1:
2:
3: \documentclass[12pt,preprint]{aastex}
4:
5: \shorttitle{Non-spherical quasi-relaxed stellar systems}
6: \shortauthors{Bertin and Varri}
7:
8: \bibpunct[; ]{(}{)}{;}{a}{}{;}
9:
10: \begin{document}
11:
12: \title{The construction of non-spherical models of quasi-relaxed stellar systems}
13:
14: \author{G. Bertin and A.L. Varri}
15: \affil{Universit\`{a} degli Studi di Milano, Dipartimento di Fisica, via Celoria 16, I-20133 Milano, Italy; anna.varri@studenti.unimi.it}
16:
17:
18:
19: \begin{abstract}
20:
21: Spherical models of collisionless but quasi-relaxed stellar systems
22: have long been studied as a natural framework for the description of
23: globular clusters. Here we consider the construction of self-consistent
24: models under the same physical conditions, but including explicitly
25: the ingredients that lead to departures from spherical symmetry.
26: In particular, we focus on the effects of the tidal field associated
27: with the hosting galaxy. We then take a stellar system on a circular
28: orbit inside a galaxy represented as a ``frozen" external field. The
29: equilibrium distribution function is obtained from the one describing
30: the spherical case by replacing the energy integral with the relevant
31: Jacobi integral in the presence of the external tidal field. Then the
32: construction of the model requires the investigation of a singular
33: perturbation problem for an elliptic partial differential equation with
34: a free boundary, for which we provide a method of solution to any desired
35: order, with explicit solutions to two orders. We outline the relevant
36: parameter space, thus opening the way to a systematic study of the
37: properties of a two-parameter family of physically justified
38: non-spherical models of quasi-relaxed stellar systems. The general
39: method developed here can also be used to construct models for which
40: the non-spherical shape is due to internal rotation. Eventually, the
41: models will be a useful tool to investigate whether the shapes of
42: globular clusters are primarily determined by internal rotation,
43: by external tides, or by pressure anisotropy.
44:
45: \end{abstract}
46:
47: \keywords{globular clusters: general --- methods: analytical --- stellar dynamics}
48:
49: \section{Introduction}
50:
51: Large stellar systems can be studied as collisionless systems, by
52: means of a one-star distribution function obeying the combined set
53: of the collisionless Boltzmann equation and the Poisson equation,
54: under the action of the self-consistent mean potential. For elliptical
55: galaxies the relevant two-star relaxation times do actually exceed
56: their age; an imprint of partial relaxation may be left at the time
57: of their formation \citetext{if we refer to a picture of formation
58: via incomplete violent relaxation; \citealp{Lyn67}, \citealp{Alb82}},
59: but otherwise they should be thought of as truly collisionless systems,
60: generally characterized by an anisotropic pressure tensor. In turn,
61: for globular clusters the relevant relaxation times are typically
62: shorter than their age, so that we may argue that for many of them
63: the two-star relaxation processes have had enough time to act and
64: to bring them close to a thermodynamically relaxed state, with their
65: distribution function close to a Maxwellian. This line of arguments
66: has led to the development of well-known dynamical models for globular
67: clusters \citep{Kin65,Kin66}.
68:
69: King models are based on a quasi-Maxwellian isotropic distribution
70: function $f_K(E)$ in which a truncation prescription, continuous
71: in phase space, is set heuristically to incorporate the presence
72: of external tides; but otherwise, they are fully self-consistent
73: (i.e., no external fields are actually considered) and perfectly
74: spherical. Empirically, the simplification of spherical symmetry
75: is encouraged by the fact that in general globular clusters have
76: round appearance. Indeed, as a zero-th order description, these
77: models have had remarkable success in applications to observed
78: globular clusters \citep[e.g., see][and references therein]{Spi87,
79: DjoMey94,McLMar05}.
80: In recent years, great progress has been made in the acquisition of
81: detailed quantitative information about the structure of these stellar
82: systems, especially in relation to the measurement of the proper
83: motions of thousands of individual stars \citep[see][]{Lee00,McL06},
84: with the possibility of getting a direct
85: 5-dimensional view of their phase space. Such progress calls for
86: renewed efforts on the side of modeling. More general models would
87: allow us to address the issue of the origin of the observed
88: departures from spherical symmetry. In fact, it remains to be
89: established which physical ingredient among rotation, pressure
90: anisotropy, and tides is the primary cause of the flattening of
91: globular clusters \citep[e.g., see][and references therein]{Kin61,
92: FreFal82,Gey83,WhiSha87,DavPru90,HanRyd94,Ryd96,Goo97,Ber08}
93:
94: As in the case of the study of elliptical galaxies \citep[e.g., see]
95: [and references therein]{BerSti93}, different approaches can be
96: taken to the construction of models.
97: Broadly speaking, two complementary paths can be followed. In the
98: first, ``descriptive" approach, under suitable geometrical (on the
99: intrinsic shape) and dynamical (e.g., on the absence or presence
100: of dark matter) hypotheses, the available data for an individual
101: stellar system are imposed as constraints to derive the internal
102: orbital structure (distribution function) most likely to
103: correspond to the observations. This approach is often carried out
104: in terms of codes that generalize a method introduced by
105: \citet{Sch79}; for an application to the globular cluster
106: $\omega$ Cen, see \citet{Ven06}. In the second, ``predictive"
107: approach, one proposes a formation/evolution scenario in order to
108: identify a physically justified distribution function for a wide
109: class of objects, and then proceeds to investigate, by comparison
110: with observations of several individual objects, whether the data
111: support the general physical picture that has been proposed. Indeed,
112: King models belong to this latter approach.
113:
114: The purpose of this paper is to extend the description of
115: quasi-relaxed stellar systems, so far basically limited to the
116: spherical King models, to the non-spherical case. There are at
117: least three different ways of extending spherical isotropic models
118: of quasi-relaxed stellar systems (such as King models), by
119: modifying the distribution function so as to include: (i) the
120: explicit presence of a non-spherical tidal field; (ii) the
121: presence of internal rotation; (iii) the presence of some pressure
122: anisotropy. As noted earlier in this Introduction, these
123: correspond to the physical ingredients that, separately, may be
124: thought to be at the origin of the observed non-spherical shapes.
125: We will thus focus on the construction of physically justified
126: models, as an extension of the King models, in the presence of
127: external tides and, briefly, on the extension of the models to the
128: presence of rigid internal rotation. A first-order analysis of the
129: triaxial tidal problem addressed in this paper was carried out by
130: \citet{HegRam95} and the effect of a ``frozen'' tidal field on
131: (initially) spherical King models was studied by \citet{Wei93}
132: using N-body simulations.
133: For the axisymmetric problem associated with the presence of internal
134: rotation, a first-order analysis of a simply truncated Maxwellian
135: distribution perturbed by a rigid rotation was given by \citet{KorAna71};
136: different models where differential rotation is considered were
137: proposed by \citet{PreTom70} and by \citet{Wil75}, also in view of
138: extensions to the presence of pressure anisotropy, which goes beyond
139: the scope of this paper. Models that represent a direct generalization
140: of the King family to the case with differential rotation have also
141: been examined, with particular attention to their thermodynamic
142: properties \citep{LagLon96,LonLag96}. In principle, the method of
143: solution that we will present below can deal with the extension of
144: other spherical isotropic models with finite size that are not
145: of the King form \citetext{e.g., see \citealp{WooDic62,Dav77};
146: see also the interesting suggestion by \citealp{Mad96}}.
147:
148: The study of self-consistent collisionless equilibrium models has
149: a long tradition not only in stellar dynamics, but also in plasma
150: physics \citep[e.g., see][]{Har62,AttPeg99}. We note
151: that in both research areas a study in the presence of external
152: fields, especially when the external field is bound to break the
153: natural symmetry associated with the one-component problem, is
154: only rarely considered.
155:
156: The paper is organized as follows. Section 2 introduces the
157: reference physical model, in which a globular cluster is imagined
158: to move on a circular orbit inside a host galaxy treated as a
159: frozen background field; the modified distribution function for
160: such a cluster is then identified and the relevant parameter space
161: defined. In Sect.~3 we set the mathematical problem associated
162: with the construction of the related self-consistent models. For
163: models generated by the spherical $f_K(E)$, Sect.~4 and Appendix
164: A give the complete solution in terms of matched asymptotic
165: expansions. Alternative methods of solutions are briefly discussed
166: in Sect.~5. The concluding Section 6 gives a summary of the paper,
167: with a short discussion of the results obtained. In Appendix B we
168: show how the method developed in this paper can be applied to
169: construct quasi-relaxed models flattened by rotation in the absence
170: of external tides. In Appendix C we show how the method can be
171: applied to other isotropic truncated models, different from King
172: models.
173:
174: Technically, the mathematical problem of a singular perturbation
175: with a free boundary that is faced here is very similar to the
176: problem noted in the theory of rotating stars, starting with Milne
177: \citep[see][]{Tas78,Mil23,Cha33,Kro42,ChaLeb62,MonRox65}. The
178: problem was initially dealt with inadequate tools; a satisfactory
179: solution of the singular perturbation problem was obtained only
180: later, by \citet{Smi75,Smi76}.
181:
182: \section{The physical model}
183:
184: \subsection{The tidal potential}
185:
186: As a reference case, we consider an idealized model in which the
187: center of mass of a globular cluster is imagined to move on a
188: circular orbit of radius $R_0$, characterized by orbital frequency
189: $\Omega$, inside a host galaxy. For simplicity, we focus on the
190: motion of the stars inside the globular cluster and model the
191: galaxy, taken to have very large mass, by means of a {\it frozen}
192: gravitational field (which we will call the galactic field,
193: described by the potential $\Phi_G$), with a given overall
194: symmetry. This choice makes us ignore interesting effects that are
195: generally present in the full interaction between a ``satellite"
196: and a galaxy; in a sense, we are taking a complementary view of an
197: extremely complex dynamical situation, with respect to other
198: investigations, such as those that lead to a discussion of the
199: mechanism of dynamical friction \citetext{in which the globular
200: cluster or satellite is modeled as a rigid body and the stars of
201: the galaxy are taken as the ``live" component; see \citealp{BonAlb87};
202: \citealp{Ber03}; \citealp{AreBer07}; and references therein}.
203: Therefore, we will be initially following the picture of a
204: {\it restricted three-body problem}, with one important
205: difference, that the ``secondary" is not treated as a point mass
206: but as a ``live" stellar system, described by the cluster
207: mean-field potential $\Phi_C$. In this extremely simple orbital
208: choice for the cluster center of mass, in the corotating frame the
209: associated tidal field is time-independent and so we can proceed
210: to the construction of a stationary dynamical model.
211:
212: We consider the galactic potential $\Phi_G$ to be spherically
213: symmetric, i.e. $\Phi_G = \Phi_G (R)$, with $R= \sqrt{X^2
214: +Y^2+Z^2}$, in terms of a standard set of Cartesian coordinates
215: $(X,Y,Z)$, so that $\Omega^2 = (d\Phi_G (R)/dR)_{R_0}/R_0$. Let
216: $(X,Y)$ be the orbit plane of the center of mass of the cluster.
217: We then introduce a local rotating frame of reference, so that the
218: position vector is given by ${\bf r}=(x,y,z)$, with origin in the
219: center of mass of the cluster and for which the x-axis points away
220: from the center of the galactic field, the y-axis follows the
221: direction of the cluster rotation in its orbit around the galaxy,
222: and the z-axis is perpendicular to the orbit plane (according to
223: the right-hand rule). In such rotating local frame, the relevant
224: Lagrangian, describing the motion of a star belonging to the
225: cluster, is \citetext{cf. \citealp{Cha42}, Eq.~[5.510]}:
226: \begin{equation}\label{lagran}
227: \mathcal{L}=\frac{1}{2}\{\dot{x}^2+\dot{y}^2+\dot{z}^2 +\Omega^2[(R_0+x)^2+y^2]
228: +2\Omega(R_0+x)\dot{y}-2\Omega\dot{x}y\}-\Phi_G(R)-\Phi_C(x,y,z)~,
229: \end{equation}
230: where $R =\sqrt{(R_0+x)^2+y^2+z^2}$ and the terms responsible for
231: centrifugal and Coriolis forces are explicitly displayed.
232:
233: If we suppose that the size of the cluster is small compared to
234: $R_0$, we can adequately represent the galactic field by its
235: linear approximation with respect to the local coordinates (the
236: so-called ``tidal approximation"). The corresponding equations of
237: the motion for a single star in the rotating local frame are:
238: \begin{eqnarray}
239: &&\displaystyle\ddot{x}-2\Omega \dot{y}-(4\Omega^2-\kappa^2)x=
240: -\frac{\partial \Phi_C}{\partial x}~, \label{eqmotion1} \\
241: &&\displaystyle\ddot{y}+2\Omega \dot{x}=-\frac{\partial
242: \Phi_C}{\partial y}~, \label{eqmotion2} \\
243: &&\displaystyle\ddot{z}+\Omega^2z=-\frac{\partial \Phi_C}{\partial z}~,\label{eqmotion3}
244: \end{eqnarray}
245: where $\kappa$ is the epicyclic frequency at $R_0$, given by
246: $\kappa^2 = 3\Omega^2 + (d^2\Phi_G/dR^2)_{R_0}$. Note that the
247: assumed symmetry for $\Phi_G$ introduces a cancellation between the
248: kinematic term $y \Omega^2$ and the gradient of the galactic
249: potential $\partial \Phi_G/ \partial y$ and makes the vertical
250: acceleration $ - \partial \Phi_G/ \partial z$ approximately equal
251: to $-z\Omega^2$.
252:
253: These equations admit an energy (isolating) integral of the
254: motion, known as the Jacobi integral:
255: \begin{equation}\label{Jacobi}
256: H = \frac{1}{2}(\dot{x}^2+\dot{y}^2+\dot{z}^2)+\Phi_T+\Phi_C~,
257: \end{equation}
258: where
259: \begin{equation}\label{tidalpot}
260: \Phi_T=\frac{1}{2}\Omega^2\left(z^2-\nu x^2\right)
261: \end{equation}
262: is the tidal potential. Here $\nu\equiv 4 -\kappa^2/\Omega^2$
263: is a generally positive dimensionless coefficient.
264:
265: Thus, at the level of single star orbits, we note that, in
266: general, the tidal potential leads to a compression in the
267: z-direction, a stretching in the x-direction, and leaves the
268: y-direction untouched. The tidal potential is static, breaks the
269: spherical symmetry, but is characterized by reflection symmetry
270: with respect to the three natural coordinate planes; strictly
271: speaking, the symmetry with respect to $(y,z)$ is applicable only
272: in the limit of an infinitely massive host galaxy \citep[see][]{Spi87}.
273: In turn, we will see that the geometry of the tidal
274: potential induces a non-spherical distortion of the cluster shape
275: collectively, in particular an elongation along the x-axis and a
276: compression along the z-axis. In practice, the numerical
277: coefficient $\nu$ that defines quantitatively the induced
278: distortion depends on the galactic potential. We recall that we
279: have $\nu = 3$ for a Keplerian potential, $\nu = 2$ for a
280: logarithmic potential, while for a Plummer model the dimensionless
281: coefficient depends on the location of the circular orbit with
282: respect to the model scale radius $b$, with $\nu(R_0)=3 R_0^2/
283: (b^2+R_0^2)$ \citetext{for a definition of the Plummer model see, e.g.,
284: \citealp{Ber00}}.
285:
286: Different assumptions on the geometry of the galactic field can be
287: treated with tools similar to those developed here, leading to a
288: similar structure of the equations of the motion, with a slight
289: modification of the tidal field. In particular, for an
290: axisymmetric galactic field, the tidal potential differs from the
291: one obtained here only by the z-term \citep{Cha42,HegHut03}. This
292: case is often considered, for example by
293: referring to a globular cluster in circular orbit on the
294: (axisymmetric) disk of our Galaxy \citep[see][]{HegRam95,Ern08},
295: for which $\Phi_T$ is then formulated in terms of the Oort constants.
296:
297: In the physical model outlined in this Section, the typical
298: dynamical time associated with the star orbits inside the cluster
299: is much smaller than the (external) orbital time associated with
300: $\Omega$. Therefore, in an asymptotic sense, the equilibrium
301: configurations that we will construct in the rest of the paper can
302: actually be generalized, with due qualifications, to more general
303: orbits of the cluster inside a galaxy, provided we interpret the
304: results that we are going to obtain as applicable only to a small
305: piece of the cluster orbit.
306:
307: \subsection{The distribution function}
308:
309: As outlined in the Introduction, we wish to extend the description
310: of quasi-relaxed stellar systems (so far basically limited to
311: spherical models associated with distribution functions $f =
312: f(E)$, dependent only on the single-star specific energy $E =
313: v^2/2 + \Phi_C$) to the non spherical case, by including the
314: presence of a non-spherical tidal field explicitly. Given the
315: success of the spherical King models in the study of globular
316: clusters, we will focus on the extension of models based on
317: $f_K(E)$, which is defined as a ``lowered" Maxwellian, continuous
318: in phase space, with an energy cut-off which implies the existence
319: of a boundary at the truncation radius $r_{tr}$.
320:
321: Therefore, we will consider (partially) self-consistent models
322: characterized by the distribution function:
323: \begin{equation}\label{fK}
324: f_K(H)= A [\exp(-aH)- \exp(-aH_0)]
325: \end{equation}
326: \noindent if $H \le H_0$ and $f_K(H) = 0$ otherwise, in terms of
327: the Jacobi integral defined by Eq.~(\ref{Jacobi}). Here $H_0$ is
328: the cut-off value for the Jacobi integral, while $A$ and $a$ are
329: positive constants.
330:
331: In velocity space, the inequality $H \le H_0$ identifies a
332: spherical region given by $0 \le v^2 \le 2 \psi({\bf r})/a$, where:
333: \begin{equation}\label{psi}
334: \psi({\bf r})=a\{H_0-[\Phi_C({\bf r})+\Phi_T(x,z)]\}
335: \end{equation}
336: is the dimensionless escape energy. Therefore, the boundary of the
337: cluster is defined as the relevant zero-velocity surface by the
338: condition $\psi({\bf r}) = 0$ and is given only implicitly by an
339: equipotential ({\it Hill}) surface for the total potential $\Phi_C +
340: \Phi_T$; in fact, its geometry depends on the properties of the tidal
341: potential (of known characteristics; see Eq.~[\ref{tidalpot}]) and
342: of the cluster potential (unknown {\it a priori}, to be determined
343: as the solution of the associated Poisson equation).
344:
345: The value of the cut-off potential $H_0$ should be chosen in such
346: a way that the surface that defines the boundary is closed. The
347: last (i.e., outermost) closed Hill surface is a {\em critical}
348: surface, because it contains two saddle points that represent the
349: Lagrangian points of the restricted three-body problem outlined in
350: the previous subsection. From
351: Eqs.~(\ref{eqmotion1})-(\ref{eqmotion3}), we see that such
352: Lagrangian points are located symmetrically with respect to the
353: origin of the local frame of reference and lie on the x-axis.
354: Their distance from the origin is called the {\em tidal radius},
355: which we denote by $r_T$, and can be determined from the condition:
356: \begin{equation}\label{tidalradius}
357: \frac{\partial \psi}{\partial x}(r_T,0,0)=0~.
358: \end{equation}
359: If, as a zero-th order approximation, we use a simple Keplerian
360: potential for the cluster potential $\Phi_C$, we recover the
361: classical expression (e.g., see Spitzer 1987):
362: \begin{equation}
363: r_T^{(0)} = \left(\frac{G M}{\Omega^2 \nu} \right)^{1/3}~,
364: \end{equation}
365: where $M$ is the total mass of the cluster.
366:
367: As for the spherical King model, the density profile associated
368: with the distribution function (\ref{fK}) is given by:
369: \begin{equation}\label{rho}
370: \rho ( \psi ) = \hat{A} e^ \psi \gamma \left({5 \over 2}, \psi
371: \right) \equiv \hat{A}\hat{\rho}(\psi)~,
372: \end{equation}
373: where $\hat{A} = {8 \pi A \sqrt{2}e^{-a H_0} /(3 a^{3/2})}$. We
374: recall that the incomplete gamma function has non-negative real
375: value only in correspondence to a non-negative argument. In the
376: following, we will denote the central density of the cluster by
377: $\rho_0 = \hat{A}\hat{\rho}(\Psi)$, where $\Psi\equiv \psi({\bf 0})$
378: is the depth of the central potential well.
379:
380: \subsection{The parameter space}\label{parspa}
381:
382: The models defined by $f_K(H)$ are characterized by two physical
383: scales (e.g., the two free constants $A$ and $a$, or,
384: correspondingly, the total mass $M$ and the central density
385: $\rho_0$ of the cluster) and two dimensionless parameters. The
386: first dimensionless parameter can be defined, as in the spherical
387: King models, to measure the {\it concentration} of the cluster. We
388: can thus consider the quantity $\Psi$, introduced at the end of
389: the previous subsection, or we may refer to the commonly used
390: concentration parameter:
391: \begin{equation}\label{concentr}
392: C = \log(r_{tr}/r_0)~,
393: \end{equation}
394: where $r_0 = \sqrt{9/(4 \pi G \rho_0 a)}$ is a scale length related
395: to the size of the core and $r_{tr}$ is the truncation radius of
396: the spherical King model associated with the same value of the
397: central potential well $\Psi$ \citetext{the relation between $C$
398: and $\Psi$ is one-to-one; e.g., see \citealp{Ber00}}.
399:
400: The second dimensionless parameter characterizes the strength of
401: the (external) tidal field:
402: \begin{equation}\label{tidalp}
403: \epsilon \equiv \frac{\Omega^2}{4 \pi G \rho_0}~.
404: \end{equation}
405: The definition arises naturally from the dimensionless formulation
406: of the Poisson equation that describes the (partially)
407: self-consistent problem (to be addressed in the next Section).
408:
409: In principle, for a given choice of the dimensional scales ($A$
410: and $a$) the truncation radius or the concentration parameter of a
411: spherical King model can be set arbitrarily. In practice, the
412: physical motivation of the models suggests that the truncation
413: radius $r_{tr}$ should be taken to be of the order of (and not
414: exceed) the tidal radius $r_T$, introduced in the previous
415: subsection (see Eq.~[\ref{tidalradius}]). We may thus define an
416: {\em extension parameter}, as the ratio between the truncation
417: radius of the corresponding spherical model and the tidal radius
418: $r_T$:
419: \begin{equation}\label{delta}
420: \delta \equiv \frac{r_{tr}}{r_T}~.
421: \end{equation}
422: For a given value of the central potential well $\Psi$,
423: there exists a (maximum) critical value for the tidal strength
424: parameter, which we will denote by $\epsilon_{cr}$, corresponding
425: to the maximum value for the extension parameter $\delta_{cr}$,
426: which can be found by solving the system:
427: \begin{equation}
428: \left\{
429: \begin{array}{ll}\label{sistcrit}
430: \displaystyle \frac{\partial \psi}{\partial
431: x}(r_T,0,0;\epsilon_{cr})=0 \vspace{.1cm}\\ \displaystyle
432: \psi(r_T,0,0;\epsilon_{cr})=0~.
433: \end{array}
434: \right.
435: \end{equation}
436: From this system, if we use the zero-th order Keplerian approximation
437: for $\Phi_C$, we find that $\delta_{cr}^{(0)} = 2/3$ \citep[see][]{Spi87}.
438:
439: For our two-parameter family of models we thus expect two tidal
440: regimes to exist. For models characterized by the pairs $(\Psi,
441: \epsilon)$ near the critical condition $\delta \approx
442: \delta_{cr}$ the tidal distortion should be maximal, while for
443: models with pairs well below criticality only small departures
444: from spherical symmetry should occur. A thorough exploration of
445: the parameter space will be carried out in a separate paper (Varri
446: \& Bertin, in preparation). In closing, we note that the models
447: proposed and studied by \citet{HegRam95} correspond to
448: the pairs in parameter space that we have called critical.
449:
450:
451: \section{The mathematical problem}
452:
453: The (partially) self-consistent models associated with the
454: distribution function defined by Eq.~(\ref{fK}) are constructed by
455: solving the relevant Poisson equation. In terms of the
456: dimensionless escape energy $\psi$, given by Eq.~(\ref{psi}), the
457: Poisson equation (for $\psi \ge 0$) can be written as:
458: \begin{equation}
459: \nabla^2 (\psi + a \Phi_T) = -\frac{9}{r_0^2}\frac{\rho}{\rho_0} =
460: - \frac{9}{r_0^2}\frac{\hat{\rho}(\psi)}{\hat{\rho}(\Psi)}~,
461: \end{equation}
462: where $r_0$ is the scale length introduced in Sect.~\ref{parspa}
463: (see Eq.~[\ref{concentr}]). We then rescale the coordinates and
464: introduce the dimensionless position vector $\hat{{\bf r}} = {\bf r}
465: /r_0$, so that $\hat{\nabla}^2=r_0^2\nabla^2$ and $a \Phi_T \equiv
466: \epsilon T=9 \epsilon (\hat{z}^2- \nu\hat{x}^2)/2$, where we have
467: made use of the tidal parameter introduced in Eq.~(\ref{tidalp}).
468: Therefore, the Poisson equation, for $\psi \ge 0$, can be written
469: in dimensionless form as:
470: \begin{equation}\label{Poisson}
471: \hat{\nabla}^2\psi=-9\left[\frac{\hat{\rho}(\psi)}{\hat{\rho}(\Psi)}
472: +\epsilon(1-\nu)\right]~,
473: \end{equation}
474: while for negative values of $\psi$ we should refer to:
475: \begin{equation}\label{vacuum}
476: \hat{\nabla}^2 \psi=-9\epsilon (1-\nu)~,
477: \end{equation}
478: i.e. the Laplace equation $\hat{\nabla}^2(a\Phi_C)=0$.
479:
480: The mathematical problem is completed by specifying the
481: appropriate boundary conditions. As for the spherical King models,
482: we require regularity of the solution at the origin
483: \begin{eqnarray}
484: \psi({\bf 0})=\Psi~, \label{cond1}\\
485: \hat{\nabla} \psi({\bf 0})={\bf0}~, \label{cond2}
486: \end{eqnarray}
487: and, at large radii:
488: \begin{equation}
489: \psi + \epsilon T \rightarrow a H_0 \label{cond3}~,
490: \end{equation}
491: which corresponds to $a\Phi_C \rightarrow 0$.
492:
493: Poisson and Laplace domains are thus separated by the
494: surface defined by $\psi=0$ which is unknown {\em a priori}; in
495: other words, we have to solve an elliptic partial differential
496: equation in a {\em free boundary} problem.
497:
498: In the ordinary differential problem that characterizes the
499: construction of spherical models with finite mass, the condition
500: of vanishing cluster potential at large radii (together with the
501: regularity conditions at the origin) overdetermines the problem,
502: which can then be seen as an eigenvalue problem \citetext{e.g., see
503: Sect.~2.5 in \citealp{BerSti93}}. Indeed, for the King models the
504: integration of the Poisson equation from the origin outwards, with
505: ``initial conditions" (\ref{cond1})-(\ref{cond2}), sets the relation
506: between the ratio $r_{tr}/r_0$ and $\Psi$ in order to meet the
507: requirement (\ref{cond3}), with $r_{tr}/r_0$ thus playing the role
508: of an ``eigenvalue".
509:
510: In the more complex, three-dimensional situation that we are
511: facing here, the existence of two different domains, internal
512: (Poisson) and external (Laplace), suggests the use of the method
513: of matched asymptotic expansions in order to obtain a uniform
514: solution across the separation free surface. The solution in
515: the internal and external domains are expressed as an asymptotic
516: series with respect to the tidal parameter $\epsilon$, which is
517: assumed to be small (following the physical model described in the
518: previous Section):
519: \begin{equation}\label{asymserin}
520: \psi^{(int)}({\bf \hat{r}};\epsilon)=\sum_{k=0}^{\infty}\frac{1}
521: {k!}\psi_k^{(int)}({\bf \hat{r}}) \epsilon^k~,
522: \end{equation}
523: \begin{equation}\label{asymserext}
524: \psi^{(ext)}({\bf \hat{r}};\epsilon)=\sum_{k=0}^{\infty}\frac{1}
525: {k!}\psi_k^{(ext)}({\bf \hat{r}})
526: \epsilon^k~,
527: \end{equation}
528: with spherical symmetry assumed for the zero-th order terms. The
529: internal solution should obey the boundary conditions
530: (\ref{cond1})-(\ref{cond2}), while the external solution should
531: satisfy Eq.~(\ref{cond3}). The two representations should be
532: properly connected at the surface of the cluster.
533:
534: On the other hand, for any small but finite value of $\epsilon$
535: the boundary, defined by $\psi = 0$, will be different from the
536: unperturbed boundary, defined by $\psi_0 = 0$, so that, for each
537: of the two representations given above, there will be a small
538: region in the vicinity of the surface of the cluster where the
539: leading term is vanishingly small and actually smaller than the
540: remaining terms of the formal series. Therefore, we expect the
541: validity of the expansion to break down where the second term
542: becomes comparable to the first, i.e where $\psi_0 =
543: \mathcal{O}(\epsilon)$. This region can be considered as a {\em
544: boundary layer}, which should be examined in ``microscopic" detail
545: by a suitable rescaling of the spatial coordinates and for which
546: an adequate solution $\psi^{(lay)}$, expressed as a different
547: asymptotic series, should be constructed. To obtain a uniformly
548: valid solution over the entire space, an asymptotic matching is
549: performed between the pairs ($\psi^{(int)}$, $\psi^{(lay)}$) and
550: ($\psi^{(lay)}$, $\psi^{(ext)}$), thus leading to a solution
551: $\psi$, obeying all the desired boundary conditions, in terms of
552: three different, but matched, representations. This method of
553: solution is basically the same method proposed by \citet{Smi75} for
554: the analogous mathematical problem that arises in the
555: determination of the structure of rigidly rotating fluid
556: polytropes.
557:
558: \section{Solution in terms of matched asymptotic expansions}
559:
560: The complete solution to two significant orders in the tidal
561: parameter is now presented. The formal solution to three orders is
562: also displayed because of the requirements of the Van Dyke principle
563: of asymptotic matching \citetext{cf. \citealp{Dyk75}, Eq.~[5.24]} that we have
564: adopted.
565:
566: \subsection{Internal region}
567:
568: If we insert the series (\ref{asymserin}) in the Poisson equation
569: (\ref{Poisson}), under the conditions (\ref{cond1})-(\ref{cond2}),
570: we obtain an (infinite) set of Cauchy problems for $\psi_k$. The
571: problem for the zero-th order term (i.e., the unperturbed problem
572: with $\epsilon = 0$) is the one defining the construction of the
573: spherical and fully self-consistent King models:
574: \begin{equation}\label{ord0}
575: {\psi_0^{(int)}}^{''}+\frac{2}{\hat{r}}
576: {\psi_0 ^{(int)}}'=-9\frac{\hat{\rho}\left(\psi_0^{(int)}\right)}
577: {\hat{\rho}(\Psi)}~,
578: \end{equation}
579: with $\psi_0^{(int)}(0)=\Psi$ and ${\psi_0^{(int)}}'(0)= 0$, where
580: the symbol $'$ denotes derivative with respect to the argument
581: $\hat{r}$. We recall that the truncation radius $\hat{r}_{tr}$,
582: which defines the boundary of the spherical models, is given implicitly
583: by $\psi_0^{(int)}(\hat{r}_{tr})=0$.
584:
585: Let us introduce the quantities:
586: \begin{equation}\label{rj}
587: R_j(\hat{r};\Psi) \equiv \frac{9}{\hat{\rho}(\Psi)}\frac{d^j
588: \hat{\rho}}{d\psi^j} \biggl\vert_{\psi_0^{(int)}}~.
589: \end{equation}
590: These quantities depend on $\hat{r}$ implicitly, through
591: the function $\psi_0^{(int)}$; in turn, the dependence on $\Psi$
592: is both explicit (through the term $\hat{\rho}(\Psi)$) and
593: implicit (because the function $\psi_0^{(int)}(\hat{r})$ depends
594: on the value of $\Psi$). For convenience, we give the
595: expression of the first terms of the sequence (cf. Eq.~[\ref{rho}]):
596: $R_1 =[9/\hat{\rho}(\Psi)][\hat{\rho}(\psi_0^{(int)}) +
597: (\psi_0^{(int)})^{3/2}]$,
598: $R_2=R_1+27(\psi_0^{(int)})^{1/2}/[2\hat{\rho}(\Psi)]$,
599: $R_3=R_2+27(\psi_0^{(int)})^{-1/2}/[4\hat{\rho}(\Psi)]$.
600: Note that for $\psi_0^{(int)}\rightarrow 0$, i.e. for
601: $\hat{r} \rightarrow \hat{r}_{tr}$,
602: $R_1 \rightarrow 0$, $R_2 \rightarrow 0$, while for $j
603: \ge 3$ the quantities $R_j$ actually {\it diverge}. This is one
604: more indication of the singular character of our perturbation
605: analysis, which brings in some fractional power dependence on the
606: perturbation parameter $\epsilon$ (see also expansion [\ref{rhoexp}]).
607:
608: Therefore, the equations governing the next two orders (for
609: $\psi_k^{(int)}$ with $k = 1,2$) can be written as:
610: \begin{equation}\label{ord1}
611: \left[\hat{\nabla}^2 +
612: R_1(\hat{r};\Psi)\right]\psi_1^{(int)}=-9(1-\nu)
613: \end{equation}
614: \begin{equation}\label{ord2}
615: \left[\hat{\nabla}^2+R_1(\hat{r};\Psi)\right]\psi_2^{(int)}=
616: -R_2(\hat{r};\Psi) ({\psi_1^{(int)}})^2
617: \end{equation}
618: \noindent with $\psi_1^{(int)}({\bf 0})=\psi_2^{(int)}({\bf 0})=0$
619: and $\hat{\nabla}\psi_1^{(int)}({\bf
620: 0})=\hat{\nabla}\psi_2^{(int)}({\bf 0}) ={\bf 0}$. The equation
621: for $k=3$ is recorded in Appendix A.1, where we also describe the
622: structure of the general equation for $\psi_k^{(int)}$.
623:
624: For any given order of the expansion, the operator acting on the
625: function $\psi_k^{(int)}$ (see the left-hand side of
626: Eqs.~[\ref{ord1}] and [\ref{ord2}]) is the same, i.e. a Laplacian
627: ``shifted" by the function $R_1(\hat{r};\Psi)$. If we thus expand
628: every term $\psi_k({\bf \hat{r}})$ in spherical
629: harmonics\footnote{We use orthonormalized real spherical
630: harmonics with Condon-Shortley phase; with respect to the toroidal
631: angle $\phi$, they are even for $m \ge 0$ and odd otherwise.}:
632: \begin{equation}
633: \psi_k^{(int)}({\bf
634: \hat{r}})=\sum_{l=0}^{\infty}\sum_{m=-l}^{l}\psi_{k,lm}
635: ^{(int)}(\hat{r})Y_{lm}(\theta,\phi)~,
636: \end{equation}
637: the three-dimensional differential problem is reduced to
638: a one-dimensional (radial) problem, characterized by the following
639: second order, linear ordinary differential operator:
640: \begin{equation}
641: \mathcal{D}_l = \frac{d^2 }{d
642: \hat{r}^2}+\frac{2}{\hat{r}}\frac{d}{d\hat{r}}
643: -\frac{l(l+1)}{\hat{r}^2}+R_1(\hat{r};\Psi)~.
644: \end{equation}
645: \noindent In general, for a fixed value of $l$, two independent
646: solutions to the homogeneous problem $\mathcal{D}_l f = 0$ are
647: expected \footnote{We note that
648: $R_1(0,\Psi)=9[1+\Psi^{3/2}/\hat{\rho}(\Psi)]$, i.e. a numerical positive
649: constant. Therefore, for $\hat{r} \rightarrow 0$ the operator $\mathcal{D}_l$
650: tends to the operator associated with the spherical Bessel
651: functions of the first and of the second kind \citetext{e.g., see
652: \citealp{AbrSte}, Eq.~[10.1.1] for the equation and
653: Eqs.~[10.1.4] and [10.1.5] for the limiting values of the functions
654: for small argument}.} to behave like $\hat{r}^l$ and $1/
655: \hat{r}^{l+1}$ for $\hat{r} \rightarrow 0$. Because of the
656: presence of $R_1(\hat{r};\Psi)$, solutions to equations where
657: $\mathcal{D}_l$ appears have to be obtained numerically.
658:
659: For $k=1$ (see Eq.~[\ref{ord1}]) we thus have to address the
660: following problem. For $l=0$, the relevant equation is:
661: \begin{equation}\label{ord100}
662: \mathcal{D}_0 f_{00}=-9(1-\nu),
663: \end{equation}
664: where $f_{00} \equiv \psi_{1,00}^{(int)}/\sqrt{4\pi}$,
665: with $f_{00}(0) = f_{00}'(0) = 0$. Here we do not have to worry
666: about including solutions to the associated homogeneous problem,
667: because one of the two independent solutions would be singular at
668: the origin and the other would be forced to vanish by the required
669: condition at $\hat{r} = 0$. For $l \ge 1$ we have:
670: \begin{equation}\label{ord1lm}
671: \mathcal{D}_l\psi_{1,lm}^{(int)} = 0
672: \end{equation}
673: with $\psi_{1,lm}^{(int)}(0)={\psi_{1,lm}^{(int)}}'(0)=0$.
674: Both Eq.~(\ref{ord1lm}) and the associated boundary conditions are
675: homogeneous. Therefore, the solution is undetermined by an
676: $m$-dependent multiplicative constant:
677: $\psi_{1,lm}^{(int)}(\hat{r})=A_{lm}\gamma_l (\hat{r})$, with
678: $\gamma_l(\hat{r})\sim \hat{r}^l$ for $\hat{r} \rightarrow 0$ (the
679: singular solution is excluded by the boundary conditions at the
680: origin). Then the complete formal solution is:
681: \begin{equation}\label{sol1}
682: \psi_{1}^{(int)}({\bf \hat{r}})=f_{00}(\hat{r}) +
683: \sum_{l=1}^{\infty}
684: \sum_{m=-l}^{l}A_{lm}\gamma_l(\hat{r})Y_{lm}(\theta,\phi)~,
685: \end{equation}
686: where the constants are ready to be determined by means
687: of the asymptotic matching with $\psi_{1}^{(lay)}({\bf \hat{r}})$
688: at the boundary layer.
689:
690: For $k = 2$ (see Eq.~[\ref{ord2}]) the relevant equations are:
691: \begin{equation}\label{ord2lm}
692: \mathcal{D}_l\psi_{2,lm}^{(int)}=-R_2(\hat{r};\Psi)\left[{\psi_1^{(int)}}
693: ^2\right]_{lm}~,
694: \end{equation}
695: where on the right-hand side the function $\psi_1^{(int)}$ is that
696: derived from the solution of the first order problem (which shows
697: the progressive character of this method for the construction of
698: solutions). In Appendix A.2 the equations for the six relevant
699: harmonics are displayed explicitly. The boundary conditions to be
700: imposed at the origin are again homogeneous: $\psi_{2,lm}^{(int)}
701: (0) = {\psi_{2,lm}^{(int)}}'(0) =0$. For a fixed harmonic $(l,m)$
702: with $l > 0$, the general solution of Eq.~(\ref{ord2lm}) is the sum
703: of a particular solution (which we will denote by $g_{lm} (\hat{r})$)
704: and of a regular solution to the associated homogeneous problem given by
705: Eq.~(\ref{ord1lm}) (which we will call $B_{lm} \gamma_l(\hat{r})$,
706: with $ \gamma_l(\hat{r})$ the same functions introduced for the
707: first order problem). Obviously, the particular solution exists
708: only when Eq.~(\ref{ord2lm}) is non-homogeneous, i.e. only for
709: those values of $(l,m)$ that correspond to a non-vanishing
710: coefficient in the expansion of $({\psi_1^{(int)}})^2$ in spherical
711: harmonics. As noted in the first order problem ($k=1$), the
712: associated homogeneous problem for $l = 0$ has no non-trivial
713: solution. Then we can express the complete solution as:
714: \begin{equation}\label{sol2}
715: \psi_{2}^{(int)}({\bf \hat{r}})= g_{00}(\hat{r}) +
716: \sum_{l=1}^{\infty}\sum_{m=-l}^{l}[g_{lm}
717: (\hat{r})+B_{lm}\gamma_l(\hat{r})]Y_{lm}(\theta,\phi)~,
718: \end{equation}
719: where $B_{lm}$ are constants to be determined from the matching with the
720: boundary layer.
721:
722: Similarly, for $k=3$ the solution can be written as:
723: \begin{equation}\label{sol3}
724: \psi_{3}^{(int)}({\bf \hat{r}}) = h_{00}(\hat{r}) +
725: \sum_{l=1}^{\infty}\sum_{m=-l}^{l}
726: [h_{lm}(\hat{r})+C_{lm}\gamma_l(\hat{r})]Y_{lm}(\theta,\phi)~,\\
727: \end{equation}
728: where $h_{lm}$ are particular solutions and $C_{lm}$ are
729: constants, again to be determined from the matching with the
730: boundary layer.
731:
732: Because the differential operator $\mathcal{D}_l$ and the boundary
733: conditions at the origin are the same for the reduced radial
734: problem of every order, we have thus obtained the general
735: structure of the solution for the internal region (see Appendix
736: A.1).
737:
738: \subsection{External region}
739:
740: Here we first present the general solution and then proceed to set
741: up the asymptotic series (\ref{asymserext}).
742:
743: The solution to Eq.~(\ref{vacuum}) describing the external region,
744: i.e. in the Laplace domain, can be expressed as the sum of a
745: particular solution ($- \epsilon T({\bf \hat{r}})$) and of
746: the solutions to the radial part of the Laplacian operator
747: consistent with the boundary condition (\ref{cond3}):
748: \begin{equation}\label{extgen}
749: \psi^{(ext)}({\bf \hat{r}})=\alpha
750: -\frac{\lambda}{\hat{r}}-\sum_{l=1}^{\infty}\sum_{m=-l}^{l}
751: \frac{\beta_{lm}}{\hat{r}^{l+1}}Y_{lm}(\theta,\phi)-
752: \epsilon T({\bf \hat{r}})~.
753: \end{equation}
754: Here we note that the tidal potential contributes only with spherical
755: harmonics of order $l=0,2$ with even values of $m$:
756: \begin{eqnarray}
757: &&T_{00}(\hat{r})=-3\sqrt \pi (\nu-1)\hat{r}^2 \label{T00}~,\\
758: &&T_{20}(\hat{r})=3\sqrt{ \frac{\pi}{5}} (2+\nu)\hat{r}^2 \label{T20}~,\\
759: &&T_{22}(\hat{r})=-3\sqrt{ \frac{3\pi}{5}} \nu \hat{r}^2 \label{T22}~.
760: \end{eqnarray}
761: At this point we can proceed to set up the asymptotic series, by
762: expanding the constant coefficients $\alpha$, $\lambda$, and
763: $\beta_{lm}$ with respect to $\epsilon$:
764: \begin{equation}\label{alpha}
765: \alpha= aH_0 =
766: \alpha_0+\alpha_1\epsilon+\frac{1}{2!}\alpha_2\epsilon^2+..~,
767: \end{equation}
768: \begin{equation}
769: \lambda=\lambda_0+\lambda_1\epsilon+\frac{1}{2!}\lambda_2\epsilon^2+..~,
770: \end{equation}
771: \begin{equation}
772: \beta_{lm}=a_{lm}\epsilon
773: +\frac{1}{2!}b_{lm}\epsilon^2+\frac{1}{3!} c_{lm}\epsilon^3+..~.
774: \end{equation}
775: The last expansion starts with a first order term because the density
776: distribution of the unperturbed problem is spherically symmetric.
777:
778: For convenience, we give the explicit expression of the external
779: solution up to third order:
780: \begin{eqnarray}\label{ext}
781: &&\psi^{(ext)}({\bf \hat{r}})=
782: \alpha_0-\frac{\lambda_0}{\hat{r}}+\left\{\alpha_1-\frac{\lambda_1}
783: {\hat{r}}-\frac{T_{00}(\hat{r})}{2\sqrt \pi}-\sum_{l=1}^{\infty}
784: \sum_{m=-l}^{l}\left[ \frac{a_{lm}}{\hat{r}^{l+1}}+T_{lm}(\hat{r})\right]
785: Y_{lm}(\theta,\phi)\right\}\epsilon\nonumber\\
786: &&+\frac{1}{2!}\left[\alpha_2-\frac{\lambda_2}{\hat{r}}-
787: \sum_{l=1}^{\infty}\sum_{m=-l}^{l}\frac{b_{lm}}{\hat{r}^{l+1}}
788: Y_{lm}(\theta,\phi)\right]\epsilon^2+\frac{1}{3!}\left[\alpha_3-
789: \frac{\lambda_3}{\hat{r}} -\sum_{l=1}^{\infty}\sum_{m=-l}^{l}
790: \frac{c_{lm}}{\hat{r}^{l+1}}Y_{lm}(\theta,\phi) \right]\epsilon^3~.
791: \end{eqnarray}
792:
793: \subsection{Boundary layer}
794:
795: The boundary layer is the region where the function $\psi$ becomes
796: vanishingly small. Since the unperturbed gravitational field at
797: the truncation radius is finite, $\psi_0'(\hat{r}_{tr})\ne 0$, for
798: any value of $\Psi$, based on a Taylor expansion of $\psi_0$ about
799: $\hat{r}=\hat{r}_{tr}$ we may argue that the region in which the
800: series (\ref{asymserin}) breaks down can be defined by
801: $\hat{r}_{tr}-\hat{r}= \mathcal{O}(\epsilon)$. In this boundary
802: layer we thus introduce a suitable change of variables:
803: \begin{equation}\label{eta}
804: \eta=\frac{\hat{r}_{tr}-\hat{r}}{\epsilon}~,
805: \end{equation}
806: take the ordering $\psi^{(lay)}=\mathcal{O}(\epsilon)$, and thus
807: rescale the solution by introducing the function $\tau \equiv
808: \psi^{(lay)}/\epsilon$. For positive values of $\tau$ the Poisson
809: equation (\ref{Poisson}) thus becomes:
810: \begin{equation}\label{Poislay}
811: \frac{\partial^2 \tau}{\partial \eta^2}-\frac{2 \epsilon}{\hat{r}_{tr}-
812: \epsilon \eta}\frac{\partial \tau}{\partial \eta}+\frac{\epsilon^2}
813: {(\hat{r}_{tr}-\epsilon \eta)^2}\Lambda^2\tau=
814: -\frac{9}{\hat{\rho}(\Psi)}\,\epsilon\hat{\rho}
815: (\epsilon \tau)-9\epsilon^2(1-\nu)~,
816: \end{equation}
817: where $\Lambda^2$ is the angular part of the Laplacian in spherical
818: coordinates. For negative values of $\tau$ we can write a similar
819: equation, corresponding to Eq.~(\ref{vacuum}), which is obtained
820: from Eq.~(\ref{Poislay}) by dropping the term proportional to
821: $\hat{\rho}(\epsilon \tau)$.
822:
823: With the help of the asymptotic expansion for small argument of
824: the incomplete gamma function \citetext{e.g., see \citealp{BenOrs99},
825: Eq.~[6.2.5]}, we find:
826: \begin{equation}\label{rhoexp}
827: \hat{\rho}(\epsilon\tau) \sim \frac{2}{5}\tau^{5/2}\epsilon^{5/2}
828: +\frac{4}{35}\tau^{7/2}\epsilon^{7/2}+...~,
829: \end{equation}
830: so that, within the boundary layer, the contribution of
831: $\hat{\rho}(\epsilon \tau)$ (which is the one that distinguishes
832: the Poisson from the Laplace regime) becomes significant only
833: beyond the tidal term, as a correction
834: $\mathcal{O}(\epsilon^{7/2})$.
835:
836: Therefore, up to $\mathcal{O}(\epsilon^{2})$ we can write:
837: \begin{equation}\label{tau}
838: \tau= \tau_0 + \tau_1 \epsilon + \frac{1}{2!}\tau_2\epsilon^2.
839: \end{equation}
840: To this order, which is required for a full solution up
841: to $k=2$ of the global problem (see Eqs.~[\ref{asymserin}] and
842: [\ref{asymserext}]), by equating in Eq.~(\ref{Poislay}) the first
843: powers of $\epsilon$ separately, we obtain the relevant equations
844: for the first three terms:
845: \begin{equation}
846: \frac{\partial^2 \tau_0}{\partial \eta^2}=0~,
847: \end{equation}
848: \begin{equation}
849: \frac{\partial^2 \tau_1}{\partial \eta^2}=\frac{2}{\hat{r}_{tr}}\frac
850: {\partial \tau_0}{\partial \eta}~,
851: \end{equation}
852: \begin{equation}
853: \frac{\partial^2 \tau_2}{\partial
854: \eta^2}=\frac{4}{\hat{r}_{tr}}\left[ \frac{\partial
855: \tau_1}{\partial \eta}+\frac{\eta}{\hat{r}_{tr}}\frac {\partial
856: \tau_0}{\partial \eta}\right]-\frac{2}{\hat{r}_{tr}^2}\Lambda^2
857: \tau_0-18(1-\nu)~.
858: \end{equation}
859: The equations are easily integrated in the variable $\eta$, to
860: obtain the solutions:
861: \begin{equation}\label{tau0}
862: \tau_0=F_0(\theta,\phi)\eta+G_0(\theta,\phi)~,
863: \end{equation}
864: \begin{equation}\label{tau1}
865: \tau_1=\frac{F_0(\theta,\phi)}{\hat{r}_{tr}}\eta^2+F_1(\theta,\phi)\eta
866: +G_1(\theta,\phi)~,
867: \end{equation}
868: \begin{eqnarray}\label{tau2}
869: &&\tau_2=\frac{2 F_0(\theta,\phi)}{\hat{r}_{tr}^2}\eta^3-\frac{1}
870: {3 \hat{r}_{tr}^2}\Lambda^2 F_0(\theta,\phi)\eta^3+\frac{2F_1
871: (\theta,\phi)}{\hat{r}_{tr}}\eta^2\nonumber\\
872: &&-9(1-\nu)\eta^2-\frac{1}{\hat{r}_{tr}^2}\Lambda^2
873: G_0(\theta,\phi)\eta^2+F_2(\theta,\phi)\eta+G_2(\theta,\phi)~.
874: \end{eqnarray}
875: The six free angular functions that appear in the formal solutions
876: will be determined by the matching procedure.
877:
878: \subsection{Asymptotic matching to two orders}
879:
880: In order to obtain the solution, we must perform separately the
881: relevant matching for the pairs $(\psi^{(int)},\psi^{(lay)})$ and
882: $(\psi^{(lay)},\psi^{(ext)})$. We follow the Van Dyke matching
883: principle, which requires that we compare the second order expansion
884: of the internal and external solutions with the third order expansion
885: of the boundary layer solution. The full procedure is described in
886: Appendix A.3.
887:
888: To first order (i.e., up to $k=1$ in series [\ref{asymserin}] and
889: [\ref{asymserext}]), from the matching of the pair $(\psi^{(int)},
890: \psi^{(lay)})$ we find the free angular functions of (\ref{tau0})
891: and (\ref{tau1}):
892: \begin{equation}\label{F0}
893: F_0(\theta,\phi)=-{\psi_0^{(int)}}'(\hat{r}_{tr})~,
894: \end{equation}
895: \begin{equation}\label{G0}
896: G_0(\theta,\phi)=\psi_1^{(int)}(\hat{r}_{tr},\theta,\phi)~,
897: \end{equation}
898: \begin{equation}\label{F1}
899: F_1(\theta,\phi)=-\frac{\partial \psi_1^{(int)}}{\partial \hat{r}}
900: (\hat{r}_{tr},\theta,\phi)~,
901: \end{equation}
902: \begin{equation}\label{G1}
903: G_1(\theta,\phi)=\frac{1}{2}\psi_2^{(int)}(\hat{r}_{tr},\theta,\phi)~.
904: \end{equation}
905: From the matching of the pair $(\psi^{(ext)},\psi^{(lay)})$ we
906: connect $\psi^{(ext)}$ to the same angular functions, thus proving
907: that the matching to first order is equivalent to imposing continuity
908: of the solution up to second order and of the first radial derivative
909: up to first order. This allows us to determine the free constants that
910: are present in the first two terms of (\ref{ext}) and in (\ref{sol1}):
911: \begin{equation}\label{alpha0}
912: \alpha_0=\frac{\lambda_0}{\hat{r}_{tr}}~,
913: \end{equation}
914: \begin{equation}\label{lambda0}
915: \lambda_0=\hat{r}_{tr}^2{\psi_0^{(int)}}'(\hat{r}_{tr})~,
916: \end{equation}
917: \begin{equation}\label{alpha1}
918: \alpha_1=f_{00}(\hat{r}_{tr})+\hat{r}_{tr}f_{00}'(\hat{r}_{tr})+\frac
919: {3 T_{00}(\hat{r}_{tr})}{2\sqrt \pi}~,
920: \end{equation}
921: \begin{equation}\label{lambda1}
922: \lambda_1=\hat{r}_{tr}^2f_{00}'(\hat{r}_{tr})+\frac{\hat{r}_{tr}T_{00}
923: (\hat{r}_{tr})}{ \sqrt\pi}~,
924: \end{equation}
925: \begin{equation}\label{A2m}
926: A_{2m}=-\frac{5 T_{2m}(\hat{r}_{tr})}{\hat{r}_{tr}\gamma_2'(\hat{r}
927: _{tr})+3\gamma_2(\hat{r}_{tr})}~,
928: \end{equation}
929: \begin{equation}\label{a2m}
930: a_{2m}=-\hat{r}_{tr}^3[A_{2m}\gamma_2(\hat{r}_{tr})+
931: T_{2m}(\hat{r}_{tr})]~.
932: \end{equation}
933: Note that $A_{lm}=a_{lm}=0$ if $l\ne 2$, for every value of $m$, and that
934: the constants for $l=2$ are non-vanishing only for $m=0,2$. The constants
935: that identify the solution are thus expressed in terms of the values of the
936: unperturbed field ${\psi_0^{(int)}}'$, of the ``driving" tidal potential
937: $T_{lm}$, and of the solutions $f_{00}$ and $\gamma_2$ (see Eqs.~[\ref{ord100}]
938: and [\ref{ord1lm}]) taken at $\hat{r} = \hat{r}_{tr}$.
939:
940: The boundary surface of the first order model is defined implicitly by
941: $\psi_0^{(ext)}(\hat{r})+\psi_1^{(ext)}(\hat{r},\theta,\phi)\epsilon=0$,
942: i.e. the spherical shape of the King model is modified by monopole and
943: quadrupole contributions, which are even with respect to toroidal and
944: poloidal angles and characterized by reflection symmetry with respect to
945: the three natural coordinates planes. As might have been expected from
946: the physical model, the spherical shape is thus modified only by spherical
947: harmonics $(l,m)$ for which the tidal potential has non-vanishing coefficients.
948: Mathematically, this is non-trivial, because the first order equation
949: in the internal region Eq.~(\ref{ord1}) is non-homogeneous only for $l=0$;
950: the quadrupole contribution to the internal solution is formally ``hidden"
951: by the use of the function $\psi$ (which includes the tidal potential) and
952: is unveiled by the matching which demonstrates that $A_{2m}$
953: with $m=0,2$ are non-vanishing.
954:
955: The first order solution can be inserted into the right-hand side of
956: Eq.~(\ref{ord2lm}) to generate non-homogeneous equations (and thus
957: particular solutions) only for $l=0,2,4$ and corresponding positive
958: and even values of $m$ (see Appendix A.2). We can thus proceed to
959: construct the second order solution in the same way described above
960: for the first order solution. From the matching of the pair
961: $(\psi^{(int)},\psi^{(lay)})$ we determine the missing angular
962: functions:
963: \begin{equation}\label{F2}
964: F_2(\theta,\phi)=-\frac{\partial \psi_2^{(int)}}{\partial \hat{r}}
965: (\hat{r}_{tr},\theta,\phi)~,
966: \end{equation}
967: \begin{equation}\label{G2}
968: G_2(\theta,\phi)=\frac{1}{3}\psi_3^{(int)}(\hat{r}_{tr},\theta,\phi)~,
969: \end{equation}
970: which are then connected to the properties of $\psi^{(ext)}$ by the
971: matching of the pair $(\psi^{(ext)},\psi^{(lay)})$. This is equivalent
972: to imposing continuity of the solution up to third order and of the
973: first radial derivative up to second order and leads to the determination
974: of the free constants that appear in the third term of (\ref{ext}) and
975: in (\ref{sol2}):
976: \begin{equation}\label{alpha2}
977: \alpha_2=g_{00}(\hat{r}_{tr})+\hat{r}_{tr}g_{00}'(\hat{r}_{tr})~,
978: \end{equation}
979: \begin{equation}\label{lambda2}
980: \lambda_2=\hat{r}_{tr}^2g_{00}'(\hat{r}_{tr})~,
981: \end{equation}
982: \begin{equation}\label{B2m}
983: B_{2m}=-\frac{\hat{r}_{tr} g_{2m}'(\hat{r}_{tr})+3g_{2m}(\hat{r}_{tr})}
984: {\hat{r}_{tr}\gamma_2'(\hat{r}_{tr})+3\gamma_2(\hat{r}_{tr})}~,
985: \end{equation}
986: \begin{equation}\label{b2m}
987: b_{2m}=-\hat{r}_{tr}^3[g_{2m}(\hat{r}_{tr} )+B_{2m} \gamma_2(\hat{r}_{tr})]~,
988: \end{equation}
989: \begin{equation}\label{B4m}
990: B_{4m}=-\frac{\hat{r}_{tr} g_{4m}'(\hat{r}_{tr})+5g_{4m}(\hat{r}_{tr})}
991: {\hat{r}_{tr}\gamma_4'(\hat{r}_{tr})+5\gamma_4(\hat{r}_{tr})}~,
992: \end{equation}
993: \begin{equation}\label{b4m}
994: b_{4m}=-\hat{r}_{tr}^5[g_{4m}(\hat{r}_{tr} )+B_{4m} \gamma_4(\hat{r}_{tr})]~.
995: \end{equation}
996: Here $B_{lm}=b_{lm}=0$ if $l\ne 2,4$ for every value of $m$; the only
997: non-vanishing constants with $l=2,4$ are those with even $m$.
998:
999: Therefore, the second order solution has non-vanishing contributions only
1000: for $l=0, 2,4$, i.e. for those harmonics for which the particular solution to
1001: Eq.~(\ref{ord2}) is non-trivial. {\it By induction}, it can be proved (see
1002: Appendix A.4) that the $k$-th order solution is characterized by $l=0,2,..,2k$
1003: harmonics with corrisponding positive and even values of $m$. In reality,
1004: the discussion of the matching to higher orders ($k>3$) would require a
1005: re-definition of the boundary layer, because the density contribution on the
1006: right-hand side of Eq.~(\ref{Poislay}) (for positive values of $\tau$) comes
1007: into play. The asymptotic matching procedure carries through also in this
1008: more complex case but, for simplicity, is omitted here. We should also keep
1009: in mind that in an asymptotic analysis the inclusion of higher order terms
1010: does not necessarily lead to better accuracy in the solution; the optimal
1011: truncation in the asymptotic series depends on the value of the expansion
1012: parameter (in this case, on the value of $\epsilon$) and has to be judged
1013: empirically.
1014:
1015: In conclusion, starting from a given value of the King concentration parameter
1016: $\Psi$ and from a given strength of the tidal field $\epsilon$, the uniform
1017: triaxial solution is constructed by numerically integrating Eqs.~(\ref{ord0}),
1018: (\ref{ord100}), (\ref{ord1lm}), and (\ref{ord2lm}) and by applying the
1019: constants derived in this subsection to the asymptotic series expansion
1020: (\ref{asymserin})-(\ref{asymserext}). The numerical integrations can be
1021: performed efficiently by means of standard Runge-Kutta routines. The boundary
1022: surface of the model is thus defined by $\psi_0^{(ext)}(\hat{r})+\psi_1^{(ext)}
1023: (\hat{r},\theta,\phi)\epsilon + \psi_2^{(ext)}(\hat{r},\theta,\phi)\epsilon^2/2=0$,
1024: while the internal density distribution is given by $\rho = \rho(\psi_0^{(int)}
1025: (\hat{r})+\psi_1^{(int)}(\hat{r},\theta,\phi)\epsilon + \psi_2^{(int)}(\hat{r},
1026: \theta,\phi)\epsilon^2/2)$, with the function $\rho$ defined by Eq.~(\ref{rho}).
1027: Any other ``observable" quantity can be reconstructed by suitable integration in
1028: phase space of the distribution function $f_K(H)$ defined by Eq.~(\ref{fK}), with
1029: $H$ defined by Eq.~(\ref{Jacobi}), and $\Phi_T + \Phi_C = H_0 - [\psi_0^{(int)}
1030: (\hat{r})+\psi_1^{(int)}(\hat{r},\theta,\phi)\epsilon + \psi_2^{(int)}(\hat{r},
1031: \theta,\phi)\epsilon^2/2]/a$.
1032:
1033: In Fig.~1 and Fig.~2 we illustrate the main characteristics of one
1034: triaxial model constructed with the method described in this Section.
1035:
1036:
1037: \section{Alternative methods of solution}
1038:
1039: \subsection{The method of strained coordinates}
1040:
1041: The mathematical problem described in Sect.~3 can also be solved
1042: by the method of {\em strained coordinates}, an alternative method
1043: usually applied to
1044: non-linear hyperbolic differential equations \citetext{e.g., see
1045: \citealp{Dyk75}, Chapter 6} and considered by \citet{Smi76} in the
1046: solution of the singular free-boundary perturbation problem that
1047: arises in the study of rotating polytropes.
1048:
1049: Starting from a series representation of the form
1050: (\ref{asymserin}) and (\ref{asymserext}) for the solution defined in
1051: the Poisson and Laplace domains, respectively, a transformation is
1052: considered from spherical coordinates $(\hat{r},\theta,\phi)$ to
1053: ``strained coordinates" $(s,p,q)$:
1054: \begin{eqnarray}\label{trans}
1055: && \hat{r}=s+\epsilon \hat{r}_1(s,p,q)+\frac{1}{2} \epsilon^2
1056: \hat{r}_2(s,p,q)+... \nonumber \\ &&\theta=p \nonumber \\
1057: &&\phi=q~,
1058: \end{eqnarray}
1059: where $\hat{r}_k(s,p,q)$ are initially unspecified straining functions.
1060: We note that the zero-th order problem is defined by the same
1061: Eq.~(\ref{ord0}) with the same boundary conditions but with the variable
1062: $\hat{r}$ replaced by $s$. The unperturbed spherical boundary in the
1063: strained space is defined by $s=s_0$, where $\psi_0^{(int)}(s_0)=0$.
1064: To each order, the effective boundary of the perturbed configuration
1065: remains described by the surface $s=s_0$, while in physical coordinates
1066: the truncation radius actually changes as a result of the straining
1067: functions $\hat{r}_k$ that are determined progressively.
1068:
1069: The Laplacian expressed in the new coordinates, $\widetilde{\nabla}^2$,
1070: can be written as an asymptotic series: $\widetilde{\nabla}^2= L_0+
1071: \epsilon L_1+1/2 \epsilon^2 L_2+...$, where $L_k$ are linear second order
1072: operators \footnote{Surfaces with constant $s$ in the strained space are
1073: assumed to correspond to surfaces with constant $\psi^{(int)}$ in the
1074: physical space, i.e. $\psi^{(int)}=\psi^{(int)}(s)$; therefore, $L_k$
1075: (with $k \ge 0$) is an ordinary differential operator for $\psi^{(int)}$.}
1076: in which $\hat{r}_j(s,p,q)$ (with $j=1,..,k$) and their derivatives appear.
1077: For convenience, we record the zero-th and first order operators:
1078: \begin{equation}
1079: L_0\equiv \frac{d ^2}{d s^2}+ \frac{2}{s}\frac{d}{d s}~,
1080: \end{equation}
1081: \begin{equation}\label{L1}
1082: L_1\equiv -\left(2\frac{\partial \hat{r}_1}{\partial s}\right)
1083: \frac{d^2}{d s^2}-\left(\frac{\partial^2 \hat{r}_1}{\partial s^2}+
1084: \frac{2}{s}\frac{\partial\hat{r}_1}{\partial s} + \frac{1}{s^2}\Lambda^2
1085: \hat{r}_1+\frac{2}{s^2}\hat{r}_1\right) \frac{d }{d s}~,
1086: \end{equation}
1087: where $\Lambda^2$ is the standard angular part of the Laplacian, written
1088: with angular coordinates $(p,q)$. The general $k$-th order operator can
1089: be decomposed as $L_k=L_1+F_k$, where $F_k$ is, in turn, a second order
1090: operator in which $\hat{r}_j(s,p,q)$ (with $j=1,..,k-1$) appear and $L_1$
1091: is defined as in Eq.~(\ref{L1}) but with $\hat{r}_k(s,p,q)$ instead of
1092: $\hat{r}_1(s,p,q)$; these operators appear in the relevant equation for
1093: $\psi_k^{(int)}$:
1094: \begin{equation}
1095: [L_0+R_1(\psi;\epsilon)] \psi_k^{(int)}=L_k\psi_0^{(int)}
1096: \end{equation}
1097: which corresponds to the general $k$-th order equation of the previous method.
1098:
1099: Following a set of constraints that guarantee the regularity of the
1100: series (\ref{asymserin}) in the strained space, the equations that
1101: uniquely identify the straining functions to any desired order can
1102: be found and solved numerically; structurally, they somewhat correspond
1103: to Eqs.~(\ref{ord1lm}) and (\ref{ord2lm}) in Sect.~4.1. Therefore, the
1104: internal and external solutions can be worked out and patched by requiring
1105: continuity of the solution and of the first derivative with respect to
1106: the variable $s$ at the boundary surface defined by $s=s_0$, in general
1107: qualitative analogy with the method described in the previous section.
1108:
1109: This method is formally more elegant than the method of matched
1110: asymptotic expansions but requires a more significant
1111: numerical effort because, even though the number of equations to
1112: be solved at each order is the same, the operator that plays here
1113: a central role in the equations for the straining functions,
1114: $L_1\psi_0^{(int)}$ (interpreted as an operator acting on $\hat{r}_k[s,p,q]$),
1115: is more complex than $\mathcal{D}_l$ (defined in Sect.~4.1).
1116:
1117: \subsection{Iteration}
1118:
1119: This technique follows the approach taken by \citet{PreTom70}
1120: and by \citet{Wil75}, for the construction of
1121: self-consistent dynamical models of differentially rotating
1122: elliptical galaxies, and later by \citet{LonLag96},
1123: for their extension of King models to the rotating case.
1124:
1125: In terms of the function:
1126: \begin{equation}
1127: u({\bf \hat{r}}) \equiv a[H_0-\Phi_C({\bf \hat{r}})]= \psi({\bf
1128: \hat{r}})+ \epsilon T({\bf \hat{r}})~,
1129: \end{equation}
1130: inside the cluster the Poisson equation can be written as:
1131: \begin{equation}\label{iterPoi}
1132: \hat{\nabla}^2 u=-\frac{9}{\hat{\rho}(\Psi)}\hat{\rho}(u-\epsilon T)~,
1133: \end{equation}
1134: while outside the cluster the Laplace equation is simply:
1135: \begin{equation}\label{iterLap}
1136: \hat{\nabla}^2 u=0~.
1137: \end{equation}
1138: The boundary conditions at the origin are $u({\bf0})=\Psi$ and
1139: $\hat{\nabla}u({\bf 0})={\bf 0}$, because the tidal potential
1140: $T({\bf \hat{r}})$ is a homogeneous function; the condition at
1141: large radii is $u \rightarrow aH_0$.
1142:
1143: The basic idea is to get an improved solution $u^{(n)}$ of the
1144: Poisson equation by evaluating the ``source term" on the right-hand
1145: side with the solution obtained in the immediately previous step:
1146: \begin{equation}\label{iterPoin}
1147: \hat{\nabla}^2
1148: u^{(n)}=-\frac{9}{\hat{\rho}(\Psi)}\hat{\rho}(u^{(n-1)}-\epsilon T)~.
1149: \end{equation}
1150: The iteration is seeded by inserting as $u^{(0)}$, on the right-hand
1151: side of Eq.~(\ref{iterPoin}), the spherical solution of the King models.
1152: The iteration continues until convergence is reached.
1153:
1154: In order to solve Eq.~(\ref{iterPoin}), we expand in spherical
1155: harmonics the solution and, correspondingly, the dimensionless
1156: density distribution:
1157: \begin{equation}\label{uexp}
1158: u^{(n)}({\bf \hat{r}
1159: })=\sum_{l=0}^{\infty}\sum_{m=-l}^{l}u_{lm}^{(n)}
1160: (\hat{r})Y_{lm}(\theta,\phi)~,
1161: \end{equation}
1162: \begin{equation}\label{rhoitexp}
1163: \hat{\rho}^{(n)}({\bf \hat{r}
1164: })=\sum_{l=0}^{\infty}\sum_{m=-l}^{l}\hat
1165: {\rho}_{lm}^{(n)}(\hat{r})Y_{lm}(\theta,\phi)~,
1166: \end{equation}
1167: so that the reduced radial problems for the functions $u_{lm}^{(n)}
1168: (\hat{r})$ are:
1169: \begin{equation}\label{equlm}
1170: \left[\frac{d^2 }{ d \hat{r}^2}+\frac{2}{\hat{r}}{d \over d
1171: \hat{r}} -{l(l+1)\over \hat{r}^2}
1172: \right]u_{lm}^{(n)}=-\frac{9}{\hat{\rho}(\Psi)}
1173: \hat{\rho}_{lm}^{(n-1)}~,
1174: \end{equation}
1175: with boundary conditions $u_{00}^{(n)}(0)=\Psi$, $u_{lm}^{(n)}(0)=0$
1176: and ${u_{00}^{(n)}}'(0)={u_{lm}^{(n)}}'(0)=0$.
1177: Here, in contrast with the structure of the governing equations
1178: for $\psi_k^{(int)}$ of Sect.~4.1 and Sect.~4.2, the radial part of
1179: the Laplacian appears with no ``shift", for which the homogeneous
1180: solutions are known analytically. Thus the full solution to
1181: Eq.~(\ref{equlm}) can be obtained in integral form by the standard
1182: method of {\em variation of the arbitrary constants}:
1183: \begin{equation}\label{u00}
1184: u_{00}^{(n)}(\hat{r})=\Psi -\frac{9}{\hat{\rho}(\Psi)}\left[
1185: \int_0^{\hat{r}}\hat{r}'\hat{\rho}_{00}^{(n-1)}
1186: (\hat{r}')d\hat{r}'-\frac{1}{\hat{r}}\int_0^{\hat{r}}
1187: \hat{r}'^2\hat{\rho}_{00}^{(n-1)}(\hat{r}')d\hat{r}'\right]~,
1188: \end{equation}
1189: \begin{equation}\label{ulm}
1190: u_{lm}^{(n)}(\hat{r})= \frac{9}{(2l+1)\hat{\rho}(\Psi)}\left[\hat{r}^{l}
1191: \int_{\hat{r}}^{\infty}
1192: \hat{r}'^{1-l}\hat{\rho}_{lm}^{(n-1)}(\hat{r}')d\hat{r}'
1193: +\frac{1}{\hat{r}^{l+1}}\int_0^{\hat{r}}
1194: \hat{r}'^{l+2} \hat{\rho}_{lm}^{(n-1)}(\hat{r}')d\hat{r}'\right]~.
1195: \end{equation}
1196: The complete calculation can be found in the Appendix of \citet{PreTom70}.
1197: Here we only remark that this integral form is valid in both Poisson and
1198: Laplace domains because it contains simultaneously the regular and the
1199: singular homogeneous solutions of the Laplacian. In the derivation, all
1200: the boundary conditions have been used; in particular, the two conditions
1201: at the origin are sufficient to obtain expression (\ref{u00}), while for
1202: expression (\ref{ulm}) the one concerning the radial derivative at the origin
1203: is used together with the one that describes the behavior at large radii (i.e.
1204: $u_{lm}^{(n)}\rightarrow 0$ for $l \ge 1$). Furthermore, from the condition
1205: at large radii evaluated for the harmonic $l=0$, i.e. $u_{00}^{(n)}/\sqrt{4\pi}
1206: \rightarrow aH_0^{(n)}$ (here the notation reminds us that the value of $H_0$
1207: is known only approximately and it changes slightly at every iteration), we find:
1208: \begin{equation}
1209: aH_0^{(n)}\equiv \frac{\Psi}{\sqrt{4\pi}} -\frac{9}{\sqrt{4\pi}
1210: \hat{\rho}(\Psi)}\int_0^{\infty}\hat{r}'
1211: \hat{\rho}_{00}^{(n-1)}(\hat{r}')d\hat{r}'~,
1212: \end{equation}
1213: where we should recall that beyond a certain radius $\hat{\rho}_{00}^{(n-1)}$
1214: vanishes.
1215:
1216: In terms of the function $u$, the boundary of the cluster is given implicitly
1217: by: $u({\bf \hat{r}})=\epsilon T({\bf \hat{r}})$. Therefore, the radial
1218: location at which the $\hat{\rho}_{lm}^{(n-1)}$ vanishes is determined
1219: numerically from:
1220: \begin{equation}
1221: \hat{\rho}_{lm}^{(n-1)}(\hat{r})=\int_0^{2 \pi}\int_{-1}^{1} \hat{\rho}
1222: [u^{(n-1)}(\hat{r},\theta,\phi)-\epsilon T(\hat{r},\theta,\phi)]Y_{lm}
1223: (\theta, \phi) d(cos \theta)d\phi~.
1224: \end{equation}
1225:
1226: In practice, to perform the iteration, the definition of a
1227: grid in spherical coordinates and of a suitable algorithm, in order
1228: to perform the expansion and the resummation in spherical harmonics
1229: of $u$ and $\hat{\rho}$, is required; the number of angular points
1230: of the grid and the maximum harmonic indices $(l,m)$ admitted in
1231: the series (\ref{uexp}) and (\ref{rhoitexp}) are obviously related.
1232:
1233: \section{Conclusions}
1234:
1235: Spherical King models are physically justified models of
1236: quasi-relaxed stellar systems with a truncation radius argued to
1237: ``summarize" the action of an external tidal field. Such simple
1238: models have had great success in representing the structure and
1239: dynamics of globular clusters, even though the presence of the
1240: tidal field is actually ignored. Motivated by these considerations
1241: and by the recent major progress in the observations of globular
1242: clusters, in this paper we have developed a systematic procedure
1243: to construct self-consistent non-spherical models of quasi-relaxed
1244: stellar systems, with special attention to models for which the
1245: non-spherical shape is due to the presence of external tides.
1246:
1247: The procedure developed in this paper starts from a distribution
1248: function identified by replacing, in a reference spherical model,
1249: the single star energy with the relevant Jacobi integral, thus
1250: guaranteeing that the collisionless Boltzmann equation is
1251: satisfied. Then the models are constructed by solving the Poisson
1252: equation, an elliptic partial differential equation with free
1253: boundary. The procedure is very general and can lead to the
1254: construction of several families of non-spherical
1255: equilibrium models. In particular, we have obtained the following
1256: results:
1257:
1258: \begin{itemize}
1259: \item We have constructed models of quasi-relaxed
1260: triaxial stellar systems in which the shape is due to the presence
1261: of external tides; these models reduce to the standard spherical
1262: King models when the tidal field is absent.
1263: \item For these models we have outlined the general properties of
1264: the relevant parameter space; in a separate paper (Varri \&
1265: Bertin, in preparation) we will provide a thorough description of
1266: this two-parameter family of models, also in terms of projected
1267: quantities, as appropriate for comparisons with the observations.
1268: \item We have given a full, explicit solution to two orders in the
1269: tidal strength parameter, based on the method of matched asymptotic
1270: expansions; by comparison with studies of analogous problems in the
1271: theory of rotating polytropic stars, this method appears to be most
1272: satisfactory.
1273: \item We have also discussed two alternative methods of
1274: solution, one of which is based on iteration seeded by the
1275: spherical solution; together with the use of dedicated $N$-body
1276: simulations, the ability to solve such a complex mathematical
1277: problem in different ways will allow us to test the quality of the
1278: solutions in great detail.
1279: \item By suitable change of notation and physical re-interpretation,
1280: the procedure developed in this paper can be applied to the
1281: construction of non-spherical quasi-relaxed stellar systems
1282: flattened by rotation (see Appendix B).
1283: \item The same procedure can also be applied to extend to the triaxial
1284: case other isotropic truncated models (such as low-$n$ polytropes),
1285: that is models that do not reduce to King models in the absence of
1286: external tides (see Appendix C).
1287: \end{itemize}
1288:
1289: We hope that this contribution, in addition to extending the class
1290: of self-consistent models of interest in stellar dynamics, will be
1291: the basis for the development of simple quantitative tools to
1292: investigate whether the observed shape of globular clusters is
1293: primarily determined by internal rotation, by external tides, or
1294: by pressure anisotropy.
1295:
1296: \acknowledgments
1297:
1298: \appendix
1299:
1300: \section{Details of the solution in terms of matched asymptotic expansion}
1301:
1302: \subsection{The general equation}
1303:
1304: From the Taylor expansion about $\epsilon=0$ of the right-hand
1305: side of Eq. (\ref{Poisson}), the structure of the equations for
1306: $\psi_k^{(int)}$ (with $k\ge 2$) can be expressed as:
1307: \begin{equation}\label{gen}
1308: \left[\hat{\nabla}^2+R_1(\hat{r};\Psi)\right]\psi_k^{(int)}=-\left(
1309: \sum_{j=2}^{k}R_j(\hat{r};\Psi)\,X_{k,j}\right)~,
1310: \end{equation}
1311: where $X_{k,j}$ denotes the terms that arise from the
1312: derivatives of $\psi^{(int)}(\hat{r};\epsilon)$ with respect to
1313: $\epsilon$, thus expressed as products of $\psi_i^{(int)}$ (with
1314: $i=1 ,..,k-1$). For fixed $k$ and $j$, the quantity $X_{k,j}$ is
1315: thus a sum of products of $\psi_i^{(int)}$ with subscripts that
1316: are j-part partitions of the integer $k$. Each product of
1317: $\psi_i^{(int)}$ is multiplied by a numerical factor defined as
1318: the ratio between $k!$ and the factorials of the integers that are
1319: parts of the associated partition (if an integer appears $m$ times
1320: in the partition, the factor must also be divided by $m!$). In
1321: particular, for $k=3$ we have:
1322: \begin{equation}
1323: X_{3,2}=3\psi_2^{(int)}\psi_1^{(int)}~{\rm and} \;\;\;\;
1324: X_{3,3}=(\psi_1^{(int)})^3~,
1325: \end{equation}
1326: because the 2-part partition of 3 is $2+1$ and the
1327: 3-part partition is trivially $1+1+1$, thus the relevant equation
1328: is:
1329: \begin{equation}
1330: \left[\hat{\nabla}^2+R_1(\hat{r};\Psi)\right]\psi_3^{(int)}=
1331: -R_2(\hat{r};\Psi)\,3\psi_2^{(int)}\psi_1^{(int)}-R_3(\hat{r};\Psi)\,
1332: (\psi_1^{(int)})^3~.
1333: \end{equation}
1334: Therefore, this formulation of the right-hand side of
1335: the general equation (together with the term $R_1(\hat{r};\Psi)
1336: \psi_k^{(int)}$ on the left-hand side) brings in the {\em Fa\'{a}
1337: di Bruno formula} \citep{Faa1855} for the $k$-th order derivative
1338: of a composite function in which the inner one is expressed as a
1339: series in the variable with respect to which the derivation is
1340: performed.
1341:
1342: \subsection{The equation for the second order radial problem}
1343:
1344: The expansion in spherical harmonics of $(\psi_1^{(in)})^2$, which
1345: involves the product of two spherical harmonics with $l=0$ or
1346: $l=2$ (with $m$ positive and even), can be performed by means of
1347: the so-called 3-j Wigner symbols \footnote{For the definition of
1348: 3-j Wigner symbols and the expression of the harmonic expansion
1349: of the product of two spherical harmonics, see, e.g., \citet{Edm60},
1350: Eqs.~(3.7.3) and (4.6.5), respectively.}. Equation (\ref{ord2lm})
1351: thus corresponds to the following set of six equations:
1352: \begin{eqnarray}
1353: &&\mathcal{D}_0\psi_{2,00}^{(int)}=-R_2(\hat{r};\Psi){1\over
1354: 2\sqrt{\pi}}\;[(\psi_{1,00}^{(int)})^2+(\psi_{1,20}^{(int)})^2+
1355: (\psi_{1,22}^{(int)})^2]~,\\
1356: &&\mathcal{D}_2\psi_{2,20}^{(int)}=-R_2(\hat{r};\Psi){1 \over
1357: 7}\sqrt{{5\over\pi}}\;\left[{7\over
1358: \sqrt{5}}\psi_{1,00}^{(int)}\,\psi_{1,20}^{(int)}+(\psi_{1,20}
1359: ^{(int)})^2-(\psi_{1,22}^{(int)})^2
1360: \right]~,\\
1361: &&\mathcal{D}_2\psi_{2,22}^{(int)}=-R_2(\hat{r};\Psi){1\over
1362: \sqrt{\pi}}\;\left[\psi_{1,00}^{(int)}\,\psi_{1,22}^{(int)}-
1363: {2\sqrt{5}\over 7}\psi_{1,20}^{(int)}\,\psi_{1,22}^{(int)}\right]~,\\
1364: &&\mathcal{D}_4\psi_{2,40}^{(int)}=-R_2(\hat{r};\Psi){1\over
1365: 7\sqrt{\pi}}\;\left[3(\psi_{1,20}^{(int)})^2+\frac{1}{2}(\psi_{1,22}
1366: ^{(int)})^2\right]~,\\
1367: &&\mathcal{D}_4\psi_{2,42}^{(int)}=-R_2(\hat{r};\Psi){1\over
1368: 7}\sqrt{{15\over\pi}}\;\psi_{1,20}^{(int)}\,\psi_{1,22}^{(int)}~,\\
1369: &&\mathcal{D}_4\psi_{2,44}^{(int)}=-R_2(\hat{r};\Psi)\frac{1}{2}
1370: \sqrt{{5\over 7\,\pi}}\;(\psi_{1,22}^{(int)})^2~.
1371: \end{eqnarray}
1372:
1373: \subsection{The asymptotic matching for the first order solution}
1374:
1375: To derive the first order solution, the matching between the pairs
1376: $(\psi^{(int)}, \psi^{(lay)})$ and $(\psi^{(ext)},\psi^{(lay)})$
1377: requires that the internal (external) solution is expanded in a
1378: Taylor series around $\hat{r} = \hat{r}_{tr}$ up to terms of order
1379: $\mathcal{O}((\hat{r} - \hat{r}_{tr})^2)$, expressed with scaled
1380: variables, expanded up to $\mathcal{O}(\epsilon^2)$, and
1381: re-expressed with non-scaled variables:
1382: \begin{eqnarray}\label{intextexp}
1383: &&[\psi^{(~)}]^{(2)}(\hat{r},\theta,\phi)=\psi_0^{(~)}
1384: (\hat{r}_{tr})-{\psi_0^{(~)}}'(\hat{r}_{tr})(\hat{r}_{tr}-\hat{r})
1385: +\frac{1}{2}{\psi_0^{(~)}}^{''}(\hat{r}_{tr})(\hat{r}_{tr}-\hat{r})^2
1386: \nonumber\\ &&+\left[\psi_1^{(~)}(\hat{r}_{tr},\theta,\phi)
1387: -\frac{\partial \psi_1^{(~)}} {\partial
1388: \hat{r}}(\hat{r}_{tr},\theta,\phi)(\hat{r}_{tr}-\hat{r})\right]
1389: \epsilon+\frac{1}{2}\psi_2^{(~)}(\hat{r}_{tr},\theta,\phi)\epsilon^2~.
1390: \end{eqnarray}
1391: Here the closed parentheses include either ``int" or``ext", to denote
1392: internal or external solution, while the notation $[\;]^{(2)}$ on the
1393: left-hand side indicates that a second order expansion in $\epsilon$
1394: has been performed.
1395:
1396: The boundary layer solution in the vicinity of $\eta = 0$ up to
1397: $\mathcal{O}(\eta^2)$, expressed with non-scaled variables and
1398: expanded (formally) up to third order in $\epsilon$ is given by:
1399: \begin{eqnarray}\label{layexp}
1400: && [\psi^{(lay)}]^{(3)}(\hat{r},\theta,\phi)=F_0(\theta,\phi)(\hat{r}_{tr}-\hat{r})
1401: +\frac{F_0(\theta,\phi)}{\hat{r}_{tr}}
1402: (\hat{r}_{tr}-\hat{r})^2 \nonumber\\
1403: &&+[G_0(\theta,\phi)+F_1(\theta,\phi)
1404: (\hat{r}_{tr}-\hat{r})]\epsilon+G_1(\theta,\phi)\epsilon^2~.
1405: \end{eqnarray}
1406: \noindent By equating equal powers of $\epsilon$ and
1407: $(\hat{r}_{tr}-\hat{r})$ in (\ref{intextexp}) and (\ref{layexp}),
1408: we find:
1409: \begin{eqnarray}
1410: \psi^{(int)}_0(\hat{r}_{tr})&=\;\;\;\;\;0\;\;\;\;\;=&\alpha_0-
1411: \frac{\lambda_0}{\hat{r}_{tr}}~, \label{match0}\\
1412: {-\psi^{(int)}_0}^{'}(\hat{r}_{tr})&=F_0(\theta,\phi)=&-
1413: \frac{\lambda_0}{\hat{r}_{tr}^2}~, \label{match1}\\
1414: \frac{1}{2}{\psi^{(int)}_0}^{''}(\hat{r}_{tr})&\displaystyle =
1415: \frac{F_0(\theta,\phi)}{\hat{r}_{tr}}=&-\frac{\lambda_0}
1416: {\hat{r}_{tr}^3}~, \label{match2}\\
1417: \psi^{(int)}_1(\hat{r}_{tr},\theta,\phi)&=G_0(\theta,\phi)=&
1418: \alpha_1-\frac{\lambda_1}{\hat{r}_{tr}}-\frac{T_{00}(\hat{r}_{tr})}
1419: {2\sqrt \pi}\nonumber\\ &&-\sum_{l=1}^{\infty}\sum_{m=-l}^{l}\left[\frac{a_{lm}}
1420: {\hat{r}^{l+1}_{tr}}+T_{lm}(\hat{r}_{tr})\right]Y_{lm}(\theta,\phi)~, \label{match3}\\
1421: -\frac{\partial \psi_1^{(int)}}{\partial\hat{r}}(\hat{r}_{tr},\theta,\phi)&=
1422: F_1(\theta,\phi)=&-\frac{\lambda_1}{\hat{r}^2_{tr}}+\frac{T_{00}(\hat{r}_{tr})}
1423: {\sqrt \pi\hat{r}_{tr}}\nonumber\\ &&-\sum_{l=1}^{\infty}\sum_{m=-l}^{l}\left[\frac{(l+1)a_{lm}}
1424: {\hat{r}^{l+2}_{tr}}-\frac{2T_{lm}(\hat{r}_{tr})}{\hat{r}_{tr}}\right]Y_{lm}(\theta,\phi)~,
1425: \label{match4}\\
1426: \frac{1}{2}\psi^{(int)}_2(\hat{r}_{tr},\theta,\phi)&=G_1(\theta,\phi)=&\frac{1}{2}
1427: \left[ \alpha_2-\frac{\lambda_2}{\hat{r}_{tr}}-\sum_{l=1}^{\infty}
1428: \sum_{m=-l}^{l}\frac{b_{lm}}{\hat{r}_{tr}^{l+1}}Y_{lm}(\theta,\phi)\right]~,
1429: \label{match5}
1430: \end{eqnarray}
1431: where the equalities on the left-hand side arise from
1432: the matching of the pair $(\psi^{(int)},\psi^{(lay)})$ and
1433: identify the free angular functions (\ref{F0})-(\ref{G1}), while
1434: those on the right-hand side arise from the matching of the pair
1435: $(\psi^{(ext)}, \psi^{(lay)})$. We also note that (\ref{match0})
1436: is consistent with the definition of the truncation radius and
1437: that (\ref{match2}) is equivalent to (\ref{match1}), because from
1438: Eq.~(\ref{ord0}) we have:
1439: ${\psi_0^{(int)}}^{''}(\hat{r}_{tr})=-(2/\hat{r}_{tr}){\psi_0^{(int)}}^{'}
1440: (\hat{r}_{tr})$.
1441:
1442: The free constants (\ref{alpha0})-(\ref{lambda1}) are thus easily
1443: determined. For a given harmonic $(l,m)$ of Eqs.~(\ref{match3})
1444: and (\ref{match4}), with $l\ge 1$, the constants $A_{lm}$ and
1445: $a_{lm}$ are governed by the linear system with $i,j=1,2$:
1446: \begin{equation}\label{system}
1447: M_{ij} u_j= v_i~,
1448: \end{equation}
1449: where the matrix $M$ is given by
1450: \begin{equation}
1451: M = \left(
1452: \begin{array}{cc}
1453: \gamma_l(\hat{r}_{tr})&\hat{r}_{tr}^{-(l+1)}\\
1454: -\gamma_l'(\hat{r}_{tr})\hat{r}_{tr}&(l+1)\hat{r}_{tr}^{-(l+1)}
1455: \end{array}
1456: \right)~,
1457: \end{equation}
1458: and the vectors are defined as $(u_1,u_2)=(A_{lm},a_{lm})$ and
1459: $(v_1,v_2)=-T_{lm}(\hat{r}_{tr})(1,-2)$. Such linear system
1460: (\ref{system}) is well posed, i.e. $det M \ne 0$.
1461:
1462: To show this, we may integrate Eq~(\ref{ord1lm}) for the regular
1463: solution, under the conditions $\gamma_l(0)=\gamma'_l(0)=0$:
1464: \begin{equation}
1465: \gamma_l'(\hat{r})\hat{r}^2=\int_0^{\hat{r}}[l(l+1)-R_1(\hat{r};\Psi)\hat{r}^2]\gamma_l(\hat{r})d
1466: \hat{r}~.
1467: \end{equation}
1468: In the vicinity of $\hat{r}=0$ the quantity $R_1(\hat{r};\Psi)
1469: \hat{r}^2$ is vanishingly small, so that, if the quantity
1470: $R_1(\hat{r};\Psi)\hat{r}^2$ remains smaller than $l(l+1)$, then
1471: the regular solution, starting positive and monotonic, remains a
1472: positive and monotonically increasing function of $\hat{r}$.
1473: Indeed, we have checked that this condition occurs for $l \ge 2$,
1474: because for $\Psi \in [0.5,10]$ the quantity
1475: $R_1(\hat{r};\Psi)\hat{r}^2$ has a maximum value in the range
1476: [4.229, 3.326]. Under this condition the function $\mu_l(\hat{r})
1477: \equiv \gamma_l'(\hat{r}) \hat{r} +(l+1)\gamma_l(\hat{r})$ cannot
1478: change sign, so that $det M =
1479: \hat{r}_{tr}^{-(l+1)}\mu_l(\hat{r}_{tr})$ cannot vanish. This
1480: argument does not work for the case $l=1$, in which the function
1481: $\mu_l(\hat{r})$ does change sign at a point
1482: $0<\hat{r}_0<\hat{r}_{tr}$, but we have checked directly that the
1483: property $det M \neq 0$ is satisfied also in this case.
1484:
1485: Then expressions (\ref{A2m}) and (\ref{a2m}) are easily recovered
1486: and we conclude that for the harmonics that are not ``driven''by
1487: the tidal potential $T({\bf \hat{r}})$ the related $a_{lm}$ and
1488: $A_{lm}$ must vanish.
1489:
1490: Similarly, for a given harmonic $(l,m)$ with $l\ge 1$,
1491: Eq.~(\ref{F2}), which is obtained from the
1492: second order matching, and Eq.~(\ref{match5}) can be cast in the form of
1493: Eq.~(\ref{system}), with $(u_1,u_2)=(B_{lm},b_{lm})$ and
1494: $(v_1,v_2)=(-g_{lm} (\hat{r}_{tr}),\hat{r}_{tr}
1495: g'_{lm}(\hat{r}_{tr}))$. Therefore, the argument provided above
1496: applies and we can conclude that for those harmonics for which the
1497: particular solutions $g_{lm}(\hat{r})$ are absent (or,
1498: equivalently, Eq.~[\ref{ord2lm}] is homogeneous), the constants
1499: $B_{lm}$ and $b_{lm}$ must vanish. A linear system equivalent to
1500: (\ref{system}) can be written for a fixed harmonic $(l,m)$ with
1501: $l\ge 1$ of the solution of general order $k$ and the same
1502: argument applies. Therefore, we conclude that the $k$-th order
1503: term of the solution contributes only to those harmonics for which
1504: the particular solutions are present.
1505:
1506: \subsection{The structure of $k$-th order term}
1507:
1508: Because we have noted (see argument introduced about
1509: the system [\ref{system}]) that the $k$-th order term
1510: of the solution has non-vanishing contribution only in
1511: correspondence of those harmonics for which the component of
1512: Eq.~(\ref{gen}) is non-trivial, the discussion about the structure
1513: of the term reduces to the analysis of the structure of the
1514: expansion in spherical harmonics of the right-hand side of that
1515: equation. Recalling that the harmonic expansion of the product of
1516: two spherical harmonics $(l_1,m_1)$ and $(l_2,m_2)$ can be
1517: expressed by means of 3-j Wigner symbols \citetext{see \citealp{Edm60},
1518: Eq.~[4.6.5]}, we note that the composed harmonic $(l,m)$ must
1519: satisfy the following {\em selection rules}: (i) $|l_1-l_2| \le l
1520: \le l_1+l_2$ (``triangular inequality"), (ii) $m_1+m_2=m$, and
1521: (iii) $l_1+l_2+l$ must be even. The last condition holds because
1522: in the cited expression the composed harmonic appears multiplied
1523: by the special case of the Wigner symbol with $(l_1,l_2,l)$ as
1524: first row and $(0,0,0)$ as second row. Bearing in mind that the
1525: first order term is characterized by harmonics with $l=0,2$ and
1526: corresponding positive and even values of $m$ and that the
1527: structure of the right-hand side of Eq.~(\ref{gen}) can be
1528: interpreted by means of the partitions of the integer $k$, it can
1529: be proved {\em by induction} that the $k$-th order term is
1530: characterized by harmonics with $l=0,2,...,2k$ and corresponding
1531: positive and even values of $m$.
1532:
1533: \section{Extension to the presence of internal rotation}
1534:
1535: It is well known that in the presence of finite total angular
1536: momentum of the system, relaxation leads to solid-body rotation
1537: (e.g., see Landau \& Lifchitz 1967). If we denote by $\omega$ the angular
1538: velocity of such rigid rotation and assume that it takes place
1539: around the $z$ axis, in the statistical mechanical argument that
1540: leads to the derivation of the Maxwell-Boltzmann distribution one
1541: finds that in the final distribution function the single particle
1542: energy $E$ is replaced by the quantity $E - \omega J_z$. Following
1543: this picture, we may consider the extension of King models to the
1544: case of internal rigid rotation. This extension is conceptually
1545: simpler than that addressed in the main text of this paper,
1546: because the perturbation associated with internal rotation, while
1547: breaking spherical symmetry, preserves axial symmetry. We note
1548: that the models described below differ from those studied by
1549: \citet{KorAna71}, which were characterized by a different,
1550: discontinuous truncation, and those by \citet{PreTom70}
1551: and by \citet{Wil75}, which were characterized by a different
1552: truncation prescription and by differential rotation.
1553:
1554: The relevant physical model is that of a rigidly rotating isolated
1555: globular cluster characterized by angular velocity
1556: $\mbox{\boldmath$\omega$}=\omega \hat{e}_z$, with respect to a
1557: frame of reference with the origin in the center of mass of the
1558: cluster. We then introduce a second frame of reference,
1559: co-rotating with the cluster, in which the position vector is
1560: given by ${\bf r}=(x,y,z)$. In such rotating frame, the Lagrangian
1561: describing the motion of a star belonging to the cluster is given
1562: by:
1563: \begin{equation}
1564: \mathcal{L}=\frac{1}{2}(\dot{x}^2+\dot{y}^2+\dot{z}^2+2\omega\dot{y}x-
1565: 2\omega\dot{x}y)-\Phi_{cen}(x,y)-\Phi_C(x,y,z)~,
1566: \end{equation}
1567: \noindent where $\Phi_{cen}(x,y)=-(x^2+y^2)\omega^2/2$ is the
1568: centrifugal potential; the energy integral of the motion (called
1569: the Jacobi integral) is:
1570: \begin{equation}
1571: H=\frac{1}{2}(\dot{x}^2+\dot{y}^2+\dot{z}^2)+\Phi_{cen}+\Phi_C~.
1572: \end{equation}
1573: As in the tidal case, the extension of King models is performed by
1574: considering the distribution function $f_K(H)$, as in (\ref{fK}),
1575: for $H \le H_0$ and $f_K(H)=0$ otherwise, with $H_0$ the cut-off
1576: constant. The dimensionless energy is defined by:
1577: \begin{equation}
1578: \psi({\bf r})=a\{H_0-[\Phi_C({\bf r})+\Phi_{cen}(x,y)]\}~,
1579: \end{equation}
1580: \noindent and the boundary of the cluster, implicitly defined as
1581: $\psi({\bf r})=0$, is an equipotential surface for the total
1582: potential $\Phi_C+\Phi_{cen}$. Its geometry, reflecting the
1583: properties of the centrifugal potential, is characterized by
1584: symmetry with respect to the $z$-axis and reflection symmetry with
1585: respect to the equatorial plane $(x,y)$. The constant $\psi$
1586: family of surfaces, much like the Hill surfaces of the tidal case,
1587: is characterized by a {\em critical} surface which distinguishes
1588: the closed from the opened ones and in which the points on the
1589: equatorial plane are all saddle points. In these points the
1590: centrifugal force balances the self-gravity of the cluster; their
1591: distance from the origin, which we call {\em break-off radius}
1592: ($r_B$), can be determined from the condition:
1593: \begin{equation}\label{breakoffradius}
1594: \frac{\partial \psi}{\partial x}(r_B,0,0)=\frac{\partial
1595: \psi}{\partial y} (0,r_B,0)=0~.
1596: \end{equation}
1597: Following the argument that we gave in Sect.~2.2 for the
1598: tidal radius, we find a zero-th order approximation for the
1599: break-off radius: $r_B^{(0)}=(GM/\omega^2)^{1/3}$, where $M$ is
1600: the total mass of the cluster. The discussion about the parameter
1601: space that characterizes these models is equivalent to the one
1602: presented in Sect.~2.3, with:
1603: \begin{equation}
1604: \chi \equiv \frac{\omega^2}{4 \pi G \rho_0}~,
1605: \end{equation}
1606: a dimensionless parameter that measures the strength of
1607: the rotation with respect to the central density of the cluster,
1608: playing the role of $\epsilon$. Similarly, we may define an
1609: extension parameter $\sigma=r_{tr}/r_B$ (instead of $\delta$, see
1610: Sect.~2.3). For every value of the dimensionless central potential
1611: well $\Psi$, a maximum value for the strength parameter,
1612: $\chi_{cr}$, exists, corresponding to the critical condition in
1613: which the boundary of the cluster is given by the critical
1614: equipotential surface.
1615:
1616: The relevant equations in the construction of these (fully)
1617: self-consistent models can be expressed in dimensionless form by
1618: means of the same rescaling of variables performed in Sect.~3.
1619: Thus, for $\psi \ge 0$, the Poisson equation can be written as:
1620: \begin{equation}\label{Poissonrot}
1621: \hat{\nabla}^2
1622: \psi=-9\left[\frac{\hat{\rho}(\psi)}{\hat{\rho}(\Psi)}-2\chi
1623: \right]~,
1624: \end{equation}
1625: where $\hat{\rho}$ is defined as in (\ref{rho}). For
1626: negative values of $\psi$ we should refer to:
1627: \begin{equation}\label{vacuumrot}
1628: \hat{\nabla}^2 \psi=18 \chi~,
1629: \end{equation}
1630: \noindent with the boundary conditions at the origin written as in
1631: (\ref{cond1})-(\ref{cond2}) and at large radii given by $\psi+\chi
1632: C \rightarrow aH_0$, where $C \equiv -(9/2)(\hat{x}^2+\hat{y}^2)$
1633: is the dimensionless centrifugal potential.
1634:
1635: The solution up to second order in terms of matched asymptotic
1636: expansions presented in Sect.~4 can be adapted to this case
1637: without effort. In fact, with respect to the calculation presented
1638: in the main text only two differences occur: (i) wherever the
1639: constant term $-\hat{\nabla}^2T=-9(1-\nu)$ appears, it must be
1640: replaced here by $-\hat{\nabla} ^2C=18$ (the sign is the same in
1641: the two cases, because $1-\nu < 0$, see Sect. 2.1), and (ii)
1642: thanks to axisymmetry, in the angular part of the Laplacian the
1643: derivative with respect to the toroidal angle $\phi$ can be
1644: dropped and thus the terms of the asymptotic series
1645: (\ref{asymserin})- (\ref{asymserext}) can be expanded by means of
1646: Legendre polynomials\footnote{Following \citet{AbrSte}, we use
1647: Legendre polynomials as defined in (22.3.8), i.e. with Condon-Shortley
1648: phase, and normalized with respect to the relation (22.2.10). We
1649: remark that, although they are structurally equivalent to zonal
1650: spherical harmonics, the normalization is different.} instead of
1651: spherical harmonics. The latter property implies that the radial
1652: part of each term of the asymptotic series is characterized by
1653: only one index, $l$, i.e. we can write $\psi_{k,l}^{(\;)}$. We
1654: note that the differential operator that appears on the left-hand
1655: side of the relevant equations for the solution defined in the
1656: internal region is still $\mathcal{D}_l$, and thus also the functions
1657: $\gamma_l(\hat{r})$ can be introduced in the same way as before.
1658: As to the equations corresponding to (\ref{tau0})-(\ref{tau2}),
1659: the formal solutions of the equations in the boundary layer, the
1660: angular functions $F_i$ and $G_i$ (with $i=0,1,2$) now depend only
1661: on the poloidal angle $\theta$. About the external solution, an
1662: expression analogous to (\ref{extgen}) can be used, with the
1663: particular solution given by $\chi C$ instead of $\epsilon T$.
1664: The centrifugal potential contributes, as in the case of the
1665: tidal potential, only with monopole and quadrupole terms,
1666: explicitly:
1667: \begin{eqnarray}
1668: &&C_{0}(\hat{r})=-3\sqrt 2 \hat{r}^2 \label{C0}~,\\
1669: &&C_{2}(\hat{r})=3\sqrt{ \frac{2}{5}}\hat{r}^2 \label{C2}~.
1670: \end{eqnarray}
1671: Finally, as a result of the matching of the pair $(\psi^{(int)},
1672: \psi^{(lay)})$ up to second order, the expressions for the angular
1673: functions (\ref{F0})-(\ref{G1}) and (\ref{F2})-(\ref{G2}) are
1674: still applicable. In addition, from the matching of the pair
1675: $(\psi^{(lay)},\psi^{ext})$ up to second order, we find that the
1676: explicit expressions for the free constants follow
1677: Eqs.~(\ref{alpha0})-(\ref{a2m}) and (\ref{alpha2})-(\ref{b4m}),
1678: provided that we drop everywhere the index $m$ and we replace $3
1679: T_{00}(\hat{r}_{tr})/(2\sqrt \pi)$ with $3 C_0
1680: (\hat{r}_{tr})/\sqrt 2$ in (\ref{alpha1}) and
1681: $T_{00}(\hat{r}_{tr})/ (\sqrt \pi \hat{r}_{tr})$ with $2
1682: C_0(\hat{r}_{tr})/(\sqrt2 \hat{r}_{tr})$ in (\ref{lambda1}).
1683: Obviously, the particular solutions $f_0$ and $g_{l}$ (with
1684: $l=0,2,4$) are different from the ones obtained in the tidal case,
1685: because the right-hand side of the relevant equations is
1686: different. Also in this case, it can be proved {\it by induction}
1687: that the $k$-th order term has non-vanishing contributions only
1688: for $l=0,2,...,2k$.
1689:
1690: For completeness, we record the explicit expression of the second
1691: order equations in the internal region:
1692: \begin{eqnarray}
1693: &&\mathcal{D}_0\psi_{2,0}^{(int)}=-R_2(\hat{r};\Psi){1\over\sqrt{2}}\;
1694: [(\psi_{1,0}^{(int)})^2+(\psi_{1,2}^{(int)})^2]~,\\
1695: &&\mathcal{D}_2\psi_{2,2}^{(int)}=-R_2(\hat{r};\Psi){\sqrt{10}
1696: \over7}\;
1697: \left[{7\over\sqrt{5}}\psi_{1,0}^{(int)}\,\psi_{1,2}^{(int)}+(\psi_{1,2}
1698: ^{(int)})^2\right]~,\\
1699: &&\mathcal{D}_4\psi_{2,4}^{(int)}=-R_2(\hat{r};\Psi){3 \sqrt
1700: 2\over7}\; (\psi_{1,2}^{(int)})^2~.
1701: \end{eqnarray}
1702: \noindent We remark that the Legendre expansion of the product of
1703: two Legendre polynomials is straightforward, because the 3-j
1704: Wigner symbols of interest all belong to the special case with
1705: $(0,0,0)$ as second row.
1706:
1707:
1708: \section{Extension of other isotropic truncated models}
1709:
1710: The procedure developed in Sects.~3 and 4 can be applied also to
1711: extend other isotropic truncated models, different from the King
1712: models, to the case of tidal distortions. Here we briefly describe
1713: the case of low-$n$ polytropes ($1<n<5$), which are particularly
1714: well suited to the purpose, because they are characterized by a
1715: very simple analytical expression for the density as a function of
1716: the potential; this class of models was also considered by \citet{Wei93}.
1717: In the distribution function that defines the polytropes
1718: \citep[e.g., see][]{Ber00}, we may thus replace the single star energy
1719: with the Jacobi integral (see definition [\ref{Jacobi}]) and consider:
1720: \begin{equation}\label{fP}
1721: f_P(H)=A(H_0-H)^{n-3/2}~,
1722: \end{equation}
1723: for $H \le H_0$, and a vanishing distribution otherwise. Unlike the
1724: King family discussed in the main text, these models have no
1725: dimensionless parameter to measure the concentration of the stellar
1726: system, which depends only on the polytropic index $n$; in fact, the
1727: spherical fully self-consistent polytropes are characterized only by
1728: two physical scales, which are associated with the cut-off constant
1729: $H_0$ and the normalization factor $A$. Below we consider values of
1730: $n<5$, so that the models have finite radius.
1731: Therefore, the relevant parameter space for the tidally distorted models
1732: is represented just by the tidal strength parameter $\epsilon$ (see
1733: definition [\ref{tidalp}]), which, for a given value of the index $n$,
1734: has a (maximal) critical value. The definition (\ref{delta}) for the
1735: extension parameter $\delta$ is still valid if by $r_{tr}$ we denote
1736: the radius of the unperturbed spherical configuration. The associated
1737: density functional is given by:
1738: \begin{equation}
1739: \rho(\psi)=\rho_0 \psi^n~,
1740: \end{equation}
1741: where the dimensionless escape energy is given by:
1742: \begin{equation}
1743: \psi({\bf r})=\{H_0-[\Phi_C({\bf r})+\Phi_T({\bf r})] \}\left(\frac{c_n}
1744: {\rho_0}\right)^{1/n}~,
1745: \end{equation}
1746: with $c_n\equiv(2 \pi)^{3/2}\Gamma(n-1/2)A/n!$. The boundary of the
1747: perturbed configuration is defined by $\psi({\bf r})=0$, following the
1748: same arguments described in the main text. Here $\rho_0$ can be
1749: interpreted as the central density if we set $\psi({\bf 0})=1$.
1750:
1751: For $\psi \ge 0$, the relevant equation for the construction of the
1752: self-consistent tidally distorted models is the Poisson equation, which,
1753: in dimensionless form, is given by:
1754: \begin{equation}\label{Poissonpoly}
1755: \hat{\nabla}^2 \psi=-[\psi^n-\epsilon (1-\nu)]~,
1756: \end{equation}
1757: while for negative values of $\psi$ we must refer to Eq.~(\ref{vacuum}).
1758: Here the rescaling of variables has been performed by means of
1759: the scale length $\zeta\equiv \sqrt{\rho_0^{1/n-1}/(4 \pi G c_n^{1/n})}$.
1760: The relevant boundary conditions are given by $\psi({\bf 0})=1$ instead
1761: of (\ref{cond1}), while (\ref{cond2}) and (\ref{cond3}) hold unchanged.
1762:
1763: If the polytropic index is in the range $1 < n < 5$, the solution up to
1764: second order presented in Sect.~4 is fully applicable, provided that we
1765: note that the problem for the zero-th order term of the series (\ref{asymserin})
1766: is now given by the Lane-Emden equation \citep[see, e.g.][]{Cha39}:
1767: \begin{equation}\label{LaneEmden}
1768: {\psi_0^{(int)}}^{''}+\frac{2}{\hat{r}}{\psi_0^{(int)}}'=-\left(\psi_0^
1769: {(int)}\right)^n~,
1770: \end{equation}
1771: with $\psi_0^{(int)}(0)=1$ and ${\psi_0^{(int)}}'(0)= 0$, where the symbol
1772: $'$ denotes derivative with respect to the argument $\hat{r}$; explicitly,
1773: the truncation radius $\hat{r}_{tr}$ is now defined by $\psi_0^{(int)}
1774: (\hat{r}_{tr})=0$, i.e. it represents the radius of the so-called
1775: {\em Emden sphere}. Correspondingly, the quantities called $R_j$
1776: in the main text must be re-defined as:
1777: \begin{equation}
1778: R_j(\hat{r};n) \equiv \frac{d^j\psi^n}{d\psi^j} \biggl\vert_{\psi_0^{(int)}}~;
1779: \end{equation}
1780: the value of $j$ at which the quantity $R_j$ may start to diverge depends
1781: on the index $n$. Obviously, in Eq.~(\ref{Poislay}), i.e. in the Poisson
1782: equation defined in the boundary layer, $\hat{\rho}(\epsilon \tau)$ must
1783: be replaced by $(\epsilon \tau)^n$. This makes it clear that the value of
1784: the polytropic index $n$ directly affects the order, with respect to the
1785: perturbation parameter, at which the density contribution on the right-hand
1786: side of Eq.~(\ref{Poislay}) comes into play and therefore changes the matching
1787: procedure. If $n >1$, the density contribution emerges only after the second
1788: order and thus the full procedure described in Sect.~4 is valid. In contrast,
1789: if $n \le 1$ the procedure described in the main text is applicable only up to
1790: first order while the calculation of second order terms would require a
1791: re-definition of the boundary layer (as it happens for the case discussed in
1792: the main text when terms of order $k >3$ are desired).
1793: In closing, we note that the procedure presented in this paper can be applied
1794: also to isotropic truncated models with more complicated expressions for the
1795: density functional (e.g. the family of models $f_{1n}$ proposed by \citet{Dav77},
1796: without boundary conditions on tangential velocity, for which the density
1797: functionals are expressed in terms of the error function and of the Dawson
1798: integral), bearing in mind the last {\em caveat} about the possibility that
1799: the density contribution may affect at some order the boundary layer, thus
1800: requiring a reformulation of the results presented in Sect.~4.
1801:
1802:
1803:
1804: \begin{thebibliography}{}
1805: \bibitem[Abramowitz \& Stegun(1964)]{AbrSte} Abramowitz, M., Stegun, I.\ 1964, Handbook of Mathematical Functions (Mineola, NY: Dover Publications)
1806: \bibitem[Arena \& Bertin(2007)]{AreBer07} Arena, S.~E., \& Bertin, G.\ 2007, \aap, 463, 921
1807: \bibitem[Attico \& Pegoraro(1999)]{AttPeg99} Attico, N., \& Pegoraro, F.\ 1999, Phys. Pl., 6, 767
1808: \bibitem[Bender\& Orszag(1999)]{BenOrs99} Bender, C.M., \& Orszag, S.A.\ 1999, Advanced Mathematical Methods for Scientists and Engineers (Berlin: Springer-Verlag)
1809: \bibitem[Bertin \& Stiavelli(1993)]{BerSti93} Bertin, G., \& Stiavelli, M.\ 1993, Reports of Progress in Physics, 56, 493
1810: \bibitem[Bertin(2000)]{Ber00} Bertin, G.\ 2000, Dynamics of Galaxies (Cambridge, UK: Cambridge University Press)
1811: \bibitem[Bertin et al.(2003)]{Ber03} Bertin, G., Liseikina, T., \& Pegoraro 2003, \aap, 405, 73
1812: \bibitem[Bontekoe \& van Albada(1987)]{BonAlb87} Bontekoe, T.~R., \& van Albada, T.~S.\ 1987, \mnras, 224, 349
1813: \bibitem[Chandrasekhar(1933)]{Cha33} Chandrasekhar, S.\ 1933, \mnras, 93, 390
1814: \bibitem[Chandrasekhar(1939)]{Cha39} Chandrasekhar, S.\ 1939, An Introduction to the Study of Stellar Structure (Chicago, IL: University of Chicago Press)
1815: \bibitem[Chandrasekhar(1942)]{Cha42} Chandrasekhar, S.\ 1942, Principles of Stellar Dynamics (Chicago, IL: University of Chicago Press)
1816: \bibitem[Chandrasekhar \& Lebovitz(1962)]{ChaLeb62} Chandrasekhar, S., \& Lebovitz, N.~R.\ 1962, \apj, 136, 1082
1817: \bibitem[Davoust(1977)]{Dav77} Davoust, E.\ 1977, \aap, 61, 391
1818: \bibitem[Davoust \& Prugniel(1990)]{DavPru90} Davoust, E., \& Prugniel, P.\ 1990, \aap, 230, 67
1819: \bibitem[Djorgovski \& Meylan(1994)]{DjoMey94} Djorgovski, S., \& Meylan, G.\ 1994, \aj, 108, 1292
1820: \bibitem[Edmonds(1960)]{Edm60} Edmonds, A.R.\ 1960, Angular Momentum in Quantum Mechanics (Second ed.; Princeton, NJ: Princeton University Press)
1821: \bibitem[Ernst et al.(2008)]{Ern08} Ernst, A., Just, A., Spurzem, R., \& Porth, O.\ 2008, \mnras, 383, 897
1822: \bibitem[Fa\'{a} di Bruno(1855)]{Faa1855} Fa\'{a} di Bruno, C. F.\ 1855, Ann. di Scienze Matem. et Fisiche di Tortoloni 6, 479
1823: \bibitem[Frenk \& Fall(1982)]{FreFal82} Frenk, C.~S., \& Fall, S.~M.\ 1982, \mnras, 199, 565
1824: \bibitem[Geyer et al.(1983)]{Gey83} Geyer, E.~H., Nelles, B., \& Hopp, U.\ 1983, \aap, 125, 359
1825: \bibitem[Goodwin(1997)]{Goo97} Goodwin, S.~P.\ 1997, \mnras, 286, L39
1826: \bibitem[Han \& Ryden(1994)]{HanRyd94} Han, C., \& Ryden, B.~S.\ 1994, \apj, 433, 80
1827: \bibitem[Harris(1962)]{Har62} Harris, E.~G.\ 1962, Il Nuovo Cimento, 23, 115
1828: \bibitem[Heggie \& Hut(2003)]{HegHut03} Heggie, D., \& Hut, P.\ 2003, The Gravitational Million-Body Problem (Cambridge, UK: Cambridge University Press)
1829: \bibitem[Heggie \& Ramamani(1995)]{HegRam95} Heggie, D.~C., \& Ramamani, N.\ 1995, \mnras, 272, 317
1830: \bibitem[King(1961)]{Kin61} King, I.~R.\ 1961, \aj, 66, 68
1831: \bibitem[King(1965)]{Kin65} King, I.~R.\ 1965, \aj, 70, 376
1832: \bibitem[King(1966)]{Kin66} King, I.~R.\ 1966, \aj, 71, 64
1833: \bibitem[Kormendy \& Anand(1971)]{KorAna71} Kormendy, J., \& Anand, S.~P.~S.\ 1971, \apss, 12, 47
1834: \bibitem[Krogdahl(1942)]{Kro42} Krogdahl, W.\ 1942, \apj, 96, 124
1835: \bibitem[Lagoute \& Longaretti(1996)]{LagLon96} Lagoute, C., \& Longaretti, P.-Y.\ 1996, \aap, 308, 441
1836: \bibitem[Landau \& Lifchitz(1967)]{LanLif67} Landau, L. \& Lifchitz, E.M.,\ 1967, Physique Statistique (Moscou: MIR)
1837: \bibitem[Longaretti \& Lagoute(1996)]{LonLag96} Longaretti, P.-Y., \& Lagoute, C.\ 1996, \aap, 308, 453
1838: \bibitem[Lynden-Bell(1967)]{Lyn67} Lynden-Bell, D.\ 1967, \mnras, 136, 101
1839: \bibitem[Madsen(1996)]{Mad96} Madsen, J.\ 1996, \mnras, 280, 1089
1840: \bibitem[McLaughlin \& van der Marel(2005)]{McLMar05} McLaughlin, D.~E., \& van der Marel, R.~P.\ 2005, \apjs, 161, 304
1841: \bibitem[McLaughlin et al.(2006)]{McL06} McLaughlin, D.~E., Anderson, J., Meylan, G., Gebhardt, K., Pryor, C., Minniti, D.,
1842: \& Phinney, S.\ 2006, \apjs, 166, 249
1843: \bibitem[Milne(1923)]{Mil23} Milne, E. A.\ 1923, \mnras, 83, 118
1844: \bibitem[Monaghan \& Roxburgh(1965)]{MonRox65} Monaghan, F.~F., \& Roxburgh, I.~W.\ 1965, \mnras, 131, 13
1845: \bibitem[Prendergast \& Tomer(1970)]{PreTom70} Prendergast, K.~H., \& Tomer, E.\ 1970, \aj, 75, 674
1846: \bibitem[Ryden(1996)]{Ryd96} Ryden, B.~S.\ 1996, \apj, 461, 146
1847: \bibitem[Schwarzschild(1979)]{Sch79} Schwarzschild, M.\ 1979, \apj, 232, 236
1848: \bibitem[Smith(1975)]{Smi75} Smith, B.~L.\ 1975, \apss, 35, 223
1849: \bibitem[Smith(1976)]{Smi76} Smith, B.~L.\ 1976, \apss, 43, 411
1850: \bibitem[Spitzer(1987)]{Spi87} Spitzer, L.\ 1987, Dynamical Evolution of Globular Clusters (Princeton, NJ: Princeton University Press)
1851: \bibitem[Tassoul(1978)]{Tas78} Tassoul, J.-L.\ 1978, Theory of rotating stars (Princeton, NJ: Princeton University Press)
1852: \bibitem[van Albada(1982)]{Alb82} van Albada, T.~S.\ 1982, \mnras, 201, 939
1853: \bibitem[van den Bergh(2008)]{Ber08} van den Bergh, S.\ 2008, \aj, 135, 1731
1854: \bibitem[van de Ven et al.(2006)]{Ven06} van de Ven, G., van den Bosch, R.~C.~E., Verolme, E.~K., \& de Zeeuw, P.~T.\ 2006, \aap, 445, 513
1855: \bibitem[Van Dyke(1975)]{Dyk75} Van Dyke, M.\ 1975, Perturbation Methods in Fluid Mechanics (Annotated ed.; Stanford, CA: The Parabolic Press)
1856: \bibitem[van Leeuwen et al.(2000)]{Lee00} van Leeuwen, F., Le Poole, R.~S., Reijns, R.~A., Freeman, K.~C., \& de Zeeuw, P.~T.\ 2000, \aap, 360, 472
1857: \bibitem[Weinberg(1993)]{Wei93} Weinberg, M.~D.\ 1993, in ASP Conf. Ser. 48, The Globular Cluster-Galaxy Connection, ed. G.~H. Smith, \& J.~P. Brodie, (San Francisco: ASP), 689
1858: \bibitem[White \& Shawl(1987)]{WhiSha87} White, R.~E., \& Shawl, S.~J.\ 1987, \apj, 317, 246
1859: \bibitem[Wilson(1975)]{Wil75} Wilson, C.~P.\ 1975, \aj, 80, 175
1860: \bibitem[Woolley \& Dickens(1962)]{WooDic62} Woolley, R.v.d.R, \& Dickens, R.~J.\ 1962, ROE Bull., 54
1861: \end{thebibliography}
1862:
1863:
1864: \clearpage
1865:
1866: \begin{figure}
1867: \epsscale{0.5}
1868: \plotone{f1.eps}
1869: \caption{The critical Hill surface (in dimensionless variables) for a
1870: second-order model with $\Psi=2$ and $\epsilon =7.043\times 10^{-4}$
1871: (corresponding to $\delta_{cr}^{(2)}=0.671$); the galactic potential
1872: is Keplerian ($\nu=3$).\label{fig1}}
1873: \end{figure}
1874:
1875: \begin{figure}
1876: \epsscale{1.00}
1877: \plotone{f2.eps}
1878: \caption{Sections in the three coordinate planes for a second-order
1879: model with $\Psi=2$ and $\epsilon =7.000\times 10^{-4}$, characterized
1880: by high tidal distortion ($\delta=0.669\approx \delta_{cr}^{(2)}$, see
1881: Fig. \ref{fig1}) illustrating the boundary surface of the triaxial model (solid),
1882: of the internal region (dotted), and of the corresponding spherical King
1883: model (dashed); the filled area represents the inner region of the boundary
1884: layer. Note the compression and the elongation with respect to the unperturbed
1885: configuration in the $\hat{z}$- and $\hat{x}$-direction, respectively. The
1886: galactic potential is Keplerian ($\nu=3$).}
1887: \label{fig2}
1888: \end{figure}
1889:
1890: \end{document}
1891:
1892: