDICES
1559: %=====================================================================
1560:
1561: \appendix
1562:
1563: %=====================================================================
1564: \section{Multi-Gaussian Expansion}
1565: \label{sec:appmge}
1566: %=====================================================================
1567:
1568: We summarize the (numerically) convenient expressions of the
1569: axisymmetric intrinsic density and gravitational potential in the case
1570: of a Multi-Gaussian Expansion (MGE) of the surface density
1571: \citep{1992A&A...253..366M, 1994A&A...285..723E}. Next, we show that
1572: also the gravitational lensing properties can be readily computed, so
1573: that the MGE method can be used to efficiently construct general lens
1574: models. Finally, we give the line-of-sight second-order velocity
1575: moment as a solution of the axisymmetric Jeans equations which is
1576: discussed in detail in \cite{2008arXiv0806.0042C}.
1577:
1578: %---------------------------------------------------------------------
1579: \subsection{Axisymmetric density and potential}
1580: \label{sec:appmgedenspot}
1581: %---------------------------------------------------------------------
1582:
1583: We parameterize the surface brightness $I(x',y')$ by a sum of
1584: Gaussian components
1585: %
1586: \begin{equation}
1587: \label{eq:mgeIxyproj}
1588: I_j(x',y') = I_{0,j}
1589: \exp\left\{ -\frac{1}{2{\sigma'_j}^2}
1590: \left[ x'^2 + \frac{y'^2}{{q'_j}^2} \right] \right\}.
1591: \end{equation}
1592: %
1593: each with three parameters: the central surface brightness $I_{0,j}$,
1594: the dispersion $\sigma'_j$ along the major $x'$-axis and the
1595: flattening $q'_j$. In case of an (oblate) axisymmetric system viewed
1596: at an inclination $i>0$, the corresponding intrinsic luminosity
1597: density is
1598: %
1599: \begin{equation}
1600: \label{eq:mgenuRz}
1601: \nu_j(R,z) =
1602: \frac{q'_j I_{0,j}}{\sqrt{2\pi}\sigma_j q_j}
1603: \exp\left\{ -\frac{1}{2\sigma_j^2}
1604: \left[ R^2 + \frac{z^2}{q_j^2} \right] \right\},
1605: \end{equation}
1606: %
1607: with intrinsic dispersion $\sigma_j = \sigma'_j$ and intrinsic
1608: flattening $q_j$ given by $q_j^2 \sin^2i = {q'_j}^2 - \cos^2i$.
1609:
1610: The intrinsic mass density follows as $\rho_j = \MLj \, \nu_j$, where
1611: the mass-to-light ratio $\MLj$ per Gaussian component is a free
1612: parameter. Alternatively, if the surface mass density $\Sigma(x',y')$
1613: is available (e.g.\ from gravitational lensing), an
1614: MGE-parametrization as in \eqref{eq:mgeIxyproj} directly yields
1615: $\rho_j$ as in \eqref{eq:mgenuRz}, but with $I_{0,j}$ in both
1616: expressions replaced by the central surface mass density
1617: $\Sigma_{0,j}$.
1618:
1619: The corresponding gravitational potential follows upon (numerical)
1620: evaluation of \citep[][eq.~39]{1994A&A...285..723E}
1621: %
1622: \begin{equation}
1623: \label{eq:mgepotRz}
1624: \Phi_j(R,z) =
1625: -\frac{2 G M_j}{\sqrt{2\pi}\sigma_j} \int_0^1
1626: \mathcal{F}_j(u) \, \du u,
1627: \end{equation}
1628: %
1629: where we have introduced
1630: %
1631: \begin{equation}
1632: \label{eq:mgedefFj}
1633: \mathcal{F}_j(u) = \exp\left\{ -\frac{u^2}{2\sigma_j^2}
1634: \left[ R^2 + \frac{z^2}{{\mathcal{Q}_j^2(u)}} \right]
1635: \right\} \frac{1}{\mathcal{Q}_j(u)},
1636: \end{equation}
1637: %
1638: and $\mathcal{Q}_j^2(u) = 1-(1-q_j^2) \, u^2$.
1639: %
1640: The total mass per Gaussian component is given by $M_j = 2 \pi
1641: {\sigma'_j}^2 {q'_j} \Sigma_{0,j}$, with $\Sigma_{0,j} = \MLj \,
1642: I_{0,j}$ when the mass density is inferred from the surface
1643: brightness.
1644:
1645: In the latter case, we have implicitly assumed a \emph{constant}
1646: mass-to-light ratio $\MLj$ per Gaussian component by taking it outside
1647: the integral. Nevertheless, we can still mimic a (radially) varying
1648: mass-to-light ratio by considering the $\MLj$ of the Gaussian
1649: components as free parameters \citep[see e.g.][]{2006A&A...445..513V,
1650: 2006ApJ...641..852V}. However, it is common in dynamical studies of
1651: the inner parts of galaxies \citep[e.g.][]{2006MNRAS.366.1126C} to
1652: assume the total mass-to-light ratio to be constant, i.e., $\MLj =
1653: \MLtot$ for each Gaussian component $j$. Since $\MLtot$ may be larger
1654: than the stellar mass-to-light ratio $\MLstar$, this still allows for
1655: possible dark matter contribution, but with a constant fraction.
1656:
1657: %---------------------------------------------------------------------
1658: \subsection{Axisymmetric lens model}
1659: \label{sec:appmgeaxilens}
1660: %---------------------------------------------------------------------
1661:
1662: Under the thin-lens approximation, the gravitational lensing
1663: properties of a galaxy are characterized by the deflection potential
1664: $\phi(x',y')$ and its (partial) derivatives. The deflection potential
1665: follows from projecting the potential along the line-of-sight or by
1666: solving the two-dimensional Poisson equation $\nabla^2 \phi = 2
1667: \kappa$. Here, $\kappa = \Sigma/\Sigc$ is the normalized surface mass
1668: density with the critical lensing value $\Sigma_c$ defined in
1669: equation~\eqref{eq:critsurfdens}.
1670:
1671: Similar to the surface brightness in equation~\eqref{eq:mgeIxyproj},
1672: we parameterize this so-called ``convergence'' $\kappa(x',y')$ by a
1673: sum of Gaussian components
1674: %
1675: \begin{equation}
1676: \label{eq:mge_kappa_k}
1677: \kappa_k(x',y') = \frac{\Sigma_{0,j}}{\Sigc}
1678: \exp\left\{ -\frac{1}{2{{\sigma'}_k}^2}
1679: \left[ {x'}^2 + \frac{{y'}^2}{{q'_k}^2} \right] \right\}.
1680: \end{equation}
1681: %
1682: The corresponding deflection potential is then
1683: %
1684: \begin{equation}
1685: \label{eq:mge_phi_k}
1686: \phi_k(x',y') = - \frac{M_k}{\pi\Sigc}
1687: \int_0^1 {\mathcal{F}'_k}(u) \, \frac{\du u}{u},
1688: \end{equation}
1689: %
1690: where $M_k = 2 \pi {\sigma'_k}^2 {q'_k} \Sigma_{0,k}$ is the total
1691: mass per Gaussian component, and we have introduced
1692: %
1693: \begin{equation}
1694: \label{eq:mge_Fp_k}
1695: {\mathcal{F}'_k}(u) = \exp\left\{ -\frac{u^2}{2{\sigma'}_k^2}
1696: \left[ {x'}^2 + \frac{{y'}^2}{{\mathcal{Q}'_k}^2(u)} \right]
1697: \right\} \frac{1}{{\mathcal{Q}'_k}(u)},
1698: \end{equation}
1699: %
1700: and ${\mathcal{Q}'_k}^2(u) = 1-(1-{q'_k}^2) \, u^2$.
1701:
1702: The first-order partial derivatives of $\phi_k(x',y')$ follow as
1703: %
1704: \begin{eqnarray}
1705: \frac{\partial \phi_k}{\partial {x'}}
1706: & = &
1707: {x'} \frac{M_k}{\pi{\sigma'_k}^2\Sigc}
1708: \int_0^1 {\mathcal{F}'_k}(u) \,
1709: u \, \du u,
1710: \nonumber \\ \label{eq:mge_dphi_k}
1711: \frac{\partial \phi_k}{\partial {y'}}
1712: & = &
1713: {y'} \frac{M_k}{\pi{\sigma'_k}^2\Sigc}
1714: \int_0^1 {\mathcal{F}'_k}(u) \,
1715: \frac{u \, \du u}{{\mathcal{Q}'_k}^2(u)},
1716: \end{eqnarray}
1717: %
1718: whereas the second-order partial derivatives are given by
1719: %
1720: \begin{eqnarray}
1721: \frac{\partial^2 \phi_k}{\partial {x'}^2}
1722: & = &
1723: \frac{M_k}{\pi{\sigma'_k}^2\Sigc}
1724: \int_0^1 {\mathcal{F}'_k}(u)
1725: \left[ 1 - \frac{{x'}^2}{{\sigma'_k}^2} \, u^2 \right]
1726: u \, \du u,
1727: \nonumber \\
1728: \frac{\partial^2 \phi_k}{\partial {y'}^2}
1729: & = &
1730: \frac{M_k}{\pi{\sigma'_k}^2\Sigc}
1731: \int_0^1 {\mathcal{F}'_k}(u)
1732: \left[ 1 - \frac{{x'}^2}{{\sigma'_k}^2}
1733: \frac{u^2}{{\mathcal{Q}'_k}^2(u)} \right]
1734: \frac{u \, \du u}{{\mathcal{Q}'_k}^2(u)},
1735: \nonumber \\ \label{eq:mge_dphi2_k}
1736: \frac{\partial^2 \phi_k}{\partial {x'} \partial {y'}}
1737: & = &
1738: \frac{M_k}{\pi{\sigma'_k}^2\Sigc}
1739: \int_0^1 {\mathcal{F}'_k}(u)
1740: \left[ \frac{{x'}{y'}}{{\sigma'_k}^2} \, u^2 \right]
1741: \frac{u \, \du u}{{\mathcal{Q}'_k}^2(u)}.
1742: \end{eqnarray}
1743: %
1744: The above single integrals can be readily evaluated numerically, so
1745: that the lensing properties follow in a straightforward way.
1746: %
1747: The image positions $(x',y')$ are related to the source position
1748: $(\xi,\eta)$ by the lens equation
1749: %
1750: \begin{equation}
1751: \label{eq:mge_lenseq}
1752: \xi = x' - \sum_k \frac{\partial \phi_k}{\partial {x'}},
1753: \qquad
1754: \eta = y' - \sum_k \frac{\partial \phi_k}{\partial {y'}}.
1755: \end{equation}
1756: %
1757: Given the parity $p$ of the image, the magnification $\mu$ is given by
1758: %
1759: \begin{equation}
1760: \label{eq:mge_magnification}
1761: \frac{1}{p\mu} =
1762: \left(1 - \sum_k \frac{\partial^2 \phi_k}{\partial {x'}^2} \right)
1763: \left(1 - \sum_k \frac{\partial^2 \phi_k}{\partial {y'}^2} \right)
1764: - \left( \sum_k \frac{\partial^2 \phi_k}
1765: {\partial {x'} \partial {y'}} \right).
1766: \end{equation}
1767: %
1768: The Einstein radius $\Rein$ follows from solving $\bar{\kappa}(\Rein) =
1769: 1$, where $\bar{\kappa}(R')$ is the average convergence within the
1770: projected radius $R'$. After substituting $x'=R'\cos\theta'$ and
1771: $y'=R'\sin\theta'$ in equation~\eqref{eq:mge_kappa_k} and performing the
1772: integral over $R'$, we are left with
1773: %
1774: \begin{equation}
1775: \label{eq:mge_avkappa}
1776: \bar{\kappa}_k(R') =
1777: \frac{{\sigma'_k}^2\Sigma_{0,k}}{\pi {R'}^2 \Sigc}
1778: \int_0^{2\pi} \left\{
1779: 1 - \exp\left[-\frac{{R'}^2{\mathcal{P}'}(\theta')}{2{\sigma'_k}^2}\right]
1780: \right\} \frac{\du\theta'}{{\mathcal{P}'}(\theta')},
1781: \end{equation}
1782: %
1783: where ${\mathcal{P}'}(\theta') = \cos^2\theta' +
1784: \sin^2\theta'/{q'_k}^2$. The mass within the Einstein radius then
1785: follows as $\Mein = \Sigc \pi \Rein^2$.
1786:
1787: %---------------------------------------------------------------------
1788: \subsection{Axisymmetric Jeans equations}
1789: \label{sec:appmgeaxijeans}
1790: %---------------------------------------------------------------------
1791:
1792: We consider a luminous component with Gaussian intrinsic luminosity
1793: density $\nu_j$ given by \eqref{eq:mgenuRz}, of which the observed
1794: kinematics trace the underlying gravitational potential $\Phi = \sum_k
1795: \Phi_k$, with the sum over all (luminous and dark) components $\Phi_k$
1796: given by \eqref{eq:mgepotRz}.
1797:
1798: Assuming that the velocity ellipsoid is aligned with the cylindrical
1799: $(R,\phi,z)$ coordinate system, so that $\overline{v_R v_z}=0$, we can
1800: readily solve Jeans equation~\eqref{eq:cylcbevz} as
1801: \citep[][eq.~42]{1994A&A...285..723E}
1802: %
1803: \begin{equation}
1804: \label{eq:cylvz2mge}
1805: [\overline{v_z^2}]_j =
1806: 4 \pi G
1807: \sum_k \frac{q'_k \Sigma_{0,k}}{\sqrt{2\pi} \sigma_k}
1808: \int_0^1 q_j^2 \sigma_j^2
1809: \frac{\mathcal{F}_k(u) u^2}{1-F_{jk}u^2} \; \du u,
1810: \end{equation}
1811: %
1812: with $\mathcal{F}_k(u)$ defined in \eqref{eq:mgedefFj} and
1813: %
1814: \begin{equation}
1815: \label{eq:cyldefFjku}
1816: F_{jk} = 1 - q_k^2 - q_j^2 \sigma_j^2/\sigma_k^2.
1817: \end{equation}
1818: %
1819: If we next also assume a constant flattening of the velocity ellipsoid
1820: in the meridional plane, we can write $[\overline{v_R^2}]_j = b_j \,
1821: [\overline{v_z^2}]_j$, and solve Jeans equation~\eqref{eq:cylcbevR} as
1822: %
1823: \begin{equation}
1824: \label{eq:cylvphi2mge}
1825: [\overline{v_\phi^2}]_j =
1826: 4 \pi G
1827: \sum_k \frac{q'_k \Sigma_{0,k}}{\sqrt{2\pi} \sigma_k}
1828: \int_0^1 \left[ b_j q_j^2 \sigma_j^2 + \mathcal{G}_{jk}(u) R^2 \right]
1829: \frac{\mathcal{F}_k(u) u^2}{1-F_{jk}u^2} \; \du u,
1830: \end{equation}
1831: %
1832: where we have introduced
1833: %
1834: \begin{equation}
1835: \label{eq:cyldefGjku}
1836: \mathcal{G}_{jk}(u) =
1837: 1 - b_jq_j^2 - [b_j(1-q_k^2) + (1-b_j)F_{jk}]u^2.
1838: \end{equation}
1839: %
1840: Since the constant $b_j$ may be different for each luminous component
1841: $j$, the total anisotropy in the meridional plane
1842: %
1843: \begin{equation}
1844: \label{eq:cylbetamge}
1845: \beta = 1 - \frac{\overline{v_z^2}}{\overline{v_R^2}}
1846: = 1 - \frac{\sum_j \nu_j [\overline{v_z^2}]_j}
1847: {\sum_j b_j \nu_j [\overline{v_z^2}]_j},
1848: \end{equation}
1849: %
1850: is allowed to vary throughout the system. In case of isotropy in the
1851: meridional plane all $b_j = 1$ and thus $\beta = 0$, corresponding to a
1852: two-integral DF $f(E,L_z)$ \citep[][]{1994A&A...285..723E}.
1853:
1854: After substitution of these expressions in
1855: equation~\eqref{eq:cylvlosmge}, the line-of-sight integral can be
1856: solved, resulting in \citep[][eq.~27]{2008arXiv0806.0042C}
1857: %
1858: \begin{multline}
1859: \label{eq:cylvlos2mge}
1860: [\overline{v_\mathrm{los}^2}]_j =
1861: %\frac{4 \pi^{3/2} G \nu_{0,j}}{I_j(x',y')}
1862: \frac{4 \pi^{3/2} G}{I_j(x',y')} \,
1863: \frac{q'_j I_{0,j}}{\sqrt{2\pi}\sigma_j q_j}
1864: \sum_k \frac{q'_k \Sigma_{0,k}}{\sqrt{2\pi} \sigma_k}
1865: \int_0^1 \frac{q_j^2 \sigma_j^2 (\cos^2i + b_j\sin^2i) +
1866: \mathcal{G}_{jk}(u) x'^2\sin^2i}
1867: {(1-F_{jk}u^2)\sqrt{[1-(1-q_k^2)u^2]
1868: (\mathcal{A}+\mathcal{B}\cos^2i)}}
1869: \\ \times
1870: u^2 \exp\left\{ - \mathcal{A} \left[ x'^2 +
1871: \frac{(\mathcal{A}+\mathcal{B})y'^2}
1872: {\mathcal{A}+\mathcal{B}\cos^2i}
1873: \right] \right\} \, \du u,
1874: \end{multline}
1875: %
1876: where $\mathcal{A}$ and $\mathcal{B}$ are functions of $u$ defined as
1877: %
1878: \begin{eqnarray}
1879: \label{eq:cyldefA}
1880: \mathcal{A} & = & \frac12 \left(
1881: \frac{1}{\sigma_j^2} + \frac{u^2}{\sigma_k^2} \right),
1882: \\
1883: \label{eq:cyldefB}
1884: \mathcal{B} & = & \frac12 \left(
1885: \frac{1-q_j^2}{q_j^2\sigma_j^2} +
1886: \frac{(1-q_k^2)u^4}{[1-(1-q_k^2)u^2]\sigma_k^2} \right).
1887: \end{eqnarray}
1888: %
1889: The remaining single integral can be readily evaluated numerically.
1890:
1891: This provides a direct prediction of the combination $V^2 + \sigma^2$
1892: of the observed mean line-of-sight velocity $V$ and dispersion
1893: $\sigma$ at a given position $(x',y')$ on the sky-plane, through the
1894: (luminosity weighted) sum $\overline{v_\mathrm{los}^2}(x',y') = \sum_j
1895: I_j [\overline{v_\mathrm{los}^2}]_j / \sum_j I_j$ over the
1896: corresponding luminous components.
1897:
1898: %=====================================================================
1899: % REFERENCES
1900: %=====================================================================
1901:
1902: %% \bibliographystyle{apj}
1903: %% \bibliography{ms}
1904: %% replace by contents of 'bbl' file when finished writing
1905:
1906: \begin{thebibliography}{94}
1907: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
1908:
1909: \bibitem[{{Agol} {et~al.}(2000){Agol}, {Jones}, \&
1910: {Blaes}}]{2000ApJ...545..657A}
1911: {Agol}, E., {Jones}, B., \& {Blaes}, O. 2000, \apj, 545, 657
1912:
1913: \bibitem[{{Barnab{\`e}} \& {Koopmans}(2007)}]{2007ApJ...666..726B}
1914: {Barnab{\`e}}, M. \& {Koopmans}, L.~V.~E. 2007, \apj, 666, 726
1915:
1916: \bibitem[{{Barnes} {et~al.}(1999){Barnes}, {Webster}, {Schmidt}, \&
1917: {Hughes}}]{1999MNRAS.309..641B}
1918: {Barnes}, D.~G., {Webster}, R.~L., {Schmidt}, R.~W., \& {Hughes}, A. 1999,
1919: \mnras, 309, 641
1920:
1921: \bibitem[{{Bell} \& {de Jong}(2001)}]{2001ApJ...550..212B}
1922: {Bell}, E.~F. \& {de Jong}, R.~S. 2001, \apj, 550, 212
1923:
1924: \bibitem[{{Binney} \& {Merrifield}(1998)}]{1998gaas.book.....B}
1925: {Binney}, J. \& {Merrifield}, M. 1998, {Galactic astronomy} (Princeton, NJ,
1926: Princeton University Press)
1927:
1928: \bibitem[{{Binney} \& {Tremaine}(1987)}]{BT87}
1929: {Binney}, J. \& {Tremaine}, S. 1987, {Galactic Dynamics} (Princeton, NJ,
1930: Princeton University Press)
1931:
1932: \bibitem[{{Blanton} {et~al.}(1998){Blanton}, {Turner}, \&
1933: {Wambsganss}}]{1998MNRAS.298.1223B}
1934: {Blanton}, M., {Turner}, E.~L., \& {Wambsganss}, J. 1998, \mnras, 298, 1223
1935:
1936: \bibitem[{{Cappellari}(2002)}]{2002MNRAS.333..400C}
1937: {Cappellari}, M. 2002, \mnras, 333, 400
1938:
1939: \bibitem[{{Cappellari}(2008)}]{2008arXiv0806.0042C}
1940: ---. 2008, \mnras, accepted (arXiv:0806.0042)
1941:
1942: \bibitem[{{Cappellari} {et~al.}(2006){Cappellari}, {Bacon}, {Bureau}, {Damen},
1943: {Davies}, {de Zeeuw}, {Emsellem}, {Falc{\'o}n-Barroso}, {Krajnovi{\'c}},
1944: {Kuntschner}, {McDermid}, {Peletier}, {Sarzi}, {van den Bosch}, \& {van de
1945: Ven}}]{2006MNRAS.366.1126C}
1946: {Cappellari}, M., {Bacon}, R., {Bureau}, M., {Damen}, M.~C., {Davies}, R.~L.,
1947: {de Zeeuw}, P.~T., {Emsellem}, E., {Falc{\'o}n-Barroso}, J., {Krajnovi{\'c}},
1948: D., {Kuntschner}, H., {McDermid}, R.~M., {Peletier}, R.~F., {Sarzi}, M., {van
1949: den Bosch}, R.~C.~E., \& {van de Ven}, G. 2006, \mnras, 366, 1126
1950:
1951: \bibitem[{{Cappellari} \& {Copin}(2003)}]{2003MNRAS.342..345C}
1952: {Cappellari}, M. \& {Copin}, Y. 2003, \mnras, 342, 345
1953:
1954: \bibitem[{{Cappellari} \& {Emsellem}(2004)}]{2004PASP..116..138C}
1955: {Cappellari}, M. \& {Emsellem}, E. 2004, \pasp, 116, 138
1956:
1957: \bibitem[{{Carollo} {et~al.}(1995){Carollo}, {de Zeeuw}, {van der Marel},
1958: {Danziger}, \& {Qian}}]{1995ApJ...441L..25C}
1959: {Carollo}, C.~M., {de Zeeuw}, P.~T., {van der Marel}, R.~P., {Danziger}, I.~J.,
1960: \& {Qian}, E.~E. 1995, \apjl, 441, L25
1961:
1962: \bibitem[{{Chae} {et~al.}(1998){Chae}, {Turnshek}, \&
1963: {Khersonsky}}]{1998ApJ...495..609C}
1964: {Chae}, K.-H., {Turnshek}, D.~A., \& {Khersonsky}, V.~K. 1998, \apj, 495, 609
1965:
1966: \bibitem[{{C{\^o}t{\'e}} {et~al.}(2003){C{\^o}t{\'e}}, {McLaughlin}, {Cohen},
1967: \& {Blakeslee}}]{2003ApJ...591..850C}
1968: {C{\^o}t{\'e}}, P., {McLaughlin}, D.~E., {Cohen}, J.~G., \& {Blakeslee}, J.~P.
1969: 2003, \apj, 591, 850
1970:
1971: \bibitem[{{Czoske} {et~al.}(2008){Czoske}, {Barnab{\`e}}, {Koopmans}, {Treu},
1972: \& {Bolton}}]{2008MNRAS.384..987C}
1973: {Czoske}, O., {Barnab{\`e}}, M., {Koopmans}, L.~V.~E., {Treu}, T., \& {Bolton},
1974: A.~S. 2008, \mnras, 384, 987
1975:
1976: \bibitem[{{Dai} {et~al.}(2003){Dai}, {Chartas}, {Agol}, {Bautz}, \&
1977: {Garmire}}]{2003ApJ...589..100D}
1978: {Dai}, X., {Chartas}, G., {Agol}, E., {Bautz}, M.~W., \& {Garmire}, G.~P. 2003,
1979: \apj, 589, 100
1980:
1981: \bibitem[{{de Jong} \& {Bell}(2007)}]{2007iuse.book..107D}
1982: {de Jong}, R.~S. \& {Bell}, E.~F. 2007, {Comparing Dynamical and Stellar
1983: Population Mass-To-Light Ratio Estimates} (Island Universes - Structure and
1984: Evolution of Disk Galaxies), 107
1985:
1986: \bibitem[{{Dejonghe}(1987)}]{1987MNRAS.224...13D}
1987: {Dejonghe}, H. 1987, \mnras, 224, 13
1988:
1989: \bibitem[{{Dolphin}(2000)}]{2000PASP..112.1397D}
1990: {Dolphin}, A.~E. 2000, \pasp, 112, 1397
1991:
1992: \bibitem[{{Emsellem} {et~al.}(1994){Emsellem}, {Monnet}, \&
1993: {Bacon}}]{1994A&A...285..723E}
1994: {Emsellem}, E., {Monnet}, G., \& {Bacon}, R. 1994, \aap, 285, 723
1995:
1996: \bibitem[{{Evans} \& {Witt}(2001)}]{2001MNRAS.327.1260E}
1997: {Evans}, N.~W. \& {Witt}, H.~J. 2001, \mnras, 327, 1260
1998:
1999: \bibitem[{{Evans} \& {Witt}(2003)}]{2003MNRAS.345.1351E}
2000: ---. 2003, \mnras, 345, 1351
2001:
2002: \bibitem[{{Falco} {et~al.}(1996){Falco}, {Lehar}, {Perley}, {Wambsganss}, \&
2003: {Gorenstein}}]{1996AJ....112..897F}
2004: {Falco}, E.~E., {Lehar}, J., {Perley}, R.~A., {Wambsganss}, J., \&
2005: {Gorenstein}, M.~V. 1996, \aj, 112, 897
2006:
2007: \bibitem[{{Fitte} \& {Adam}(1994)}]{1994A&A...282...11F}
2008: {Fitte}, C. \& {Adam}, G. 1994, \aap, 282, 11
2009:
2010: \bibitem[{{Foltz} {et~al.}(1992){Foltz}, {Hewett}, {Webster}, \&
2011: {Lewis}}]{1992ApJ...386L..43F}
2012: {Foltz}, C.~B., {Hewett}, P.~C., {Webster}, R.~L., \& {Lewis}, G.~F. 1992,
2013: \apjl, 386, L43
2014:
2015: \bibitem[{{Franx} {et~al.}(1994){Franx}, {van Gorkom}, \& {de
2016: Zeeuw}}]{1994ApJ...436..642F}
2017: {Franx}, M., {van Gorkom}, J.~H., \& {de Zeeuw}, T. 1994, \apj, 436, 642
2018:
2019: \bibitem[{{Fukazawa} {et~al.}(2006){Fukazawa}, {Botoya-Nonesa}, {Pu}, {Ohto},
2020: \& {Kawano}}]{2006ApJ...636..698F}
2021: {Fukazawa}, Y., {Botoya-Nonesa}, J.~G., {Pu}, J., {Ohto}, A., \& {Kawano}, N.
2022: 2006, \apj, 636, 698
2023:
2024: \bibitem[{{Gavazzi} {et~al.}(2008){Gavazzi}, {Treu}, {Koopmans}, {Bolton},
2025: {Moustakas}, {Burles}, \& {Marshall}}]{2008ApJ...677.1046G}
2026: {Gavazzi}, R., {Treu}, T., {Koopmans}, L.~V.~E., {Bolton}, A.~S., {Moustakas},
2027: L.~A., {Burles}, S., \& {Marshall}, P.~J. 2008, \apj, 677, 1046
2028:
2029: \bibitem[{{Gavazzi} {et~al.}(2007){Gavazzi}, {Treu}, {Rhodes}, {Koopmans},
2030: {Bolton}, {Burles}, {Massey}, \& {Moustakas}}]{2007ApJ...667..176G}
2031: {Gavazzi}, R., {Treu}, T., {Rhodes}, J.~D., {Koopmans}, L.~V.~E., {Bolton},
2032: A.~S., {Burles}, S., {Massey}, R.~J., \& {Moustakas}, L.~A. 2007, \apj, 667,
2033: 176
2034:
2035: \bibitem[{{Gebhardt} {et~al.}(2003){Gebhardt}, {Richstone}, {Tremaine},
2036: {Lauer}, {Bender}, {Bower}, {Dressler}, {Faber}, {Filippenko}, {Green},
2037: {Grillmair}, {Ho}, {Kormendy}, {Magorrian}, \&
2038: {Pinkney}}]{2003ApJ...583...92G}
2039: {Gebhardt}, K., {Richstone}, D., {Tremaine}, S., {Lauer}, T.~R., {Bender}, R.,
2040: {Bower}, G., {Dressler}, A., {Faber}, S.~M., {Filippenko}, A.~V., {Green},
2041: R., {Grillmair}, C., {Ho}, L.~C., {Kormendy}, J., {Magorrian}, J., \&
2042: {Pinkney}, J. 2003, \apj, 583, 92
2043:
2044: \bibitem[{{Gerhard} {et~al.}(2001){Gerhard}, {Kronawitter}, {Saglia}, \&
2045: {Bender}}]{2001AJ....121.1936G}
2046: {Gerhard}, O., {Kronawitter}, A., {Saglia}, R.~P., \& {Bender}, R. 2001, \aj,
2047: 121, 1936
2048:
2049: \bibitem[{{Gerhard}(1993)}]{1993MNRAS.265..213G}
2050: {Gerhard}, O.~E. 1993, \mnras, 265, 213
2051:
2052: \bibitem[{{Hook} {et~al.}(2004){Hook}, {J{\o}rgensen}, {Allington-Smith},
2053: {Davies}, {Metcalfe}, {Murowinski}, \& {Crampton}}]{2004PASP..116..425H}
2054: {Hook}, I.~M., {J{\o}rgensen}, I., {Allington-Smith}, J.~R., {Davies}, R.~L.,
2055: {Metcalfe}, N., {Murowinski}, R.~G., \& {Crampton}, D. 2004, \pasp, 116, 425
2056:
2057: \bibitem[{{Huchra} {et~al.}(1985){Huchra}, {Gorenstein}, {Kent}, {Shapiro},
2058: {Smith}, {Horine}, \& {Perley}}]{1985AJ.....90..691H}
2059: {Huchra}, J., {Gorenstein}, M., {Kent}, S., {Shapiro}, I., {Smith}, G.,
2060: {Horine}, E., \& {Perley}, R. 1985, \aj, 90, 691
2061:
2062: \bibitem[{{Hunter}(1977)}]{1977AJ.....82..271H}
2063: {Hunter}, C. 1977, \aj, 82, 271
2064:
2065: \bibitem[{{Irwin} {et~al.}(1989){Irwin}, {Webster}, {Hewett}, {Corrigan}, \&
2066: {Jedrzejewski}}]{1989AJ.....98.1989I}
2067: {Irwin}, M.~J., {Webster}, R.~L., {Hewett}, P.~C., {Corrigan}, R.~T., \&
2068: {Jedrzejewski}, R.~I. 1989, \aj, 98, 1989
2069:
2070: \bibitem[{{Jeans}(1915)}]{1915MNRAS..76...70J}
2071: {Jeans}, J.~H. 1915, \mnras, 76, 70
2072:
2073: \bibitem[{{Kassin} {et~al.}(2006){Kassin}, {de Jong}, \&
2074: {Weiner}}]{2006ApJ...643..804K}
2075: {Kassin}, S.~A., {de Jong}, R.~S., \& {Weiner}, B.~J. 2006, \apj, 643, 804
2076:
2077: \bibitem[{{Kauffmann} \& {van den Bosch}(2002)}]{2002SciAm.286..36K}
2078: {Kauffmann}, G. \& {van den Bosch}, F. 2002, Scientific American, 286, 36
2079:
2080: \bibitem[{{Kent} \& {Falco}(1988)}]{1988AJ.....96.1570K}
2081: {Kent}, S.~M. \& {Falco}, E.~E. 1988, \aj, 96, 1570
2082:
2083: \bibitem[{{Kochanek}(1991)}]{1991ApJ...373..354K}
2084: {Kochanek}, C.~S. 1991, \apj, 373, 354
2085:
2086: \bibitem[{{Kochanek} {et~al.}(2000){Kochanek}, {Falco}, {Impey}, {Leh{\'a}r},
2087: {McLeod}, {Rix}, {Keeton}, {Mu{\~n}oz}, \& {Peng}}]{2000ApJ...543..131K}
2088: {Kochanek}, C.~S., {Falco}, E.~E., {Impey}, C.~D., {Leh{\'a}r}, J., {McLeod},
2089: B.~A., {Rix}, H.-W., {Keeton}, C.~R., {Mu{\~n}oz}, J.~A., \& {Peng}, C.~Y.
2090: 2000, \apj, 543, 131
2091:
2092: \bibitem[{{Koopmans} \& {Treu}(2003)}]{2003ApJ...583..606K}
2093: {Koopmans}, L.~V.~E. \& {Treu}, T. 2003, \apj, 583, 606
2094:
2095: \bibitem[{{Koopmans} {et~al.}(2006){Koopmans}, {Treu}, {Bolton}, {Burles}, \&
2096: {Moustakas}}]{2006ApJ...649..599K}
2097: {Koopmans}, L.~V.~E., {Treu}, T., {Bolton}, A.~S., {Burles}, S., \&
2098: {Moustakas}, L.~A. 2006, \apj, 649, 599
2099:
2100: \bibitem[{{Krist} \& {Hook}(2004)}]{TinyTim}
2101: {Krist}, J. \& {Hook}, R. 2004, {The Tiny Tim User's Guide} (v6.3)
2102:
2103: \bibitem[{{Kroupa}(2001)}]{2001MNRAS.322..231K}
2104: {Kroupa}, P. 2001, \mnras, 322, 231
2105:
2106: \bibitem[{{Lambas} {et~al.}(1992){Lambas}, {Maddox}, \&
2107: {Loveday}}]{1992MNRAS.258..404L}
2108: {Lambas}, D.~G., {Maddox}, S.~J., \& {Loveday}, J. 1992, \mnras, 258, 404
2109:
2110: \bibitem[{{Lynden-Bell}(1962)}]{1962MNRAS.123..447L}
2111: {Lynden-Bell}, D. 1962, \mnras, 123, 447
2112:
2113: \bibitem[{{Mandelbaum} {et~al.}(2008){Mandelbaum}, {van de Ven}, , \&
2114: {Keeton}}]{Mandelbaumetal2008}
2115: {Mandelbaum}, R., {van de Ven}, G., , \& {Keeton}, C.~R. 2008, \mnras, to be
2116: submitted
2117:
2118: \bibitem[{{Maraston}(1998)}]{1998MNRAS.300..872M}
2119: {Maraston}, C. 1998, \mnras, 300, 872
2120:
2121: \bibitem[{{Maraston}(2005)}]{2005MNRAS.362..799M}
2122: ---. 2005, \mnras, 362, 799
2123:
2124: \bibitem[{{Mediavilla} {et~al.}(1998){Mediavilla}, {Arribas}, {del Burgo},
2125: {Oscoz}, {Serra-Ricart}, {Alcalde}, {Falco}, {Goicoechea}, {Garcia-Lorenzo},
2126: \& {Buitrago}}]{1998ApJ...503L..27M}
2127: {Mediavilla}, E., {Arribas}, S., {del Burgo}, C., {Oscoz}, A., {Serra-Ricart},
2128: M., {Alcalde}, D., {Falco}, E.~E., {Goicoechea}, L.~J., {Garcia-Lorenzo}, B.,
2129: \& {Buitrago}, J. 1998, \apjl, 503, L27
2130:
2131: \bibitem[{{Merritt} {et~al.}(2005){Merritt}, {Navarro}, {Ludlow}, \&
2132: {Jenkins}}]{2005ApJ...624L..85M}
2133: {Merritt}, D., {Navarro}, J.~F., {Ludlow}, A., \& {Jenkins}, A. 2005, \apjl,
2134: 624, L85
2135:
2136: \bibitem[{{Metcalf} {et~al.}(2004){Metcalf}, {Moustakas}, {Bunker}, \&
2137: {Parry}}]{2004ApJ...607...43M}
2138: {Metcalf}, R.~B., {Moustakas}, L.~A., {Bunker}, A.~J., \& {Parry}, I.~R. 2004,
2139: \apj, 607, 43
2140:
2141: \bibitem[{{Monnet} {et~al.}(1992){Monnet}, {Bacon}, \&
2142: {Emsellem}}]{1992A&A...253..366M}
2143: {Monnet}, G., {Bacon}, R., \& {Emsellem}, E. 1992, \aap, 253, 366
2144:
2145: \bibitem[{{Morganti} {et~al.}(1997){Morganti}, {Sadler}, {Oosterloo},
2146: {Pizzella}, \& {Bertola}}]{1997AJ....113..937M}
2147: {Morganti}, R., {Sadler}, E.~M., {Oosterloo}, T., {Pizzella}, A., \& {Bertola},
2148: F. 1997, \aj, 113, 937
2149:
2150: \bibitem[{{Murray} {et~al.}(2003){Murray}, {Allington-Smith}, {Content},
2151: {Davies}, {Dodsworth}, {Miller}, {Jorgensen}, {Hook}, {Crampton}, \&
2152: {Murowinski}}]{2003SPIE.4841.1750M}
2153: {Murray}, G.~J., {Allington-Smith}, J.~R., {Content}, R., {Davies}, R.~L.,
2154: {Dodsworth}, G.~N., {Miller}, B., {Jorgensen}, I., {Hook}, I., {Crampton},
2155: D., \& {Murowinski}, R.~G. 2003, SPIE, 4841, 1750
2156:
2157: \bibitem[{{Navarro} {et~al.}(1997){Navarro}, {Frenk}, \&
2158: {White}}]{1997ApJ...490..493N}
2159: {Navarro}, J.~F., {Frenk}, C.~S., \& {White}, S.~D.~M. 1997, \apj, 490, 493
2160:
2161: \bibitem[{{Palunas} \& {Williams}(2000)}]{2000AJ....120.2884P}
2162: {Palunas}, P. \& {Williams}, T.~B. 2000, \aj, 120, 2884
2163:
2164: \bibitem[{{Primack}(2004)}]{2004IAUS..220...53P}
2165: {Primack}, J.~R. 2004, in IAU Symp.\ 220: Dark Matter in Galaxies, eds. S.~D.\
2166: Ryder, D.~J.\ Pisano, M.~A.\ Walker, and K.~C.\ Freeman, 53
2167:
2168: \bibitem[{{Racine}(1991)}]{1991AJ....102..454R}
2169: {Racine}, R. 1991, \aj, 102, 454
2170:
2171: \bibitem[{{Rix} {et~al.}(1997){Rix}, {de Zeeuw}, {Cretton}, {van der Marel}, \&
2172: {Carollo}}]{1997ApJ...488..702R}
2173: {Rix}, H.-W., {de Zeeuw}, P.~T., {Cretton}, N., {van der Marel}, R.~P., \&
2174: {Carollo}, C.~M. 1997, \apj, 488, 702
2175:
2176: \bibitem[{{Rix} {et~al.}(1992){Rix}, {Schneider}, \&
2177: {Bahcall}}]{1992AJ....104..959R}
2178: {Rix}, H.-W., {Schneider}, D.~P., \& {Bahcall}, J.~N. 1992, \aj, 104, 959
2179:
2180: \bibitem[{{Rix} \& {White}(1992)}]{1992MNRAS.254..389R}
2181: {Rix}, H.-W. \& {White}, S.~D.~M. 1992, \mnras, 254, 389
2182:
2183: \bibitem[{{Romanowsky} {et~al.}(2003){Romanowsky}, {Douglas}, {Arnaboldi},
2184: {Kuijken}, {Merrifield}, {Napolitano}, {Capaccioli}, \&
2185: {Freeman}}]{2003Sci...301.1696R}
2186: {Romanowsky}, A.~J., {Douglas}, N.~G., {Arnaboldi}, M., {Kuijken}, K.,
2187: {Merrifield}, M.~R., {Napolitano}, N.~R., {Capaccioli}, M., \& {Freeman},
2188: K.~C. 2003, Science, 301, 1696
2189:
2190: \bibitem[{{Salpeter}(1955)}]{1955ApJ...121..161S}
2191: {Salpeter}, E.~E. 1955, \apj, 121, 161
2192:
2193: \bibitem[{{Schlegel} {et~al.}(1998){Schlegel}, {Finkbeiner}, \&
2194: {Davis}}]{1998ApJ...500..525S}
2195: {Schlegel}, D.~J., {Finkbeiner}, D.~P., \& {Davis}, M. 1998, \apj, 500, 525
2196:
2197: \bibitem[{{Schmidt} {et~al.}(1998){Schmidt}, {Webster}, \&
2198: {Lewis}}]{1998MNRAS.295..488S}
2199: {Schmidt}, R., {Webster}, R.~L., \& {Lewis}, G.~F. 1998, \mnras, 295, 488
2200:
2201: \bibitem[{{Schneider} {et~al.}(1992){Schneider}, {Ehlers}, \&
2202: {Falco}}]{1992grle.book.....S}
2203: {Schneider}, P., {Ehlers}, J., \& {Falco}, E.~E. 1992, {Gravitational Lenses}
2204: (Springer-Verlag Berlin Heidelberg New York)
2205:
2206: \bibitem[{{Schwarzschild}(1979)}]{1979ApJ...232..236S}
2207: {Schwarzschild}, M. 1979, \apj, 232, 236
2208:
2209: \bibitem[{{Spergel} {et~al.}(2007){Spergel}, {Bean}, {Dor{\'e}}, {Nolta},
2210: {Bennett}, {Dunkley}, {Hinshaw}, {Jarosik}, {Komatsu}, {Page}, {Peiris},
2211: {Verde}, {Halpern}, {Hill}, {Kogut}, {Limon}, {Meyer}, {Odegard}, {Tucker},
2212: {Weiland}, {Wollack}, \& {Wright}}]{2007ApJS..170..377S}
2213: {Spergel}, D.~N., {Bean}, R., {Dor{\'e}}, O., {Nolta}, M.~R., {Bennett}, C.~L.,
2214: {Dunkley}, J., {Hinshaw}, G., {Jarosik}, N., {Komatsu}, E., {Page}, L.,
2215: {Peiris}, H.~V., {Verde}, L., {Halpern}, M., {Hill}, R.~S., {Kogut}, A.,
2216: {Limon}, M., {Meyer}, S.~S., {Odegard}, N., {Tucker}, G.~S., {Weiland},
2217: J.~L., {Wollack}, E., \& {Wright}, E.~L. 2007, \apjs, 170, 377
2218:
2219: \bibitem[{{Suyu} {et~al.}(2008){Suyu}, {Marshall}, {Blandford}, {Fassnacht},
2220: {Koopmans}, {McKean}, \& {Treu}}]{2008arXiv0804.2827S}
2221: {Suyu}, S.~H., {Marshall}, P.~J., {Blandford}, R.~D., {Fassnacht}, C.~D.,
2222: {Koopmans}, L.~V.~E., {McKean}, J.~P., \& {Treu}, T. 2008, ArXiv e-prints,
2223: 0804.2827
2224:
2225: \bibitem[{{Thomas} {et~al.}(2007){Thomas}, {Saglia}, {Bender}, {Thomas},
2226: {Gebhardt}, {Magorrian}, {Corsini}, \& {Wegner}}]{2007MNRAS.382..657T}
2227: {Thomas}, J., {Saglia}, R.~P., {Bender}, R., {Thomas}, D., {Gebhardt}, K.,
2228: {Magorrian}, J., {Corsini}, E.~M., \& {Wegner}, G. 2007, \mnras, 382, 657
2229:
2230: \bibitem[{{Tremaine} {et~al.}(2002){Tremaine}, {Gebhardt}, {Bender}, {Bower},
2231: {Dressler}, {Faber}, {Filippenko}, {Green}, {Grillmair}, {Ho}, {Kormendy},
2232: {Lauer}, {Magorrian}, {Pinkney}, \& {Richstone}}]{2002ApJ...574..740T}
2233: {Tremaine}, S., {Gebhardt}, K., {Bender}, R., {Bower}, G., {Dressler}, A.,
2234: {Faber}, S.~M., {Filippenko}, A.~V., {Green}, R., {Grillmair}, C., {Ho},
2235: L.~C., {Kormendy}, J., {Lauer}, T.~R., {Magorrian}, J., {Pinkney}, J., \&
2236: {Richstone}, D. 2002, \apj, 574, 740
2237:
2238: \bibitem[{{Treu} \& {Koopmans}(2004)}]{2004ApJ...611..739T}
2239: {Treu}, T. \& {Koopmans}, L.~V.~E. 2004, \apj, 611, 739
2240:
2241: \bibitem[{{Trott} \& {Webster}(2002)}]{2002MNRAS.334..621T}
2242: {Trott}, C.~M. \& {Webster}, R.~L. 2002, \mnras, 334, 621
2243:
2244: \bibitem[{{Valluri} {et~al.}(2004){Valluri}, {Merritt}, \&
2245: {Emsellem}}]{2004ApJ...602...66V}
2246: {Valluri}, M., {Merritt}, D., \& {Emsellem}, E. 2004, \apj, 602, 66
2247:
2248: \bibitem[{{van Albada} {et~al.}(1985){van Albada}, {Bahcall}, {Begeman}, \&
2249: {Sancisi}}]{1985ApJ...295..305V}
2250: {van Albada}, T.~S., {Bahcall}, J.~N., {Begeman}, K., \& {Sancisi}, R. 1985,
2251: \apj, 295, 305
2252:
2253: \bibitem[{{van de Ven} {et~al.}(2008{\natexlab{a}}){van de Ven}, {de Zeeuw}, \&
2254: {van den Bosch}}]{2008MNRAS.385..614V}
2255: {van de Ven}, G., {de Zeeuw}, P.~T., \& {van den Bosch}, R.~C.~E.
2256: 2008{\natexlab{a}}, \mnras, 385, 614
2257:
2258: \bibitem[{{van de Ven} {et~al.}(2008{\natexlab{b}}){van de Ven}, {Mandelbaum},
2259: \& {Keeton}}]{vandeVenetal2008}
2260: {van de Ven}, G., {Mandelbaum}, R., \& {Keeton}, C.~R. 2008{\natexlab{b}},
2261: \mnras, to be submitted
2262:
2263: \bibitem[{{van de Ven} {et~al.}(2006){van de Ven}, {van den Bosch}, {Verolme},
2264: \& {de Zeeuw}}]{2006A&A...445..513V}
2265: {van de Ven}, G., {van den Bosch}, R.~C.~E., {Verolme}, E.~K., \& {de Zeeuw},
2266: P.~T. 2006, \aap, 445, 513
2267:
2268: \bibitem[{{van de Ven} {et~al.}(2003){van de Ven}, {van Dokkum}, \&
2269: {Franx}}]{2003MNRAS.344..924V}
2270: {van de Ven}, G., {van Dokkum}, P.~G., \& {Franx}, M. 2003, \mnras, 344, 924
2271:
2272: \bibitem[{{van den Bosch} {et~al.}(2006){van den Bosch}, {de Zeeuw},
2273: {Gebhardt}, {Noyola}, \& {van de Ven}}]{2006ApJ...641..852V}
2274: {van den Bosch}, R., {de Zeeuw}, T., {Gebhardt}, K., {Noyola}, E., \& {van de
2275: Ven}, G. 2006, \apj, 641, 852
2276:
2277: \bibitem[{{van den Bosch} {et~al.}(2008){van den Bosch}, {van de Ven},
2278: {Verolme}, {Cappellari}, \& {de Zeeuw}}]{2008MNRAS.385..647V}
2279: {van den Bosch}, R.~C.~E., {van de Ven}, G., {Verolme}, E.~K., {Cappellari},
2280: M., \& {de Zeeuw}, P.~T. 2008, \mnras, 385, 647
2281:
2282: \bibitem[{{van der Marel}(1991)}]{1991MNRAS.253..710V}
2283: {van der Marel}, R.~P. 1991, \mnras, 253, 710
2284:
2285: \bibitem[{{van der Marel} {et~al.}(1998){van der Marel}, {Cretton}, {de Zeeuw},
2286: \& {Rix}}]{1998ApJ...493..613V}
2287: {van der Marel}, R.~P., {Cretton}, N., {de Zeeuw}, P.~T., \& {Rix}, H. 1998,
2288: \apj, 493, 613
2289:
2290: \bibitem[{{van der Marel} \& {Franx}(1993)}]{1993ApJ...407..525V}
2291: {van der Marel}, R.~P. \& {Franx}, M. 1993, \apj, 407, 525
2292:
2293: \bibitem[{{van Dokkum}(2001)}]{2001PASP..113.1420V}
2294: {van Dokkum}, P.~G. 2001, \pasp, 113, 1420
2295:
2296: \bibitem[{{Vazdekis} {et~al.}(1996){Vazdekis}, {Casuso}, {Peletier}, \&
2297: {Beckman}}]{1996ApJS..106..307V}
2298: {Vazdekis}, A., {Casuso}, E., {Peletier}, R.~F., \& {Beckman}, J.~E. 1996,
2299: \apjs, 106, 307
2300:
2301: \bibitem[{{Vazdekis} {et~al.}(2003){Vazdekis}, {Cenarro}, {Gorgas}, {Cardiel},
2302: \& {Peletier}}]{2003MNRAS.340.1317V}
2303: {Vazdekis}, A., {Cenarro}, A.~J., {Gorgas}, J., {Cardiel}, N., \& {Peletier},
2304: R.~F. 2003, \mnras, 340, 1317
2305:
2306: \bibitem[{{Wambsganss} \& {Paczynski}(1994)}]{1994AJ....108.1156W}
2307: {Wambsganss}, J. \& {Paczynski}, B. 1994, \aj, 108, 1156
2308:
2309: \bibitem[{{Weijmans} {et~al.}(2008){Weijmans}, {Krajnovi{\'c}}, {van de Ven},
2310: {Oosterloo}, {Morganti}, \& {de Zeeuw}}]{2008MNRAS.383.1343W}
2311: {Weijmans}, A.-M., {Krajnovi{\'c}}, D., {van de Ven}, G., {Oosterloo}, T.~A.,
2312: {Morganti}, R., \& {de Zeeuw}, P.~T. 2008, \mnras, 383, 1343
2313:
2314: \bibitem[{{Yee}(1988)}]{1988AJ.....95.1331Y}
2315: {Yee}, H.~K.~C. 1988, \aj, 95, 1331
2316:
2317: \end{thebibliography}
2318:
2319: %=====================================================================
2320: % END DOCUMENT
2321: %=====================================================================
2322:
2323: \end{document}
2324:
2325: %%% Local Variables:
2326: %%% mode: latex
2327: %%% TeX-master: t
2328: %%% End:
2329: