1: \documentclass[11pt,preprint2]{aastex}
2: \usepackage{amsmath}
3:
4:
5: \newcommand{\vdag}{(v)^\dagger}
6: \newcommand{\myemail}{msp@phys.au.dk}
7:
8: \shorttitle{Spin down of cooling hybrid stars}
9: \shortauthors{M. Stejner, F. Weber \& J. Madsen}
10:
11: \begin{document}
12:
13: \title{Signature of deconfinement with spin down compression in
14: cooling hybrid stars}
15:
16: \author{Morten Stejner}
17: \affil{Department of Physics and Astronomy, University of Aarhus}
18: \affil{Ny Munkegade, Bld. 1520, DK-8000 Aarhus C, Denmark.}
19: \email{\myemail}
20: \and
21: \author{Fridolin Weber}
22: \affil{Department of Physics, San Diego State University}
23: \affil{5500 Campanile Dr, San Diego CA 92182-1233}
24: \and
25: \author{Jes Madsen}
26: \affil{Department of Physics and Astronomy, University of Aarhus}
27: \affil{Ny Munkegade, Bld. 1520, DK-8000 Aarhus C, Denmark.}
28:
29:
30: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
31: %% Abstract
32: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
33:
34: \begin{abstract}
35: \noindent The thermal evolution of neutron stars is coupled to their spin down
36: and the resulting changes in structure and chemical composition. This
37: coupling correlates stellar surface temperatures with rotational state
38: as well as time. We report an extensive investigation of the coupling
39: between spin down and cooling for hybrid stars which undergo a phase
40: transition to deconfined quark matter at the high densities present in
41: stars at low rotation frequencies. The thermal balance of neutron
42: stars is re-analyzed to incorporate phase transitions and the related
43: latent heat self-consistently, and numerical calculations are
44: undertaken to simultaneously evolve the stellar structure and
45: temperature distribution. We find that the changes in stellar
46: structure and chemical composition with the introduction of a pure
47: quark matter phase in the core delay the cooling and produce a period
48: of increasing surface temperature for strongly superfluid stars of
49: strong and intermediate magnetic field strength. The latent heat of
50: deconfinement is found to reinforce this signature if quark matter is
51: superfluid and it can dominate the thermal balance during the
52: formation of a pure quark matter core. At other times it is less
53: important and does not significantly change the thermal evolution.
54: \end{abstract}
55:
56: \keywords{Stars:neutron --- stars:rotation --- dense
57: matter --- equation of state}
58:
59: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
60: %% Text
61: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
62:
63:
64: \section{Introduction}
65: The chemical composition of neutron stars at densities beyond nuclear
66: saturation remains uncertain with alternatives ranging from purely
67: nucleonic compositions through hyperon or meson condensates to
68: deconfined quark matter -- see e.g. \cite{Weber:2005} and
69: \cite{Page:2006} for recent reviews with emphasis on quark matter. A
70: future understanding of neutron star structure gained through
71: confrontation of theoretical models with the now steadily growing body
72: of observational facts will therefore simultaneously constrain
73: fundamental elements of particle and nuclear physics
74: \citep{Lattimer:2007}. The interlinked processes of spin down
75: and thermal cooling present intriguing prospects of gaining insight in
76: the properties of matter in neutron star cores by confrontation with
77: soft X-ray observations of thermal radiation from neutron star
78: surfaces as they both depend sensitively on and to some extent
79: determine the chemical composition.
80:
81: As neutron stars spin down and contract, their structure and
82: composition change with the increasing density -- drastically if phase
83: boundaries are crossed so new forms of matter become possible. We
84: shall here investigate how this influences the thermal evolution of
85: hybrid stars which contain large amounts of deconfined quark
86: matter. The increasing density and changing chemical composition
87: further imply additional entropy production in bulk and the release of
88: latent heat as particles cross any phase boundaries present. We
89: therefore re-analyze the thermal equilibrium of compact stars to show
90: how mixed phases may be incorporated. We thus arrive at a natural
91: description of the latent heat of phase transitions in compact stars,
92: but also find -- through direct numerical calculations -- that unless
93: the stellar structure changes very rapidly the effects of latent heat
94: on the thermal evolution are insignificant when compared to those of
95: the changing chemical composition and the surface area reduction.\\
96:
97:
98: Neutron stars are extremely compact objects and densities in their
99: cores reach well in excess of the nuclear saturation density. At such
100: densities the distance between particles is on the order of the
101: characteristic range of the nuclear forces. Therefore, as was stressed
102: recently by \cite{Baym:2006}, perturbative treatments in terms of few-
103: or even many-body forces -- although highly successful in describing
104: the properties of matter below nuclear saturation -- are no longer
105: well defined in the core of neutron stars. Further, the relevant
106: degrees of freedom should include the appearance of hyperons and
107: possibly deconfined quarks, so treatments in terms of nucleons alone
108: must also be seen as approximative. Hyperons are expected to appear at
109: densities around $2-3\rho_0$, where $\rho_0 = 0.153$ fm$^{-3}$ is the
110: baryon density at nuclear saturation. The appearance of hyperons so
111: softens the equation of state that purely hadronic equations of state
112: may not allow stable models compatible with the accurately measured
113: masses of neutron stars in binary systems
114: \citep{Baldo:2003,Schulze:2006}. \cite{Schulze:2006} further
115: demonstrate that this conclusion is highly robust with respect to
116: different assumptions about hyperon interactions and the nucleonic
117: equation of state. At best these studies are strong arguments against
118: purely hadronic compositions and they certainly do show the relevance
119: of considering alternatives such as a
120: transition at high densities to deconfined quark matter.\\
121:
122: Quark matter represents an entirely new type of matter -- as opposed
123: to just an additional degree of freedom as in the case of hyperons in
124: the hadronic phase -- and it cannot be assumed to soften the equation
125: of state to the same extent
126: \citep{Alford:2007a,Schaffner-Bielich:2007}. Hybrid stars can be
127: be consistent with the high masses and radii indicated by recent
128: observations (e.g. \cite{Ozel:2006},
129: \cite{Freire:2008,Freire:2008a,Freire:2007a}), and the quark matter
130: equation of state can fulfill the
131: constraints imposed by heavy-ion collision transverse flow data and
132: $K^+$ production (which nucleonic equations of state do not). Further,
133: \cite{Drago:2008} found that only stars with a quark matter component
134: can rotate stably without losing angular momentum by emission of
135: gravitational waves through the r-mode instability at the 1122 Hz
136: indicated by recent observations for the X-ray transient XTE J1739-285
137: \citep{Kaaret:2007}. This conclusion is tentative as the observed
138: neutron star spin frequency awaits confirmation, but it again shows
139: how the composition of neutron stars is an open question to be
140: determined by a confrontation of theory and observation, and that
141: a deconfined quark matter phase remains a viable alternative.\\
142:
143: For definiteness we shall work with the equation of state suggested
144: by \cite{Glendenning:1992}; hereafter the G$^{300}_{180}$ equation of
145: state. While this equation of state is not sophisticated in its
146: treatment of the quark matter phase, it is illustrative in that it
147: allows a very large pure quark matter core with a transition through a
148: mixed phase at relatively low densities around 2-5$\rho_0$. For our
149: purposes it is interesting in that it implies drastic changes in
150: composition with spin down -- the pure quark matter core disappears at
151: high spin frequencies for instance -- and it therefore represents
152: something close to a limiting case. We shall then be able to
153: investigate both the effects of steady conversion of hadronic matter
154: to quark matter and the sudden appearance of a pure quark matter core
155: in the hadronic phase.
156:
157: In the so-called minimal
158: scenario, which excludes exotic and very rapid neutrino processes
159: \citep{Page:2004}, the internal temperature of neutron stars
160: drops from beyond $10^{10}$~K to around $10^9$~K within a few minutes
161: after their birth, and neutrino cooling continues to dominate for at
162: least the next few thousand years until the internal temperature has
163: dropped below $10^8$~K and photon cooling takes over (see,
164: e.g. \cite{Page:2004, Yakovlev:2004a} for reviews). If the highly
165: efficient direct Urca neutrino emission (i.e., essentially beta decays,
166: $n\rightarrow p+e^-+\bar{\nu}_e$ and related processes in quark
167: matter) is active, the stars may cool very rapidly and reach very low
168: temperatures on a timescale of a few hundred years. As we shall see
169: the nuclear direct Urca process is active in the mixed phase of hybrid
170: stars, and may thus control the thermal evolution. These processes may
171: be suppressed by pairing of the participating particles however, and
172: further the extent of the mixed phase depends strongly on the rotation
173: frequency thus giving rise to a diverse range of possible cooling
174: paths.
175:
176: The latent heat of the phase transition and the related release of
177: entropy in bulk with changing density are generally found to be of
178: smaller importance than the changing structure and chemical
179: composition. It does not significantly delay or enhance the cooling,
180: although it does briefly balance or even dominate the cooling terms
181: when the pure quark matter core is first formed for certain choices of
182: stellar parameters. The latent heat of the quark matter phase
183: transition was previously considered by \cite{Miao:2007, Miao:2007a} and,
184: \cite{Xiaoping:2008}. These authors took a different approach to calculate
185: the latent heat than what is discussed below and found very
186: significant heating terms which we do not recover.
187:
188: The work of \cite{Reisenegger:1995} and lately \cite{Fernandez:2005}
189: considered the related process of roto-chemical heating for neutron
190: stars in which weak reactions are driven out of equilibrium by the
191: changing density. The resulting release of energy was found able to
192: maintain old stars at relatively high temperatures determined by their
193: rate of spin down. This term naturally appears in our equations for
194: the thermal balance and should be included in a full model. It is to
195: some extent complementary to the effects discussed here, but the
196: treatment of weak reactions beyond equilibrium is beyond our scope in
197: this work, and we shall assume chemical equilibrium throughout. We
198: return briefly to discuss this issue in Sect.~\ref{discuss}. \\
199:
200: The link between the thermal evolution of neutron stars and
201: their spin down has become a subject of some interest as an
202: observational correlation between the temperatures and inferred
203: magnetic fields of neutron stars was recently discovered by
204: \cite{Pons:2007}. This was interpreted as evidence for magnetic field
205: decay and detailed work by \cite{Aguilera:2007, Aguilera:2008}
206: strongly supports this conclusion. An alternative interpretation was
207: offered recently by \cite{Niebergal:2007} in terms of magnetic flux
208: expulsion from color-flavor locked quark stars (a hypothetical class
209: of stars consisting entirely of quark matter which in this case is
210: assumed to be absolutely stable). The previously mentioned work of
211: \cite{Drago:2008} also indicates a link between the spin down and
212: cooling of hybrid stars. These correlations -- if they can indeed be
213: shown to exist -- complement the traditional cooling calculations
214: which relate only temperature and age, and they may help break the
215: degeneracy seen between such calculations with different underlying
216: assumptions about
217: the state of matter at very high density.\\
218:
219: In the following, we shall first revisit the equations of
220: thermal balance for compact stars in Sect.~\ref{thequilibrium} to show
221: the effects of a time-dependent density and discuss how the presence
222: of phase transitions may be included. In Sect.~\ref{eosandmodel}, we
223: discuss the G$_{180}^{300}$ equation of state, the resulting stellar
224: models and some simple estimates for the additional terms in the
225: thermal balance equations in more detail. In
226: Sect.~\ref{calculations} we show the results of including these terms
227: in spherical isothermal cooling calculations. We conclude with a
228: discussion in Sect.~\ref{discuss}.
229:
230: \section{Thermal Equilibrium in the Mixed Phase}\label{thequilibrium}
231: The thermal evolution of compact stars is determined by the equations
232: of local energy conservation and transport in the framework of general
233: relativity. These equations balance any energy radiated away from the
234: star by photons and neutrinos against changes in rest mass due to
235: nuclear reactions and changes in gravitational or internal energy as
236: the stellar structure evolves. In the most widely studied scenario the
237: stellar structure is assumed constant and the only energy source
238: available to neutron stars is then their original endowment of thermal
239: energy (e.g.,
240: \cite{van-Riper:1991,Schaab:1996,Page:2004,Page:2006a,Yakovlev:2004a}
241: and references therein). Neutrino production in the core and photon
242: emission from the stellar surface then ensures a monotonical and
243: sometimes very rapid cooling of the star depending strongly on any
244: superfluid properties of the core. But in addition to this a number of
245: powerful energy sources may play a role in delaying or even reversing
246: the cooling at various stages in a neutron stars life -- see
247: e.g. \cite{Schaab:1999} for an overview of possible sources and their
248: effects. Here we discuss the effects of including a phase transition
249: in the energy balance, which, as the stellar structure changes, must
250: then also include redistribution of entropy between the two phases as
251: their proportion changes as well as changes in surface and Coulomb
252: energy for transitions through a mixed phase.
253:
254: Following \cite{Thorne:1966} or the equivalent discussion in
255: \cite{Weber:1999} we consider energy conservation for a spherical
256: shell inside of which is $a$ baryons and which itself contains $\delta
257: a$ baryons. The treatment given in this section therefore assumes
258: spherical symmetry which is sufficient to show the effects of a phase
259: transition on local energy conservation -- we shall return later to
260: discuss possible consequences of a multidimensional cooling
261: calculation. The shell under consideration will change its
262: internal energy during a coordinate time interval d$t$ by
263: \begin{eqnarray}\label{balance}
264: &&d(\textrm{internal energy})=\nonumber\\
265: &&(\textrm{amount of rest mass-energy converted
266: to} \nonumber\\
267: &&\textrm{internal energy by reactions})\nonumber\\
268: +&&(\textrm{work done on shell by
269: gravitational}\nonumber\\
270: &&\textrm{forces to change its volume during} \nonumber\\
271: &&\textrm{quasi-static contraction})\nonumber\\
272: -&&(\textrm{energy radiated, conducted} \nonumber\\
273: &&\textrm{or convected away from the shell}).
274: \end{eqnarray}
275:
276: It is important to note at this point that if the two phases are in
277: equilibrium -- as should certainly be expected for transitions
278: governed by strong reactions such as that from hadronic to quark
279: matter -- there is {\it no binding energy} involved in the transition
280: and therefore no contribution to the first term on the right hand side
281: of Eq.~(\ref{balance}). If this was the case and the phase transition did
282: involve a binding energy the two phases could not be in
283: equilibrium and the star would adjust itself accordingly -- eventually
284: becoming a strange star in the case of the quark matter transition if
285: quark matter was assumed bound relative to hadronic matter at zero
286: external pressure. In hybrid stars this is not the assumption however,
287: and so any latent heat evolved or absorbed in the phase transition
288: follows from the different thermodynamical properties of the
289: two phases as with any other phase transition.
290:
291: In the mixed phase of the G$^{300}_{180}$ equation of state regions
292: containing negatively charged quark matter appear at densities of
293: about 2 times nuclear saturation and dominate completely at 5
294: times nuclear saturation. They allow the hadronic matter to
295: lower its isospin asymmetry energy and become positively charged by
296: including more protons with charge neutrality achieved globally. The
297: geometry and structure of the mixed phase is determined by a balance
298: between surface tension and Coulomb repulsion between regions of like
299: charge. For details on the phase transition we refer to
300: e.g. \cite{Glendenning:1992,Heiselberg:1993,Glendenning:2000,Glendenning:2001a,
301: Voskresensky:2003,Endo:2006}.
302:
303: Returning to our spherical shell of baryon number $\delta a$ we will
304: assume that its volume $V$ is large enough to contain a macroscopic number of unit
305: cells each containing a region filled by the rare phase whose presence
306: may then be considered a microscopic property of the equation of
307: state. The work done to change the shells volume during
308: contraction or expansion of the star must then include changes in
309: surface and Coulomb energy as well as the usual pressure term
310: \begin{equation}\label{work}
311: \mathrm{d}W=-P\mathrm{d}V + \alpha \mathrm{d}\mathbb{S}+\mathrm{d}E_\mathrm{C}
312: \end{equation}
313: where $\alpha$ is the surface tension, $\mathbb{S}$ is the amount of
314: surface area in the shell dividing the two phases and $E_\mathrm{C}$
315: is the Coulomb energy contained in the shell. These terms can be
316: similarly included in the first law of thermodynamics which we write
317: in the same notation (units with $G=c=k_\mathrm{B}=1$ will be used
318: here and throughout this paper)
319: \begin{eqnarray}\label{firstlaw}
320: \mathrm{d}\epsilon=&\frac{P+\epsilon}{\rho}\mathrm{d}\rho+T\rho\mathrm{d}s
321: +\sum_k \mu_k \rho \mathrm{d}Y_k\nonumber\\
322: &+\rho\frac{\alpha
323: \mathrm{d}\mathbb{S}}{\delta a}+\rho\frac{\mathrm{d}
324: E_\mathrm{C}}{\delta a}
325: \end{eqnarray}
326: where $s$ is the entropy per baryon and $Y_\mathrm{k}=\rho_k/\rho$ is
327: the fraction of the baryon number density in the form $k$ with $k$
328: running over all particle species and $\sum_k Y_k =1$. In the quark
329: matter phase each quark contributes by only $\frac{1}{3}$ to the
330: baryon number density and $s$ is then the entropy per three quarks. The
331: last two terms in Eq.~(\ref{firstlaw}) are the local densities of
332: surface and Coulomb energy written in a form useful for our purpose.
333:
334: Inserting this in Eq.~(\ref{balance}) and using that the amount of energy
335: radiated, conducted or convected away from the shell can be written in
336: terms of the gradient of the total luminosity, $L_\mathrm{tot}$, we
337: show in Appendix~\ref{appendix} that local energy balance may be expressed as
338: \begin{equation}\label{balance2}
339: \frac{\mathrm{d}}{\mathrm{d}a} (L_\mathrm{tot} e^{2\Phi}) =
340: -e^{\Phi}\left [
341: T\left(\frac{\mathrm{d}s}{\mathrm{d}t}\right)_{a}+\sum_k \mu_k \left(\frac{\mathrm{d}Y_k}{\mathrm{d}t}\right)_{a} \right ]
342: \end{equation}
343: thus giving the contribution to $L_\mathrm{tot}$ from the shell in
344: terms of the entropy production and the difference between the
345: chemical potentials of particles participating in any reactions taking
346: place in the shell. $e^{2\Phi}$ is the time component of a spherically
347: symmetric metric
348: \begin{equation}
349: ds^2=-e^{2\Phi}dt^2+e^{2\Lambda}dr^2+r^2d\theta^2+r^2\sin{\theta}^2d\Phi^2
350: \end{equation}
351: found from the general relativistic structure equations for compact
352: stars.
353:
354: $L_\mathrm{tot}$ includes the neutrino
355: luminosity, but since neutrinos can be taken to immediately escape
356: from the star when they are created without converting into any other
357: form of energy along the way, they fulfill their own separate
358: equation of energy conservation
359: \begin{equation}\label{neutrinobalance}
360: \frac{\mathrm{d}}{\mathrm{d} a}(L_\nu e^{2\Phi}) = \frac{\epsilon_\nu}{\rho}e^{2\Phi}
361: \end{equation}
362: where $L_\nu$ is the neutrino luminosity and $\epsilon_\nu$ is the neutrino emissivity; the rate per unit
363: volume at which neutrino energy is created. In neutron stars
364: convection is negligible compared to electron conduction and photon diffusion
365: and so the remainder of
366: $L_\mathrm{tot}$ can be shown to fulfill a transport equation
367: \citep{Thorne:1966,Weber:1999}
368: \begin{equation}\label{transport}
369: \frac{\mathrm{d}}{\mathrm{d} a} \left(T e^\Phi \right) =
370: -\frac{3}{16\sigma}\frac{\kappa
371: \epsilon}{T^3}\frac{Le^\Phi}{16\pi^2 r^4 \rho}
372: \end{equation}
373: where $\sigma$ is the Stefan-Boltzmann constant, $\kappa$ is the
374: total thermal conductivity and $L=L_\mathrm{tot}-L_\nu$. At the
375: stellar center we must have $L_\mathrm{tot}(a=0)=0$ while at the
376: stellar surface $L$ must equal the total stellar photon luminosity
377: which may depend on assumptions about properties of the stellar
378: atmosphere or lack thereof.
379:
380: Eqs.~(\ref{balance2})-(\ref{transport}) with the appropriate boundary
381: conditions can be solved to evolve the thermal structure of a stellar
382: model. They have exactly the same form as would be expected in the
383: absence of any phase transitions. However, the phase transition influences
384: the entropy density and this, as we shall see, gives rise to additional
385: terms in the heat balance including latent heat and the surface and
386: Coulomb energies. An equivalent form of Eq.~(\ref{balance2})
387: showing the contributions from surface and Coulomb terms more
388: explicitly can be found in Eq.~(\ref{cool1}) of Appendix~\ref{appendix}. Combining Eqs.~(\ref{balance2}) and
389: (\ref{neutrinobalance}), noting that $c_\mathrm{v}=\rho T(\partial
390: s/\partial T)_V$ and assuming constant structure and chemical
391: composition we also recover the standard cooling equation for static
392: stars
393: \begin{equation}
394: \frac{\mathrm{d}}{\mathrm{d} a}(L e^{2\Phi})
395: =-\rho^{-1}\left(\epsilon_\nu e^{2\Phi}+c_\mathrm{v}
396: \frac{\mathrm{d}(Te^{\Phi})}{\mathrm{d} t} \right)
397: \end{equation}
398:
399: Eq.~(\ref{balance2}) is useful because it allows a
400: particularly simple analysis in the presence of a phase
401: transition. The first thing to note is, that particles crossing the
402: phase boundary would not contribute to the second term on the right
403: hand side of Eq.~(\ref{balance2}) if the two phases are in or close to
404: equilibrium, because their chemical potentials (or those of their
405: constituents) must then be continuous across the phase boundary. In
406: the following we therefore neglect this term, but we shall return to
407: discuss its potential importance for reactions not in equilibrium. The
408: first term is a different matter however. The entropy per particle is
409: a function of density, temperature and chemical composition which is
410: not required to be continuous across the phase boundary, so particles
411: making the transition will release or absorb heat accordingly. To see
412: how this works we write $s$ as a sum with bulk contributions from each
413: phase according to their volume or mass fraction as well as
414: contributions from the surface and Coulomb energies
415: \begin{align}
416: s&=\frac{1}{\rho}((1-\chi_\mathrm{v})S_{\mathrm{1}}+\chi_\mathrm{v}S_\mathrm{2}+S_\mathrm{S}+S_\mathrm{C})\\
417: &=(1-\chi_\rho)s_1+\chi_\rho s_2+\frac{1}{\rho}(S_\mathrm{S}+S_\mathrm{C})
418: \end{align}
419: where
420: $\rho=(1-\chi_\mathrm{v})\rho_\mathrm{1}+\chi_\mathrm{v}\rho_\mathrm{2}$
421: is the average of the two particle densities $\rho_{1,2}$ weighed by
422: the volume fraction of the dense phase, $\chi_\mathrm{v}$. $S_{1,2}$
423: is the bulk entropy density in each phase and $S_\mathrm{S}$ and
424: $S_\mathrm{C}$ are the surface and Coulomb contributions to the
425: entropy density respectively. We have further
426: introduced the baryon number fraction $\chi_\rho$ of baryons in the
427: dense phase in a sample of matter related to the volume fraction by
428: $\chi_\rho=\chi_\mathrm{v}(\rho_2/\rho)$, as well as the entropies
429: per baryon in the respective phases, $s_{1,2}=S_{1,2}/\rho_{1,2}$. At
430: constant $a$ and assuming the fraction of matter in the dense phase and
431: the particle densities do not depend on temperature we then have
432: \begin{eqnarray}\label{dsdt}
433: T\frac{\mathrm{d}s}{\mathrm{d}t}&=&\frac{c_\mathrm{v}}{\rho}\frac{\mathrm{d}T}{\mathrm{d}t}+T\frac{\mathrm{d}s}{\mathrm{d}\rho}\frac{\mathrm{d}\rho}{\mathrm{d}t}\nonumber\\&=&\frac{c_\mathrm{v}}{\rho}\frac{\mathrm{d}T}{\mathrm{d}t}+T\frac{\mathrm{d}\chi_\rho}{\mathrm{d}\rho}\frac{\mathrm{d}\rho}{\mathrm{d}t}\left(s_2-s_1\right)\nonumber\\
434: &&+(1-\chi_\rho)T\frac{\mathrm{d}\rho}{\mathrm{d}t}\frac{\mathrm{d}s_1}{\mathrm{d}\rho}
435: +\chi_\rho
436: T\frac{\mathrm{d}\rho}{\mathrm{d}t}\frac{\mathrm{d}s_2}{\mathrm{d}\rho}\nonumber\\
437: &&+T\frac{\mathrm{d}}{\mathrm{d}t}\left[\frac{1}{\rho}\left(S_\mathrm{S}+S_\mathrm{C}\right)\right]
438: \end{eqnarray}
439: where the heat capacity is again a weighed volume
440: average in the mixed phase
441: \begin{equation}
442: c_\mathrm{v}=(1-\chi_\mathrm{v})c_\mathrm{V,1} +
443: \chi_\mathrm{v}c_\mathrm{V,2}\; .
444: \end{equation}
445:
446: The latent heat absorbed by a particle crossing the boundary between
447: two phases in equilibrium is the temperature times the difference in
448: entropy per particle between the two phases, $q=T[s_2-s_1]$
449: \citep{Landau:1980}; we recover this in the second term on the right
450: hand side of Eq.~(\ref{dsdt}). If any term in Eq.~(\ref{dsdt}) is
451: negative, heat is evolved by this term which then heats the star and
452: adds to the luminosity $L$. In particular the latent heat is a heating
453: term for increasing density when the entropy per baryon of the dense
454: phase is less than that of the low density phase -- a situation which
455: will arise when considering superfluid quark matter.\\
456:
457: For future reference we identify the terms in Eq.~(\ref{dsdt}) as
458: follows. $T\frac{\mathrm{d}\chi_\rho}{\mathrm{d}\rho}\frac{\mathrm{d}\rho}{\mathrm{d}t}(s_2-s_1)$
459: is identified as the latent
460: heat, $(1-\chi_\rho)T\frac{\mathrm{d}\rho}{\mathrm{d}t}\frac{\mathrm{d}s_1}{\mathrm{d}\rho}$
461: is identified as the hadronic bulk contribution, $\chi_\rho
462: T\frac{\mathrm{d}\rho}{\mathrm{d}t}\frac{\mathrm{d}s_2}{\mathrm{d}\rho}$
463: is identified as the quark bulk contribution, and the
464: last term $T(\mathrm{d/d}t)[(S_\mathrm{S}+S_\mathrm{C})/\rho]$ is identified as the surface and Coulomb contribution. Further we often
465: refer to these terms collectively as additional entropy production (or
466: release) beyond what would be expected at constant density.
467:
468: We discuss in the following section how to calculate the bulk
469: entropy density of the hadronic and quark matter phases in the
470: relativistic mean field theory framework of the
471: $\mathrm{G}_{180}^{300}$ equation of state. For now let us just
472: remark that the entropy per particle for an ideal relativistic
473: degenerate Fermi gas is \citep{Landau:1980}
474: \begin{equation}\label{s}
475: s=\frac{(3\pi^2)^{\frac{2}{3}}}{3\hbar c}T\rho^{-\frac{1}{3}}=0.02\frac{T}{\mathrm{MeV}}\left(\frac{\rho}{\mathrm{fm}^{-3}}\right)^{-\frac{1}{3}}\;.
476: \end{equation}
477: Taking the transition to quark matter as a transition between such
478: gases -- note that the bag constant does not contribute to the
479: entropy -- and remembering that each hadron contains three quarks
480: which further have a color degeneracy of three, the latent heat in
481: Eq.~(\ref{dsdt}) is of the order of
482: \begin{align}\label{estimate}
483: 10^{33}T_9^2\bigg[9\bigg(\frac{\rho_\mathrm{Q}}{\mathrm{fm}^{-3}}\bigg)^{-\frac{1}{3}}-&
484: \left(\frac{\rho_\mathrm{H}}{\mathrm{fm}^{-3}}\right)^{-\frac{1}{3}}\bigg]\\\nonumber
485: &\times\frac{\mathrm{d}a_\mathrm{Q}/\mathrm{d}t}{10^{57}/10^7\mbox{ yr}}\mbox{ erg s}^{-1}
486: \end{align}
487: where $T_9=T/10^9$~K, and the rate of baryons making the transition to
488: quark matter, $\mathrm{d}a_\mathrm{Q}/\mathrm{d}t$, was (arbitrarily)
489: scaled to an entire star being converted steadily over $10^7$
490: years. Unless the star is very hot or changing structure fast this is
491: a very modest contribution, and we further note that it is positive for
492: a contracting star and so acts to cool the star down. However the
493: sign is subject to the very rough assumption that both gases may be
494: treated as ideal Fermi gases for the purpose of calculating their
495: entropy, and it will change in a more detailed treatment. Specifically
496: the quark phase may be color-superconducting with energy gaps as high
497: as 100 MeV and corresponding critical temperatures of the order of
498: $10^{11}$~K. Below the critical temperature the quark specific heat
499: and entropy density would be exponentially suppressed and could be
500: ignored relative to the hadronic contribution in Eq.~(\ref{estimate})
501: which would then be a heating term. We shall return to both
502: possibilities in later sections.
503:
504: We shall calculate the two bulk terms numerically in the following
505: sections, so for now we just note that from Eq.~(\ref{s}) the entropy
506: per baryon decreases with increasing density. In a contracting star
507: these terms will hence act as heating terms and locally be of the same
508: order of magnitude as the release or absorption of latent heat in
509: Eq.~(\ref{estimate}) -- but of course they contribute throughout the
510: star and are therefore potentially far more important than the latent
511: heat which is only significant in the mixed phase.
512:
513: Since the surface and Coulomb energies are related by
514: $E_\mathrm{S}=2E_\mathrm{C}$ in equilibrium \citep{Glendenning:2000}, their contributions to the
515: thermodynamic potential, and hence to the entropy, are
516: similarly fixed in proportion, and we need only consider one of them
517: here. Specifically the surface part of the entropy may be found from
518: \citep{Landau:1980}
519: \begin{eqnarray}
520: &&\Omega=\Omega_0+\Omega_\mathrm{S}=\Omega_0+\alpha\mathbb{S}\\
521: &&S_\mathrm{S}=-\frac{\partial \Omega_\mathrm{S}}{\partial
522: T}=-\mathbb{S}\frac{\partial \alpha}{\partial T}\;,
523: \end{eqnarray}
524: where $\Omega_\mathrm{S}$ is the surface contribution to the thermodynamic
525: potential $\Omega$. The surface
526: tension for the quark-hadron interface, $\alpha$, determines the
527: geometry and extent of the mixed phase \citep{Glendenning:1992,
528: Heiselberg:1993, Voskresensky:2003, Endo:2006}. The surface tension
529: of strangelets in vacuum has been evaluated by \cite{Berger:1987}, but
530: the surface tension for the mixed phase remains essentially
531: unknown. It is commonly parameterized as being proportional to the
532: difference in energy density between the two phases and the length
533: scale $L\sim 1$ fm of the strong interaction \citep{Glendenning:2000}
534: \begin{equation}
535: \alpha=K\times[\epsilon_\mathrm{Q}-\epsilon_\mathrm{H}]\times L\;.
536: \end{equation}
537: Assuming for simplicity that $K$ and $L$ are constant we then get
538: \begin{align}
539: &\frac{\partial \alpha}{\partial T}=\alpha\frac{c_\mathrm{V,Q}-c_\mathrm{V,H}}{\epsilon_\mathrm{Q}-\epsilon_\mathrm{H}}\\
540: &\rho^{-1}(S_\mathrm{S}+S_\mathrm{C})=-\frac{3}{2}\frac{\epsilon_\mathrm{S}}{\rho}\frac{c_\mathrm{V,Q}-c_\mathrm{V,H}}{\epsilon_\mathrm{Q}-\epsilon_\mathrm{H}}
541: \end{align}
542: The corresponding term in Eq.~(\ref{dsdt}) is then of the order of the
543: surface energy per baryon times the ratio between thermal and total
544: energy density. It can therefore not be expected to contribute
545: significantly, and this expectation is confirmed by the numerical results.
546:
547: \section{Equation of State and Stellar Models}\label{eosandmodel}
548: In our numerical work we have employed the rotating neutron star code
549: developed by Weber and the G$_{180}^{300}$ equation of state used by
550: \cite{Glendenning:1992,Glendenning:2000} and
551: \cite{Glendenning:2001a}. Here we shall briefly describe each of these
552: and the resulting stellar models.
553:
554: The G$_{180}^{300}$ equation of state treats the deconfined quark
555: matter phase in a simple version of the bag model which ignores gluon
556: interactions. The confined hadronic phase is described in terms of the
557: mean field solution to a covariant Lagrangian that involves the baryon
558: octet interacting through scalar, vector and vector-isovector
559: mesons. The precise values of the coupling constants for the hadronic
560: part of the model correspond to the first set of Table 5.5 of
561: \cite{Glendenning:2000} or Table 1 of \cite{Glendenning:2001a}. For
562: details we refer to these works where the resulting equation of state as well as the underlying theory are
563: carefully described at zero temperature. Finite temperature expressions for pressure, energy and particle
564: densities can be found by reinserting the Fermi distribution in the
565: phase space integrals of their zero temperature expressions which we
566: shall not write explicitly here. We refer to
567: e.g. \cite{Glendenning:1990} for the full temperature dependent
568: expressions (this reference also includes gluon interactions to first
569: order which we ignore here).
570:
571: The entropy is calculated as
572: \begin{equation}
573: s=\frac{1}{\rho T}\left[P+\epsilon-\sum_i \mu_i\rho_i\right]
574: \end{equation}
575: using the finite temperature expressions outlined above and using the
576: highly accurate publicly available code described in
577: \cite{Miralles:1996} to solve the Fermi integrals. In keeping with our
578: assumption that temperatures remain too low to significantly influence the
579: chemical composition we neglect contributions from thermally exited
580: particle-antiparticle pairs.
581:
582: Figs.~\ref{composition} and \ref{entropy} show the resulting chemical
583: composition and entropy. In the quark phase we note
584: that u-quarks are suppressed initially giving the phase a negative net
585: charge, and that as expected the entropy per baryon -- though not per
586: quark -- is higher in the quark matter phase in the absence of any
587: pairing phenomena. We have checked numerically that in the absence of
588: pairing the entropy simply scales linearly with temperature within
589: the temperature range we shall need.
590:
591: The sum of the surface and Coulomb energies has a maximum as a
592: function of density around $\rho=0.5-0.6 \textrm{ fm}^{-3}$ with a
593: corresponding minimum in the related entropy in Fig.~\ref{ssurf}. The
594: surface and Coulomb term in Eq.~(\ref{dsdt}) can therefore be either
595: positive or negative and either heat or cool the star
596: accordingly. The surface and
597: Coulomb contribution to the entropy is negative with the total entropy
598: remaining positive, which confirms that a structured phase has lower
599: entropy than an unstructured one.
600: \begin{figure}[!htb]
601: \plotone{f1.eps}
602: \caption{Chemical composition of the G$^{300}_{180}$ equation of state
603: as a function of total baryon density. Individual particle densities
604: $\rho_i$ refer to the density in regions filled with the respective
605: phase. $\chi_v$ is the (dimensionless) volume fraction of quark matter.}
606: \label{composition}
607: \end{figure}
608:
609: \begin{figure}[!htb]
610: \plotone{f2.eps}
611: \caption{Entropy per baryon at $T=1$~MeV in each phase, $s_\mathrm{H}$ and
612: $s_\mathrm{Q}$, and per particle for each particle, $s_i$, as
613: functions of total baryon density for the G$^{300}_{180}$ equation
614: of state. Protons and $\Lambda$ particles have high $s_i$ because
615: they are so rare, but contribute little to $s_\mathrm{H}$ for
616: the same reason.}
617: \label{entropy}
618: \end{figure}
619:
620: \begin{figure}[!htb]
621: \plotone{f3.eps}
622: \caption{Sum of surface and Coulomb entropy as a
623: function of total baryon density for the G$^{300}_{180}$
624: equation of state at $T=1$~MeV}
625: \label{ssurf}
626: \end{figure}
627:
628:
629: We have calculated the structure of a sequence of stars with rotation
630: frequencies (as seen by an observer at infinity) between zero and the
631: limiting mass-shedding Kepler frequency for the G$^{300}_{180}$
632: equation of state. For this purpose we use the perturbative method of
633: \cite{Hartle:1967} and \cite{Hartle:1968} as implemented in the
634: numerical code developed by Weber which also solves self-consistently
635: for the general relativistic Kepler frequency $\Omega_\mathrm{K}$ --
636: see \cite{Weber:1999} for the derivation of $\Omega_\mathrm{K}$ and
637: \citep{Weber:1991,Weber:1992} for further details. The sequence has
638: constant total baryon number $A=1.87\,\times\,10^{57}$ and
639: nonrotating total gravitational mass $M=1.42$ M$_\odot$. The Kepler
640: frequency is then found to be $\Omega_\mathrm{K}=6168 \mbox{ rad
641: s}^{-1}$ corresponding to a period of 1.02 ms.
642:
643: \begin{figure}[!htb]
644: \plotone{f4.eps}
645: \caption{Circumferential stellar radius in the equatorial and polar directions for
646: rotating stars of total nonrotating gravitational mass $M=1.42$
647: M$_\odot$.}
648: \label{OmegaR}
649: \end{figure}
650:
651: Figs.~\ref{OmegaR} and~\ref{AObound} show a few
652: properties of the models. The stars are significantly distorted by
653: rotation and increase their equatorial radius at the Kepler frequency
654: by half the nonrotating radius (Fig.~\ref{OmegaR}) while losing the
655: pure quark matter core at $\Omega=1400\mbox{ rad s}^{-1}$
656: (Fig.~\ref{AObound}). Since the stellar photon luminosity is
657: proportional to the surface area and must match the energy flux
658: emerging from the core, higher surface temperatures must also be
659: expected at low rotation frequencies for this reason alone. The
660: distortion from spherical symmetry depends on polar angle, however, and
661: above the frequency at which the quark matter core is lost the star
662: actually contracts in the polar direction.
663:
664: In Fig.~\ref{AObound} we show the location of the phase boundaries
665: between pure hadronic matter, the various geometries of the mixed
666: phase and the pure quark matter phase. Here we use as the free
667: variable the baryon number, $a$, contained within a surface on which
668: the density is uniform (i.e spatially but {\it not} temporally
669: constant). The Eulerian
670: density change $(\mathrm{d}\rho/\mathrm{d}\Omega)_r$ can be positive
671: or negative depending on location while the Lagrangian derivative
672: $(\mathrm{d}\rho/\mathrm{d}\Omega)_a$ is always negative, and so $a$
673: is a more convenient variable for some purposes. In Fig.~\ref{AObound} borders are shown at the densities which correspond to each
674: transition and $a$ is scaled to the total baryon number, $A$. We note that a large fraction of the star is
675: converted to quark matter as the star spins down. The pure quark
676: matter core appears below $\Omega=1400 \mbox{ rad s}^{-1}$ and
677: grows to eventually comprise $30$ \% of the stellar
678: gravitational mass and 26 \% of the stellar baryon number.
679:
680: \begin{figure}[!htb]
681: \plotone{f5.eps}
682: \caption{Location of phase boundaries in stars of nonrotating mass
683: $M=1.42$ M$_\odot$. The locations are given as the fractions of the
684: total stellar baryon number contained within a surface of spatially constant
685: density corresponding to each transition. The inset shows the region
686: where the phase boundaries almost -- but not quite -- cross.}
687: \label{AObound}
688: \end{figure}
689:
690:
691: Within the framework of \cite{Hartle:1967} the changes in pressure and
692: density with respect to a nonrotating star are second order effects
693: in the rotation frequency, and as discussed by \cite{Weber:1992}
694: this remains true even at the limiting Kepler
695: frequency, $\Omega_K$, where the star would shed mass from the
696: equator. The Lagrangian density change during spindown can then be
697: reasonably approximated by the order of magnitude estimate (see also
698: \cite{Fernandez:2005})
699: \begin{equation}\label{drho}
700: \left(\frac{\partial \rho}{\partial \Omega^2}\right)_{a,A}\sim -\frac{\rho}{\Omega_\mathrm{K}^2}
701: \end{equation}
702: Fig.~\ref{OdrdO} shows how the numerically calculated
703: spin down compression rate compares to this estimate in the mixed
704: phase -- we have checked it at other positions as well. For stars with
705: high spin frequencies and no pure quark matter core Eq.~(\ref{drho}) is
706: a reasonable approximation although an overestimate in some
707: regions. It is off by approximately a factor
708: of 0.1 below the frequency where the pure quark matter core
709: forms.
710:
711: \begin{figure}[!htb]
712: \plotone{f6.eps}
713: \caption{Frequency derivative of density scaled to the estimate in
714: Eq.~(\ref{drho}) at $a/A=0.33$ as a function of spin frequency. The
715: estimate is best at high spin frequencies but holds within a factor
716: of $~30$ in all stars.}
717: \label{OdrdO}
718: \end{figure}
719:
720: Eq.~(\ref{drho}) can be used to give a rough estimate of
721: the effects of the additional entropy production related to density
722: change in Eq.~(\ref{dsdt})
723: \begin{eqnarray}\label{dsdt2}
724: T\frac{\mathrm{d}s}{\mathrm{d}t}&=&\frac{c_\mathrm{v}}{\rho}\frac{\mathrm{d}T}{\mathrm{d}t}+T\frac{\mathrm{d}s}{\mathrm{d}\rho}\frac{\mathrm{d}\rho}{\mathrm{d}t}\\
725: &&\simeq\frac{c_\mathrm{v}}{\rho}\frac{\mathrm{d}T}{\mathrm{d}t}-T\frac{\mathrm{d}s}{\mathrm{d}\rho}\frac{\rho}{\Omega_\mathrm{K}^2}\times2\Omega{\dot\Omega}\;.\nonumber
726: \end{eqnarray}
727: Taking the entropy per particle from Eq.~(\ref{s}) this implies that the
728: bulk terms in Eq.~(\ref{dsdt}) are of the order of
729: \begin{align}
730: T\frac{\mathrm{d}\rho}{\mathrm{d}t}\frac{\mathrm{d}s}{\mathrm{d}\rho}&\simeq\frac{2}{3}\frac{\Omega{\dot\Omega}}{\Omega_\mathrm{K}^2}Ts\nonumber\\
731: &\simeq-10^{-15}T_9^2\left(\frac{B}{10^{14}\mbox{
732: G}}\right)^2\\
733: &\times\left(\frac{\Omega/6000}{\mbox{ rad
734: s}^{-1}}\right)^4\left(\frac{\rho}{\mbox{fm}^{-3}}\right)^{-\frac{1}{3}}\mbox{erg s}^{-1}\;,\nonumber
735: \end{align}
736: where we have taken $\Omega_\mathrm{K}=6000\mbox{
737: rad s}^{-1}$, assumed a standard dipole model for the spindown so
738: $\dot{P}=(B_{19}/3.2)^2 P^{-1}$ with the spin period, $P$, measured
739: in seconds and the magnetic field, $B$ in units of $10^{19}\mbox{
740: G}$. As a rough estimate the total heating power from the bulk terms
741: in Eq.~(\ref{dsdt}) in the absence of
742: pairing and for a star of constant density and temperature with a baryon
743: number of $10^{57}$ is then of the order of
744: \begin{eqnarray}
745: W_\mathrm{Bulk}&\sim &10^{42}T_9^2\left(\frac{B}{10^{14}\mbox{
746: G}}\right)^2\left(\frac{\Omega/6000}{\mbox{ rad
747: s}^{-1}}\right)^4\nonumber\\
748: &&\times\left(\frac{\bar{\rho}}{\mbox{ fm}^{-3}}\right)^{-\frac{1}{3}}\mbox{erg s}^{-1}\;. \label{wbulk}
749: \end{eqnarray}
750:
751: This includes only the bulk terms of Eq.~(\ref{dsdt}), but on general
752: grounds one would expect this to be a reasonable approximation of the
753: total additional entropy production. The latent heat has the same
754: basic origin -- the release of entropy with changing density -- and is
755: therefore expected to be of the same order of magnitude as the bulk
756: terms in Eq.~(\ref{wbulk}). In the following section we shall find that
757: this is justified under most, but not all, circumstances.
758:
759: The estimated heating in Eq.~(\ref{wbulk}) is to be compared with the
760: neutrino and photon luminosities which in the absence of pairing
761: phenomena can roughly be estimated as \citep{Page:2006a}
762: \begin{eqnarray}
763: L_\nu^\mathrm{slow}\sim 10^{40}T_9^8\mbox{ erg s}^{-1}\\
764: L_\nu^\mathrm{fast}\sim 10^{45}T_9^6\mbox{ erg s}^{-1}\\
765: L_\gamma\sim 10^{33}T_9^2\mbox{ erg s}^{-1}
766: \end{eqnarray}
767: where 'slow' refers to stars dominated by relatively inefficient
768: neutrino emission processes such as the modified Urca cycle,
769: bremsstrahlung or the pair breaking and formation process while 'fast'
770: refers to stars dominated by the highly efficient direct Urca
771: process.
772:
773: From Eq.~(\ref{wbulk}) we then see that while it may be possible to find
774: combinations of temperature, magnetic field and rotation frequency for
775: which the additional entropy production dominates, such configurations
776: may be difficult to realize in nature. For instance, in order for the
777: heating term in Eq.~(\ref{wbulk}) to dominate the neutrino luminosities
778: the star must be a hot millisecond magnetar. Such an object would be
779: very short lived if it could be created in nature at all. For the
780: heating term to dominate the photon luminosity -- which is the more
781: relevant term at late times and low temperatures -- extremely high
782: magnetic fields would again be required assuming a dipole model for
783: the spin down. Such conditions -- if they were to be realized -- would
784: be very short lived and therefore the more difficult to observe.
785:
786:
787: From these simple estimates we then expect that the release of entropy
788: related to changing bulk density with spin down, the latent heat of
789: phase transitions and changes in the surface and Coulomb energy of
790: mixed phases (smaller still) will have little impact on the thermal
791: evolution of compact stars and be difficult to observe.
792:
793: \section{Spherical Isothermal Cooling} \label{calculations}
794: Estimates cannot replace detailed calculations, and the ones discussed
795: above further ignore all effects of superfluidity, the changing
796: chemical composition and the rapid changes in structure at certain
797: frequencies. We therefore proceed in this section to explore these
798: effects through numerical calculations taking the simplest possible
799: approach to solve the thermal balance in Eq.~(\ref{balance2}).
800:
801: \subsection{Numerical Setup}
802:
803: Neutron stars are commonly taken to be isothermal
804: after an initial thermal relaxation lasting 10-100 years in the
805: sense that the redshifted temperature $T^\infty=Te^\Phi$ is constant
806: below a thermally insulating layer in the outer crust \citep{Gudmundsson:1983}. Since we intend only to investigate the late thermal
807: evolution of neutron stars, and as we do not expect the heating terms
808: discussed here to change the isothermality by disturbing the thermal
809: balance of any part of the star with its surroundings, we shall work
810: within this same approximation. To show the full range of effects,
811: plots in the following also include young stars not expected to be
812: isothermal and our results should only be taken as indicative at such
813: early times.
814:
815: Assuming the star to be thermally relaxed, so the redshifted
816: temperature $T^\infty=Te^\Phi$ is constant below the outer crust, we then
817: integrate Eq.~(\ref{balance2}) to get a simplified equation for global
818: thermal balance
819: \begin{equation}\label{sphbal}
820: \frac{dT^\infty}{dt}=\frac{1}{C}\left(W^\infty-L_\nu^\infty-L_\gamma^\infty\right)
821: \end{equation}
822: with
823: \begin{eqnarray}
824: C&=&\int \mathrm{d}a\, T\frac{\partial s}{\partial T} \label{C}\\
825: W^\infty&=&-\int \mathrm{d}a\, e^{\Phi}T\frac{\partial
826: s}{\partial\rho}\frac{\partial\rho}{\partial\Omega}{\dot \Omega} \label{Wdef}
827: \\
828: L^\infty_\nu&=& \int \mathrm{d}a\, \frac{\epsilon_\nu}{\rho} e^{2\Phi}\\
829: L^\infty_\gamma&=&4\pi R^2\sigma (T_\mathrm{S}^{(0)})^4F(B)e^{2\Phi_s}\label{Lgamma}\\
830: T_{\mathrm{S}_6}^{(0)}&=&
831: g_{14}^{1/4}[(7\zeta)^{2.25}+(\zeta/3)^{1.25}]^{1/4}\\\label{Ts}
832: \zeta&=&T_9-0.001\;g_{14}^{1/4}\sqrt{7T_9}
833: \end{eqnarray}
834: where $\epsilon_\nu/\rho=Q_\nu$ is the neutrino emissivity, the
835: surface temperature is given in units of $10^6$~K,
836: $T_\mathrm{S_6}=T_\mathrm{S}/10^6$~K and $T_\mathrm{S}^{(0)}$ is
837: surface temperature at zero magnetic field strength. We have
838: used a fit appropriate for iron envelopes to the relation between the
839: surface temperature, $T_\mathrm{S}$, and the inner temperature below
840: the heat-blanketing envelope, $T$, in which $g_\mathrm{{14}}$ is the
841: surface gravity in units of $10^{14}$ cm s$^{-2}$
842: \citep{Potekhin:1997,Potekhin:2001}.
843:
844: The heat conductivity
845: in the crust, which derives mainly from the motion of electrons, becomes
846: anisotropic in a strong magnetic field \citep{Potekhin:2001, Geppert:2004}. It is slightly enhanced in the
847: direction along the magnetic field lines and strongly suppressed in
848: the direction orthogonal to the field lines. The heat blanketing
849: relation and the surface temperature
850: then becomes nonuniform,
851: $T_\mathrm{local}(B,\theta,g,T)=T_\mathrm{S}^{(0)}(g,T)\chi(B,\theta,T)$.
852: The photon luminosity must therefore be found by integrating the locally
853: emerging flux and it will depend on magnetic field strength,
854: $L_\gamma(B)=L_\gamma(0)F(B)$. For this purpose we use the fits given
855: by \cite{Potekhin:2001} for $\chi(B,\theta,T)$ and $F(B)$ in a dipole
856: magnetic field. These are good for magnetic fields below $10^{16}$~G,
857: interior temperature between $10^7$~K and $10^9$~K and surface
858: temperature above $10^5$~K. $F(B)$ reaches a minimum of $\sim 0.7$
859: around $B=10^{13}$~G and grows to about 2 at $B=10^{15}$~G for a star
860: of internal temperature $T=10^9$~K. The location of the minimum moves to
861: lower magnetic field strength for lower internal temperature and the
862: change in photon luminosity can delay or accelerate the cooling at
863: late times accordingly. The figures in Sect.~\ref{numericalresults}
864: show the effective surface temperature at infinity
865: $T_\mathrm{S}^\infty=T_\mathrm{S}^{(0)}F(B)^{1/4}e^\Phi$. A strong
866: magnetic field in the inner crust can also lead to anisotropic heat
867: flow there (see e.g. \cite{Geppert:2004} and \cite{Aguilera:2007}),
868: but these effects cannot be included here.\\
869:
870: For given assumptions about the neutrino emissivity and superfluid
871: gaps Eq.~(\ref{sphbal}) is solved along with the spin down model to give
872: the surface temperature at infinity as a function of time, rotation
873: frequency and magnetic field strength. At each time step Eqs.~(\ref{C}) to~(\ref{Ts}) are solved to update the stellar structure at the
874: appropriate rotation frequency.\\
875:
876:
877: Our equations of thermal balance assume spherical symmetry. This
878: effectively treats rotation as a perturbation whose only effects are
879: to change the size and chemical composition of the stars and to
880: provide additional terms in the thermal balance. This approach
881: ignores the effects of nonradial heat flows and nonspherical
882: perturbations of the metric which can be significant when the star is
883: not spherically symmetric at very high spin frequencies (see
884: Fig.~\ref{OmegaR}). Two-dimensional cooling calculations are beyond
885: our scope, but they have been performed at constant rotation frequency
886: by \cite{Schaab:1998a} and \cite{Miralles:1993}. For stars rotating at
887: a large fraction of their Kepler frequency these authors found
888: significant effects on the thermal evolution during the nonisothermal
889: epoch, and polar temperatures up to 31\% higher than the equatorial
890: temperatures even for internally isothermal stars. The recent work of
891: \cite{Geppert:2004, Aguilera:2007, Aguilera:2008} further includes the
892: influence of large magnetic fields in the crust and Joule heating by
893: magnetic field decay in two dimensions. This results in nonradial
894: heat flow and significant heating from magnetic field decay. We focus
895: on the effects of deconfinement during spin down however and shall not
896: pursue these aspects here.
897:
898: Unless otherwise stated we shall use the equatorial radius in
899: Eqs.~(\ref{Lgamma}) to~(\ref{Ts}). We have performed calculations with
900: the polar radius too and will show one such example. At high rotation
901: frequency the surface temperature is then higher and responds less to
902: changes in rotation frequency. We expect results in a two-dimensional
903: code would be intermediate between these and such a calculation would
904: be a significant improvement on what is presented here. With this in
905: mind we use the equatorial radius to show the most pronounced effects
906: of spin down. However, we also note that the strongest effect we have
907: found from spin down stems from the formation of a pure quark matter
908: core. At the rotation frequency where this happens the polar and equatorial
909: radii are very similar (see Fig.~\ref{OmegaR})
910: and a spherical approximation is reasonable.\\
911:
912: Since the stellar moment of inertia $I$ and radius also change with rotation
913: frequency we modify the spin down model described in the previous
914: section to include such effects. The spin down is then determined by
915: \citep{Glendenning:2000}
916: \begin{align}
917: \dot{\Omega}&=-\frac{K}{I}\left[1+\frac{I'\Omega}{2I}
918: \right]^{-1}\Omega^m, \quad
919: I'=\frac{\mathrm{d}I}{\mathrm{d}\Omega}\label{Odot}\\
920: K&=(2/3)R^6 B^2\sin^2{\alpha}
921: \end{align}
922: where $\alpha$ is the inclination angle between the magnetic axis and
923: the rotation axis and $m-1$ is the multipolarity -- usually $m=3$ for
924: magnetic dipole braking or $m=5$ for gravitational quadrupole
925: radiation. We take $\sin^2\alpha=1$ and use the canonical value $m=3$ for a dipole spin down.\\
926:
927: As for the neutrino emissivities and superfluid properties of the
928: neutron star we are interested in the most important effects only and
929: aim for transparency in the model. In the quark phase we
930: include the direct and modified Urca cycles as well as bremsstrahlung
931: using the emissivities in \cite{Blaschke:2000}. The $us$-branch of the
932: direct Urca cycle fulfills the triangle inequality,
933: \begin{equation}
934: k_{\mathrm{F}_s}<k_{\mathrm{F}_u}+k_{\mathrm{F}_e}
935: \end{equation}
936: where $k_{\mathrm{F}_i}$ is the Fermi momentum of the strange-quark, up-quark and electron respectively. It is thus allowed throughout the
937: quark phase, but its $ud$-counterpart is forbidden as it cannot
938: conserve both energy and momentum. This difference stems from the
939: assumed strange quark mass of 150 MeV.
940:
941: In the hadronic phase we again include the direct and modified Urca
942: cycles as well as bremsstrahlung using the emissivities given in
943: \cite{Yakovlev:2001}. We ignore the effects of hyperons here because
944: of their low abundance. In neutron stars the direct Urca reaction is
945: often allowed only at very high densities because it cannot fulfill
946: both energy and momentum conservation unless the proton
947: fraction is above the value where both charge neutrality and the
948: triangle inequality can be observed \citep{Lattimer:1991}. It is
949: critical to note that this is not so in hybrid stars with a mixed
950: phase. The mixed phase is possible precisely because charge neutrality
951: does not have to be observed locally, and it is favorable because the
952: hadronic phase lowers its nuclear symmetry energy by increasing the
953: proton fraction (Fig.~\ref{composition}). Hence the hadronic direct
954: Urca cycle is active in the mixed phase, and as it is about three orders
955: of magnitude faster than the $us$-branch of the quark direct Urca
956: cycle it controls the cooling of hybrid stars with a mixed phase
957: unless reduced by pairing effects.
958:
959: Additional neutrino processes should be included in a more
960: sophisticated treatment -- particularly in the crust -- but as stated
961: above we are mainly interested in the general properties of the
962: additional entropy production and the changing chemical composition
963: and we shall here leave out such terms.\\
964:
965: If there is any attractive interaction among particles in a degenerate
966: Fermi system they will pair, and the resulting superfluidity has
967: important consequences for the thermal properties of neutron star
968: matter and the thermal evolution of neutron stars. Pairing in general
969: delays the cooling because it suppresses most of the neutrino
970: emissivities and enhances the heat capacity at temperatures just below
971: the critical. However it also opens additional neutrino emission
972: processes, suppresses the heat capacity far below the critical
973: temperature and enhances the thermal conductivity, and it may
974: therefore also accelerate the cooling at certain epochs (see
975: e.g. \cite{Yakovlev:1999} for a detailed account).
976:
977:
978: There is considerable uncertainty concerning the relevant superfluid
979: phase in quark matter -- see \cite{Alford:2007a} for a comprehensive
980: review of the effects of pairing among quarks and the possible range
981: of phases. For this first investigation of consequences for the
982: thermal evolution of a quark-hadron phase transition during spin down
983: we assume a simplified but physically transparent model for quark
984: matter pairing which essentially corresponds to pairing in the
985: color-flavor locked phase \citep{Alford:1999,Alford:2001}. In this
986: model each flavor participates equally, the gap $\Delta_{Q,0}$ is
987: independent of density and the critical temperature is
988: $T_c=0.72\Delta_{Q,0}$ (see \cite{Schmitt:2002}). We shall consider a
989: wide range in $\Delta_{Q,0}$ partly because it is interesting for the
990: present calculation, partly because very low gaps may be realized for
991: high strange quark mass as this implies unequal quark Fermi momenta in
992: the unpaired phase. The color-flavor locked phase is generally
993: understood to be so strongly suppressed in all its thermal properties
994: as to be virtually inert with respect to cooling. Since we shall also
995: consider small gaps and since the latent heat derived from the phase
996: transition depends strongly on pairing in the quark phase this will not
997: always apply here, but for strong quark pairing the thermal evolution
998: is generally controlled by the hadronic phase. We further note that as
999: shown by \cite{Jaikumar:2002} (and stressed by \cite{Alford:2007a})
1000: Goldstone modes and their related neutrino emissivities and heat
1001: capacities are not exponentially suppressed, scaling instead as
1002: $T^{15}$ and $T^3$ respectively. They will therefore dominate in
1003: color-flavor locked quark matter at low temperatures, but they are
1004: exceedingly small and do not influence our results.\\
1005:
1006: The effects of superfluidity relevant for our purpose are that below
1007: the critical temperatures it suppresses the neutrino emissivities and
1008: entropies of the participating particles while also allowing
1009: additional neutrino emission through Cooper pair breaking and
1010: formation. The specific heat is enhanced just below the critical
1011: temperature and suppressed at lower temperatures. These effects are
1012: incorporated through control functions, $R(T,T_c)$, which multiply the
1013: relevant quantities. They depend only on the pairing channel,
1014: temperature and gap size, and may therefore be calculated given
1015: expressions for the gap size momentum dependence.\\
1016:
1017: In the quark matter phase we use the simplest possible control
1018: functions and suppress the direct Urca emissivity by a factor of
1019: $e^{-\Delta_\mathrm{Q}/T}$ and the modified Urca and bremsstrahlung
1020: emissivity by a factor of $e^{-2\Delta_\mathrm{Q}/T}$, where
1021: (following \cite{Steiner:2002})
1022: $\Delta_\mathrm{Q}=\Delta_{Q,0}\sqrt{1-(T/T_c)^2}$ is the pairing gap
1023: at the local temperature, $T=T^\infty e^{-\Phi}$.
1024:
1025: The specific heat and entropy of superfluid particles are also
1026: modified at temperatures below the critical temperature. In the
1027: quark matter phase we use the fit to the $^1S_0$ control function for
1028: the specific heat given by \cite{Yakovlev:1999}. We further fit the results of
1029: \cite{Muhlschlegel:1959} for the entropy suppression to obtain
1030: \begin{align}\label{ssf}
1031: s_\mathrm{sf}&=0.95\,s_0\times(T/T_\mathrm{c})e^{-T_\mathrm{c}/T}\nonumber\\
1032: &\times\left[0.43+3.82(T/T_\mathrm{c})-1.41\left(T/T_\mathrm{c}\right)^2\right]\,.
1033: \end{align}
1034: Where $s_0$ is the nonsuperfluid entropy. This expression is applied
1035: to reduce the entropy of both quarks and hadrons (thus here neglecting
1036: the difference between different pairing states of neutrons). The
1037: entropy difference between the superfluid and the normal phase depends
1038: on temperature as $\Delta s\propto - (1-T/T_\mathrm{c})$ close to the
1039: critical temperature \citep{Landau:1980} with $s_0$ becoming
1040: negligible below 0.2$T_\mathrm{c}$. This has the effect of gradually
1041: turning the latent heat in Eq.~(\ref{dsdt}) from a cooling term into a
1042: heating term, but it also suppresses the bulk terms in
1043: Eq.~(\ref{dsdt}).
1044:
1045: In the absence of detailed calculations of control functions for the
1046: entropy of superfluid quark matter phases in the literature we have
1047: employed Eq.~(\ref{ssf}) in the quark matter phase though it is
1048: strictly relevant only for BCS superfluidity. Eq.~(\ref{ssf}) is,
1049: however, consistent with the exponential suppression of entropy below
1050: the critical temperature which would be expected in any such
1051: calculation, and we do not expect that our results are sensitive to
1052: this choice.
1053:
1054: Pairing further opens the important possibility of neutrino emission
1055: by Cooper pair breaking and formation in the quark phase and we
1056: include the associated emissivity as described in \cite{Jaikumar:2001}
1057: using the fit to the $^1S_0$ control function in
1058: \cite{Yakovlev:1999}.\\
1059:
1060: Among the nucleons pairing is predicted for neutrons in the $^1$S$_0$
1061: and $^3$P$_2$ channels and for protons in the $^1$S$_0$ channel. The
1062: gaps can be calculated self-consistently but results are uncertain and
1063: vary greatly both in terms of the maximum gap size and in terms of
1064: density dependence -- see e.g. \cite{Lombardo:2001,
1065: Page:2004,Page:2006a} for collections of these results. We shall
1066: here use gaps with a phenomenological momentum dependence as suggested
1067: by \cite{Kaminker:2001}, \cite{Andersson:2005} and recently
1068: \cite{Aguilera:2007}
1069: \begin{align}
1070: \Delta(k_{F,N})=\Delta_0&\frac{(k_{F,N}-k_0)^2}{(k_{F,N}-k_0)^2+k_1^2}\\
1071: &\times\frac{(k_{F,N}-k_2)^2}{(k_{F,N}-k_2)^2+k_3^2}\label{deltaofk}
1072: \end{align}
1073: where $k_{F,N}$ is the Fermi momentum of the relevant nucleon,
1074: $N=n,p$. Our choices for the parameters $\Delta_0$ and $k_{i}$ correspond to sets $a$, $e$ and $h$ of
1075: \cite{Aguilera:2007} (see that work for references to the model
1076: calculation underlying these fits). Eq.~(\ref{deltaofk}) is valid only
1077: in the range $k_0<k<k_2$ with $\Delta(k_{F,N})=0$ outside this range,
1078: and where the gaps for $^1$S$_0$ and $^3$P$_2$ pairing of neutrons are
1079: both nonzero we use the largest of the two gaps. We further use the
1080: relations between critical temperature and pairing gap listed by
1081: \cite{Yakovlev:1999}.\\
1082:
1083:
1084: In the hadronic phase neutrons and protons may pair simultaneously in
1085: different channels and so the resulting calculations for the control functions can be quite involved. We use the fits to numerically
1086: calculated control functions for each process and each pairing channel
1087: compiled in \cite{Yakovlev:1999} (or
1088: \cite{Yakovlev:2001} -- see these works for detailed
1089: references). Where protons and neutrons pair simultaneously we use the
1090: combined factors listed in these works if available and if not then
1091: the smallest of the two independent control functions.
1092:
1093: The hadronic pair breaking and formation process can be very powerful
1094: and may dominate the thermal evolution when pairing first sets in, but
1095: it was recently shown by \cite{Leinson:2006} to be strongly suppressed
1096: in the singlet state channel by approximately a factor of $10^{-6}$
1097: relative to the $^3$P$_2$ channel (see also the recent work of \cite{Steiner:2008}). In the hadronic phase we therefore
1098: include pair breaking and formation only for neutrons at high
1099: densities where they pair in the $^3$P$_2$ channel and for this we
1100: again use the emissivity and control function given in
1101: \cite{Yakovlev:1999,Yakovlev:2001}.
1102:
1103: \subsection{Numerical Results}\label{numericalresults}
1104: Figures~\ref{Tt} to~\ref{TtDelta} explore consequences for the
1105: thermal history of hybrid stars of including time dependent structure
1106: and entropy production by spin down
1107: compression in the cooling calculations with the approximations discussed above. They show the
1108: surface temperature and additional entropy production at infinity as
1109: functions of time, magnetic field and quark pairing gap energy.
1110:
1111: \begin{figure}[!htb]
1112: \plotone{f7.eps}
1113: \caption{The effect of spin down compression on cooling curves for
1114: stars with initial spin frequency $\Omega_\mathrm{start}=6000 \mbox{ rad
1115: s}^{-1}$ and constant magnetic field as labeled for each curve. The
1116: quark core is assumed superconducting with $\Delta_{Q,0}=10$
1117: MeV. The increasing quark matter fraction produces slower cooling
1118: and a jump in temperature with the appearance of the pure quark
1119: matter phase.}
1120: \label{Tt}
1121: \end{figure}
1122:
1123: In Fig.~\ref{Tt} we show the thermal evolution of stars with quark
1124: pairing gap $\Delta_{Q,0} = 10$~MeV, a range of magnetic fields and
1125: initial spin period around 1~ms. The extremely rapid initial rotation
1126: is interesting because it means the stars start out with no pure quark
1127: matter core and undergo drastic changes in structure as a pure quark
1128: matter core develops around $\Omega=1400$ rad s$^{-1}$. We see this in
1129: Fig.~\ref{Tt} as a short period of increasing temperature at the time
1130: corresponding to this angular velocity for a given magnetic
1131: field. During this short period the heating term from the additional
1132: entropy production actually dominates which is shown in more detail in
1133: Fig.~\ref{Wtscaled} discussed below. Further, the changing neutrino
1134: emissivity and heat capacity with the introduction of more quarks in
1135: the star and the reduction in stellar surface area as the pure quark matter
1136: core develops all affect the thermal evolution. This implies somewhat
1137: slower cooling -- with our specific choice of parameters -- and higher
1138: surface temperature. For the internal temperature, $T$, we have found
1139: essentially a transition between two otherwise similar cooling tracks;
1140: one for stars with no pure quark matter core and one for stars with a
1141: fully developed quark matter core. The effective surface temperature
1142: also depends on magnetic field strength through $F(B)$ however. For
1143: this reason such an easy interpretation is difficult from
1144: Fig.~\ref{Tt} alone.\\
1145:
1146: \begin{figure}[!htb]
1147: \plotone{f8.eps}
1148: \caption{Variation of surface temperature at infinity with magnetic
1149: field for stars at ages $10^2$, $10^4$ and $10^5$ years from
1150: above. Thick lines have initial spin frequency $\Omega_\mathrm{start}=6000
1151: \mbox{ rad s}^{-1}$ while thin dashed lines start with
1152: $\Omega_\mathrm{start}=600 \mbox{ rad s}^{-1}$. Thin continuous lines use the
1153: polar rather than the equatorial radius. The quark core is assumed
1154: superconducting with $\Delta_{Q,0}=10$~MeV. We note a slow increase
1155: in temperature for old stars and a sudden jump at the magnetic field corresponding
1156: to significant spin down at specific ages. The suppression in
1157: temperature at low magnetic field and increase at high field is due
1158: to the effects of the magnetic field of the heat blanketing relation.}
1159: \label{BT}
1160: \end{figure}
1161:
1162:
1163: The magnetic field strength determines the spin down and affects the
1164: heat blanketing relation. Fig.~\ref{BT} shows how the temperature varies at
1165: specific ages with the magnetic field strength (which is kept constant in time
1166: itself). This is shown for initial spin frequencies of 6000 rad
1167: s$^{-1}$ and 600 rad s$^{-1}$ and using the polar radius in addition to
1168: the equatorial when calculating the surface temperature.
1169:
1170: For low initial spin frequency
1171: we find no significant effects of spin down in Fig.~\ref{BT}. The
1172: change in temperature at a specific age
1173: with increasing magnetic field is due to the effects of the magnetic
1174: field on the heat blanketing relation and photon luminosity. If these
1175: effects were left out the curves for low initial spin frequency would
1176: be almost constant.
1177:
1178: If the initial rotation frequency is sufficiently high to exclude the
1179: pure quark matter phase it is a different matter. A strong magnetic
1180: field initially suppresses the surface temperature of young stars with
1181: high rotation frequencies, but it then gives a sharp jump at the
1182: magnetic field strength which ensures that a pure quark matter core is
1183: formed at a particular age. For older stars there is also a slight
1184: increase in surface temperature before the formation of a pure quark
1185: matter core when the equatorial radius is used. If the effects of the
1186: magnetic field on the heat blanketing relation were left out the
1187: curves at high initial spin frequency become nearly constant for
1188: $B>10^{13}$~G and for fields too weak for spin down to set in at the
1189: particular age. In this case there would still be a jump in
1190: temperature at the magnetic field strength corresponding to formation
1191: of a pure quark matter core for a specific age however.
1192:
1193:
1194: Replacing the equatorial radius with the polar in Eqs.~(\ref{Lgamma})
1195: and~(\ref{Ts}) (thin continuous lines in Fig.~\ref{BT}) we find higher
1196: surface temperatures before the introduction of pure quark matter in
1197: the core and still a clear jump in temperature after. The gradual
1198: increase in temperature before the jump found in the other curves is
1199: absent except for old stars. This can be understood if we remember
1200: that the polar radius is smaller than the equatorial and less
1201: sensitive to spin down at high rotation frequencies. We have checked that a similar pattern may
1202: be seen in plots of temperature versus time using the polar
1203: radius. We expect that full two-dimensional calculations would
1204: give results which are at high rotation frequencies intermediate
1205: between what we have found using a spherical approximation. The similarity between curves with polar and equatorial radius
1206: at the jump in temperature shows that our calculations are insensitive
1207: to the approximation at the spin frequency where the pure quark
1208: matter core forms and where we find the strongest effect of spin down.\\
1209:
1210:
1211:
1212: \begin{figure}[!htb]
1213: \plotone{f9.eps}
1214: \caption{Numerical value of the entropy production as defined in
1215: Eq.~(\ref{Wdef}) as a function of time for a star with magnetic field
1216: $B=10^{11}$~G and quark pairing gap $\Delta_{Q,0}=10$~MeV. $W_\mathrm{Tot}$ is split into
1217: contributions with different physical origins as explained in the
1218: text. The surface and Coulomb term changes sign and the negative of
1219: this term is shown as well. The latent heat and bulk terms are dominant, and the spike is
1220: caused by the appearance of a pure quark matter phase.}
1221: \label{Wt}
1222: \end{figure}
1223:
1224: \begin{figure}[!htb]
1225: \plotone{f10.eps}
1226: \caption{The total entropy production scaled to the combined neutrino
1227: and photon luminosity for stars with quark paring gap
1228: $\Delta_{Q,0}=10$~MeV and magnetic fields as indicated just above
1229: each curve. The different line styles are meant only to help guide
1230: the eye. Note that at the peaks reaching above 1 the entropy
1231: production actually dominates the cooling terms and the
1232: temperature increases. Also note that this is only possible for
1233: stars changing structure rapidly with the introduction of a pure
1234: quark matter core.}
1235: \label{Wtscaled}
1236: \end{figure}
1237:
1238: In Figs.~\ref{Wt} and \ref{Wtscaled} we explore the importance of the
1239: heating term in a little more detail. Fig.~\ref{Wt} shows the
1240: numerical value of $W^\infty$ and its various contributions as a
1241: function of time for a star with pairing gap $\Delta_{Q,0}=10$~MeV and
1242: magnetic field $B=10^{11}$~G. The total entropy production,
1243: $W^\infty$, is split into contributions corresponding to (integrals
1244: of) the terms in Eq.~(\ref{dsdt}). $W^\infty_\textrm{Latent}$
1245: corresponds to the latent heat, $W^\infty_\textrm{Hadronic bulk}$ to the
1246: hadronic bulk term, $W^\infty_\textrm{Quark bulk}$ to the quark bulk
1247: term and $W^\infty_\textrm{SC}$ to the surface and Coulomb term of
1248: Eq.~(\ref{dsdt}). Leptons do not participate in the deconfinement
1249: transition, so their entropy is assumed continuous across the phase
1250: boundary. They are separated from the other particles in the form of
1251: $W^\infty_\textrm{Lepton}$ in Fig.~\ref{Wt}.
1252:
1253: The bulk and lepton
1254: contributions always act as heating terms while the surface and
1255: Coulomb term changes sign and turns into a heating term with the onset
1256: of hadronic superfluidity around age 4000 years with the hadronic bulk
1257: term simultaneously falling away. The quark bulk term is very small
1258: because for a pairing gap of 10 MeV the quark matter is strongly
1259: superconducting at all times except the very early and the quark
1260: entropy is therefore suppressed. This also implies, however, that
1261: hadrons making the transition across the phase boundary essentially
1262: release all their entropy, and so the latent heat, which changes sign
1263: and becomes a heating term with the onset of quark superfluidity, is
1264: therefore quite significant and actually comes to dominate around the
1265: time when the hadronic bulk term falls away. The lepton term and the surface
1266: and Coulomb term are relatively small but significant when the bulk
1267: terms are eliminated through the onset of pairing. The peak in the
1268: total around age
1269: 20000 years corresponds to the appearance of a pure quark matter phase
1270: in the core and the ensuing rapid changes in structure, and it
1271: contributes to produce the short interval of increasing temperature
1272: seen in Fig.~\ref{Tt}.
1273:
1274: Fig.~\ref{Wtscaled} shows the total entropy production scaled to the
1275: total neutrino and photon luminosity, $W^\infty/(L^\infty _\gamma
1276: +L^\infty _\nu)$, for the same set of parameters which gave the
1277: cooling curves in Fig.~\ref{Tt}. The strong temperature dependence of
1278: the neutrino and photon luminosities ensures that the scaled heating
1279: term does not increase above unity where heating and cooling exactly
1280: balance -- except at the peaks caused by the appearance of a pure
1281: quark matter core. We also note that as in Fig.~\ref{Tt} the magnetic
1282: field must be above $\sim 10^{9}$~G for $W^\infty$ to have any
1283: discernible effects.
1284:
1285:
1286: \begin{figure}[!htb]
1287: \plotone{f11.eps}
1288: \caption{The total entropy production scaled to the combined neutrino
1289: and photon luminosity for stars with quark paring gap as indicated
1290: in the legend and $\Omega_\mathrm{start}=6000 \mbox{ rad s}^{-1}$. The
1291: magnetic field is constant at $10^{11}$~G. The entropy production
1292: changes sign and becomes negative for low pairing gaps and thus acts
1293: as a powerful cooling term during some epochs.}
1294: \label{WtDelta}
1295: \end{figure}
1296:
1297: While the latent heat for a transition to a strongly superfluid core
1298: is a heating term, the transition to a
1299: nonsuperfluid core with higher entropy per baryon cools the
1300: star. This is illustrated in Fig.~\ref{WtDelta} where we plot the
1301: total heating term scaled to the combined neutrino and photon
1302: luminosities at constant magnetic field for a range of quark pairing
1303: gaps $\Delta_{Q,0}$. Here we see how the entropy production changes
1304: from a heating term to a cooling term and may be quite dominant at
1305: certain times depending on the choice of stellar parameters. This can
1306: also be observed in Fig.~\ref{TtDelta} where we plot surface
1307: temperature versus age with constant magnetic field for a range of
1308: quark pairing gaps. The onset of superfluidity initially delays the
1309: cooling and then accelerates it. When the pure quark matter core forms
1310: there is a drop in temperature for very low pairing gaps and a jump
1311: for high pairing gaps. In fact, however, the major
1312: differences between the curves in Fig.~\ref{TtDelta} are due to the
1313: differences in the thermal evolution of superfluid and nonsuperfluid
1314: quark matter rather than spin down related effects. Old stars with a
1315: nonsuperfluid quark matter core cool more slowly than those without a
1316: quark matter core.
1317:
1318: \begin{figure}[!htb]
1319: \plotone{f12.eps}
1320: \caption{The effect of spin down compression on cooling curves for
1321: stars with initial spin frequency $\Omega_\mathrm{start}=6000 \mbox{ rad
1322: s}^{-1}$ and quark pairing gap $\Delta_{Q,0}$ as labelled for each
1323: curve. The magnetic field is constant at $10^{11}$~G}
1324: \label{TtDelta}
1325: \end{figure}
1326:
1327: A rich picture has now emerged from the coupling between the thermal
1328: evolution and spin down compression with deconfinement. The spin down
1329: may change the thermal evolution of hybrid stars by
1330: heating and cooling them at various ages and by changing the structure
1331: and chemical composition -- or it may be entirely inconsequential
1332: depending on quark superfluidity and initial spin frequency. The
1333: effects we found turned out to depend strongly on the inclusion of
1334: quark pairing and the timing of the appearance of the pure quark
1335: matter phase in the stellar core. These are subject to the
1336: specific assumptions made concerning the equation of state, stellar
1337: baryon number and pairing regime, so it is essential to stress that
1338: the results discussed in the present section illustrate the general
1339: point that spin down and thermal cooling may be interdependent, rather
1340: than providing quantitatively reliable predictions.\\
1341:
1342: Bearing this in mind we show in Fig.~\ref{Ttwdata} how calculations
1343: for selected pairing gaps and magnetic field strengths measure up to
1344: observational data on the thermal state of isolated neutron stars
1345: (kindly supplied by Dany Page). This figure is shown partly as a
1346: consistency check for our calculations and partly to compare the size
1347: of the effects we have found with the accuracy of actual
1348: measurements. We stress that the observational data are consistent
1349: with static cooling models \citep{Page:2004} and that we do not
1350: suggest that the effects of spin down are necessary to explain
1351: them. Further the effects we have found occur after thermal relaxation
1352: only for magnetic field strength below $\sim 10^{12}$~G. For those
1353: sources in Fig.~\ref{Ttwdata} for which the field is known it is above
1354: that value \citep{Pons:2007}.
1355:
1356: Fig.~\ref{Ttwdata} explores a range of quark pairing gaps which in our
1357: pairing scheme are density independent. In nature the pairing gap
1358: would not span a range this broad, but it could be density dependent
1359: and variations between stars in mass and density would then introduce
1360: variations in their thermal evolution. Similarly stars with different
1361: masses would have a different phase structure and acquire pure quark
1362: matter cores at different rotational frequencies thus further
1363: complicating the picture.
1364:
1365: Because the hadronic direct Urca mechanism and the neutron pair
1366: breaking and formation mechanism are active in the mixed phase our
1367: models are consistent only with the cold sources and the effects of
1368: spin down do not change this conclusion. That these mechanisms are the
1369: cause of the generally low temperatures in our models can be seen if
1370: they are artificially left out. The two dotted lines in
1371: Fig.~\ref{Ttwdata} show -- for illustration only -- that temperatures
1372: are much higher if these mechanisms are suppressed. Such a suppression
1373: could occur for certain quark pairing schemes where pairing forces the
1374: quark phase to become electrically neutral or positive rather than
1375: negative, thereby reducing the proton content in the hadronic phase
1376: below the threshold for direct Urca. A self-consistent inclusion of
1377: such effects are beyond the scope of the present investigation. We
1378: stress again that the inconsistency with hot sources seen in this
1379: figure is subject to the choice of stellar and thermal models made for
1380: the specific purpose of studying the effects of deconfinement during
1381: spin down. Thus it should not be seen as indicating a general
1382: breakdown in cooling theory.
1383:
1384:
1385: The true signal of deconfinement we have found is in the transition
1386: between cooling curves with and without pure quark matter cores, and
1387: in the brief period of increasing or rapidly decreasing temperatures
1388: thus caused for certain stellar parameters. As can be seen in
1389: Fig.~\ref{Ttwdata} the change in surface temperature caused by
1390: deconfinement is smaller than the error bars on the observational data
1391: and it will be difficult to test observationally on the basis of data
1392: relating only temperature and time.
1393:
1394: \begin{figure}[!htb]
1395: \centering
1396: \plotone{f13.eps}
1397: \caption{Comparison between observational data and cooling
1398: calculations with spin down. The stars have initial spin frequency
1399: $\Omega_\mathrm{start}=6000 \mbox{ rad s}^{-1}$, magnetic field
1400: strength $B=10^{11}$~G and quark pairing gaps as indicated. The
1401: observational data can be found in \cite{Page:2004} and sources are
1402: identified by numbers right above or below their temperature
1403: error bars. The two dotted lines illustrate slow cooling with
1404: artificial suppression of neutrino emission by direct Urca and
1405: neutron pair breaking and formation. They have $\Delta_{Q,0}=10$~MeV
1406: (thick line) and 0.01 MeV (thin line). Including these mechanisms
1407: only cold sources can be explained. The jump in temperature at the
1408: formation of a pure quark matter core is smaller than the error bars
1409: on the observational data.}
1410: \label{Ttwdata}
1411: \end{figure}
1412:
1413: \section{Discussion}\label{discuss}
1414: Our intention with the present work was to explore a possible
1415: connection between the spin down and thermal cooling of hybrid stars
1416: with a deconfined quark matter core -- the appearance of which might be
1417: expected to influence theoretical cooling tracks. Specifically we were
1418: interested in the latent heat of the phase transition and the drastic
1419: changes in structure and chemical composition resulting from the
1420: increasing density with spin down. The general formalism worked out in
1421: Sect.~\ref{thequilibrium} would also be relevant to models with no
1422: phase transition, however, as the bulk terms in the spin down derived
1423: entropy production might also be important in such cases. While
1424: estimates of the importance of the spin down derived entropy
1425: production did not give cause to expect strong signals of
1426: deconfinement in the thermal evolution, numerical calculations
1427: revealed clear effects of the changes caused by the appearance of a
1428: pure quark matter core. This signature is related to changes in
1429: radius, chemical composition, structure and under certain
1430: circumstances the additional entropy production both in bulk and in
1431: the form of latent heat.\\
1432:
1433: Our numerical work was carried out using a sophisticated and reliable
1434: code with respect to the stellar structure solving Hartles
1435: perturbative equations for the structure of rotating compact stars
1436: self-consistently. The thermal evolution, however, was treated in a
1437: spherical isothermal approximation with a somewhat ad hoc approach to the
1438: treatment of superfluid pairing in the quark matter phase. Still we believe our work is sufficiently detailed to demonstrate the
1439: general point that important signals may be derived from the interplay
1440: between spin down and cooling of compact stars, which could in the
1441: long run help answer pressing questions about the state of matter at
1442: high densities. The signals we have found do not dominate the general thermal
1443: evolution of hybrid stars but they do complement the standard
1444: picture. By correlating temperature with magnetic field strength and
1445: spin frequency they may also help break the degeneracy between models relating
1446: only temperature and time.\\
1447:
1448:
1449: The signature of
1450: deconfinement found here is below the present observational
1451: sensitivity and not of sufficient strength to set apart the cooling
1452: curves with temperature versus time for hybrid stars. The correlation
1453: between temperature and magnetic field strength provides a possible
1454: alternative. To test such a correlation it would be necessary to
1455: obtain accurate temperatures for a number of stars of approximately
1456: the same age for which the spin down could also be
1457: detected by the emission of pulses in either radio or X-ray. It might
1458: then be possible to test for features in the relation between
1459: temperature and magnetic field -- or equally interesting, the rotation
1460: frequency itself -- which as demonstrated here could result from a
1461: strong phase transition if the initial rotation frequency was
1462: sufficiently high to have spun out the high density phase.
1463:
1464: Contemplating these prospects it is important to consider that
1465: alternative effects not treated here might determine the relation
1466: between spin down and thermal evolution. Most pressingly the heat from
1467: magnetic field decay was employed by \cite{Pons:2007} to explain the
1468: observed correlation between temperature and magnetic field and it may
1469: well drown out any other effects -- although as noted by the authors
1470: the magnetic field decay itself has not yet been independently
1471: demonstrated. Recent work by \cite{Aguilera:2007} and
1472: \cite{Aguilera:2008} on the cooling of magnetized neutron stars in two
1473: dimensions also highlighted the importance of the magnetic field
1474: strength, geometry and possible decay for the thermal evolution of
1475: neutron stars. These authors find strongly anisotropic surface
1476: temperature distributions and possibly an inverted temperature
1477: distribution with hot equatorial regions for middle aged stars. They
1478: further show that the effects of magnetic field decay and Joule
1479: heating can dominate the thermal evolution at strong and intermediate
1480: field strengths -- even to the point that the effects of the direct
1481: Urca process may be hidden by this heating term in magnetars.
1482:
1483:
1484: The work of \cite{Reisenegger:1995} and \cite{Fernandez:2005} is also
1485: most important. These authors consider the second term on the right
1486: hand side of Eq.~(\ref{balance2}) which is shown to give rise to
1487: so-called roto-chemical heating as the weak interactions required to
1488: maintain chemical equilibrium are unable to keep pace with the
1489: increasing density. \cite{Fernandez:2005} found that old millisecond
1490: pulsars reach a quasi-equilibrium in which the photon luminosity is
1491: determined entirely by the spin down power and remains much higher
1492: than otherwise attainable. Their work considered only nucleonic
1493: equations of state but similar results should clearly be expected for
1494: hybrid stars just as part of the effects demonstrated here should be
1495: expected to show up in nucleonic stars.\\
1496:
1497:
1498: Given the possible importance of such alternatives and the
1499: shortcomings of the present study discussed above, a more complete
1500: treatment of the interplay between the spin down and thermal evolution
1501: of neutron stars seems desirable before strong assertions can be made
1502: concerning the specific shape of cooling tracks. As well as including
1503: all possible energy sources and sinks this should be extended to
1504: consider a wider range of stellar masses, equations of state, phase
1505: transitions and pairing regimes. A two dimensional code could further
1506: treat the influence of the magnetic field on heat flows in the inner
1507: crust and the effects of nonspherical heat flows for rotationally
1508: perturbed stars -- aspects which are all
1509: independently well described in the literature.
1510:
1511:
1512: It should further be noted that the rotation frequency at which
1513: the pure quark matter core appears depends strongly on the stellar
1514: baryon number and the equation of state, and it may well be lower than
1515: discussed here thus changing the timing of spin down related
1516: effects. The strongest signature of deconfinement found here depends
1517: entirely on the formation of a pure quark matter core where none
1518: existed previously. If, on the other hand, a pure quark matter phase
1519: is present in the core of stars at even the highest rotation
1520: frequencies -- as is the case for other equations of state -- we
1521: would not expect equally strong signals from deconfinement with spin
1522: down compression.
1523:
1524: It is possible that circumstances not considered here could
1525: provide a more pronounced link between rotation and cooling. If for
1526: instance the magnetic field varies over time -- and as discussed by
1527: \cite{Geppert:2006} this may well be the case -- the timing of any
1528: strengthening or decay of the magnetic field may provide entirely
1529: different spin down histories. Alternatively the temperature of stars
1530: being spun up in accreting binaries should to some extent be
1531: determined by the changing chemical composition and this could provide
1532: an alternative approach.
1533:
1534: Additional energy release would also be expected if the
1535: phase transition cannot maintain thermodynamic or hydrostatic
1536: equilibrium as was assumed here. As the most extreme example of this,
1537: we have entirely ignored the instabilities and corequakes which might
1538: accompany the appearance of a pure quark matter phase in a star far
1539: from equilibrium \citep{Zdunik:2006}. Such events could release large
1540: amounts of energy and significantly change the thermal
1541: evolution by reheating the star.\\
1542:
1543: While many essential issues thus remain poorly explored we hope to
1544: have demonstrated the possible usefulness of considering the
1545: connection between the spin down and thermal evolution of neutron
1546: stars. Although the results of such work may
1547: not be immediately applicable to observational tests, we believe they
1548: could prove most valuable in the long run.\\
1549:
1550:
1551: We wish to acknowledge helpful conversation with Sanjay Reddy on the
1552: hadronic pair breaking and formation process and to thank Dany Page
1553: for the use of his compilation of observational data. We are also
1554: grateful to Dima Yakovlev and Oleg Gnedin for providing us with an
1555: earlier version of their cooling code, which we took advantage of when
1556: developing the code used for this study.\\
1557: M. Stejner wishes to thank San Diego State University for its
1558: hospitality during part of this work.\\
1559: The research of F. Weber is supported by the National Science
1560: Foundation under Grant PHY-0457329, and by the Research Corporation.\\
1561: J. Madsen is supported by the Danish Natural Science Research Council.
1562:
1563: \appendix \section{Appendix A, Derivation of the Energy Balance in Mixed
1564: Phases}\label{appendix}
1565:
1566: To derive Eq.~(\ref{balance2}) from
1567: Eq.~(\ref{balance}) we proceed analogously to \cite{Thorne:1966} and
1568: \cite{Weber:1999} but work in terms of baryon number $a$ instead of
1569: radial coordinate. We first define the nuclear energy generation rate
1570: per baryon as the rate, $q$, at which rest mass is converted to
1571: internal energy as measured by an observer in the shell and at rest
1572: with respect to the shells reference frame
1573: \begin{equation}\label{first}
1574: q=-\frac{\mathrm{d}\bar{m}_\mathrm{B}}{\mathrm{d}(\mbox{proper
1575: time})}=-\frac{\mathrm{d}\bar{m}_\mathrm{B}}{e^\Phi \mathrm{d}t}\; ,
1576: \end{equation}
1577: where $\bar{m}_\mathrm{B}$ is the average rest mass per baryon and we
1578: work in units with. The
1579: first term on the right hand side of Eq.~(\ref{balance}) is the amount
1580: of rest mass converted to internal energy in the entire shell which is
1581: then $q\delta a e^\Phi \mathrm{d}t$.
1582:
1583: Next we note that by expressing the shells volume in terms of its
1584: particle density, $V=(V/\delta a)\delta a=\rho/\delta a$, the second term in
1585: Eq.~(\ref{balance}) can be written
1586: \begin{equation}\label{second}
1587: \mathrm{d}W=-P\mathrm{d}\left(\frac{1}{\rho}\delta a\right) + \alpha \mathrm{d}\mathbb{S}+\mathrm{d}E_\mathrm{C}
1588: \end{equation}
1589:
1590: The third term represents the difference between the rates at which
1591: energy enters and leaves the shell by radiative or conductive means as
1592: measured by an observer in the shell and at rest with respect to the
1593: shells reference frame, which may be written
1594: \begin{eqnarray} \label{third}
1595: L_\mathrm{tot}(a+\delta a)e^{2(\Phi(a+\delta
1596: a)-\Phi(a))}-L_\mathrm{tot}(a)&=&L_\mathrm{tot}(a+\delta
1597: a)\left(1+2\frac{\mathrm{d}\Phi}{\mathrm{d}a}\delta
1598: a\right)-L_\mathrm{tot}(a)\nonumber\\
1599: &=&\left(\frac{\mathrm{d}L_\mathrm{tot}}{\mathrm{d}
1600: a}+2L_\mathrm{tot}\frac{\mathrm{d} \Phi}{\mathrm{d} a}\right)\delta a\nonumber\\
1601: &=&\frac{\mathrm{d}}{\mathrm{d}
1602: a}\left(L_\mathrm{tot}e^{2\Phi}\right)e^{-2\Phi}\delta a
1603: \end{eqnarray}
1604: and the energy change during a coordinate time interval, d$t$, is this
1605: times the proper time interval $e^\Phi \mathrm{d}t$. The two factors
1606: of $e^{\Phi(a+\delta a)-\Phi(a)}$ account for redshift and time
1607: dilation, respectively, as the energy crosses from the inner to the outer
1608: edge of the shell, and we expanded the exponential to order $
1609: \mathcal{O}(\delta a)$ as
1610: \begin{equation}
1611: e^{2(\Phi(a+\delta a)-\Phi(a))}\approx 1+2[\Phi(a+\delta a)-\Phi(a)]=
1612: 1+2\frac{\mathrm{d}\Phi}{\mathrm{d} a}\delta a.
1613: \end{equation}
1614:
1615: Using Eqs.~(\ref{first}), (\ref{second}) and (\ref{third}) in
1616: Eq.~(\ref{balance}) and expressing the change in internal energy in
1617: terms of the density of internal energy, $\mathcal{E}_\mathrm{int}$,
1618: local energy balance can then be written as
1619: \begin{equation}
1620: \mathrm{d}\left(\frac{\mathcal{E}_\mathrm{int}}{\rho}\delta
1621: a\right)=qe^\Phi\delta a \mathrm{d}t -
1622: P\mathrm{d}\left(\frac{1}{\rho}\delta a\right) + \alpha
1623: \mathrm{d}\mathbb{S} + \mathrm{d}E_\mathrm{C} - \left[\frac{\mathrm{d}}{\mathrm{d}
1624: a}\left(L_\mathrm{tot}e^{2\Phi}\right)\right] e^{-\Phi}\delta a \mathrm{d}t\;.
1625: \end{equation}
1626: Rearranging and noting that $\delta a$ is constant in time we then get the
1627: gradient of $L_\mathrm{tot}$
1628: \begin{eqnarray}\label{cool1}
1629: \frac{\mathrm{d}}{\mathrm{d}
1630: a}\left(L_\mathrm{tot}e^{2\Phi}\right)=e^{2\Phi}\bigg\{q+\bigg[\frac{Pe^{-\Phi}}{\rho^2}\frac{\mathrm{d}\rho}{\mathrm{d}t}&+&\frac{\alpha
1631: e^{-\Phi}}{\delta
1632: a}\frac{\mathrm{d}\mathbb{S}}{\mathrm{d}t}\nonumber\\ &+&\frac{e^{-\phi}}{\delta
1633: a}\frac{\mathrm{d}E_\mathrm{C}}{\mathrm{d} t}-e^{-\Phi}\frac{\mathrm{d}}{\mathrm{d}t}\left(\frac{\mathcal{E}_\mathrm{int}}{\rho}\right)\bigg]_{a=\mathrm{constant}}\bigg\}
1634: \end{eqnarray}
1635:
1636: Eq.~(\ref{cool1}) is the relation we sought giving the luminosity
1637: gradient for stars with variable structure and including the
1638: contributions from surface and Coulomb terms in the energy density. It
1639: may be considerably simplified however by noting that we can write the
1640: internal energy density in terms of the total energy density as
1641: $\mathcal{E}_\mathrm{int}/\rho=\epsilon/\rho-\bar{m}_\mathrm{B}$ in
1642: which case from the definition of $q$
1643: \begin{equation}
1644: q-e^{-\Phi}\frac{\mathrm{d}}{\mathrm{d}t}\left(\frac{\mathcal{E}_\mathrm{int}}{\rho}\right)=
1645: -\frac{e^{-\Phi}}{\rho}\left[\frac{\mathrm{d}\epsilon}{\mathrm{d}t}-\frac{\epsilon}{\rho}\frac{\mathrm{d}\rho}{\mathrm{d}t}\right]\; .
1646: \end{equation}
1647: Inserting this in Eq.~(\ref{cool1}) and using the first law as written
1648: in Eq.~(\ref{firstlaw}) finally gives the expression
1649: \begin{eqnarray}\label{cool2}
1650: \frac{\mathrm{d}}{\mathrm{d}
1651: a}\left(L_\mathrm{tot}e^{2\Phi}\right)&=&-\frac{e^{\Phi}}{\rho}\left[\frac{\mathrm{d}\epsilon}{\mathrm{d}t}-\frac{\epsilon+P}{\rho}\frac{\mathrm{d}\rho}{\mathrm{d}t}
1652: -\rho\frac{\alpha}{\delta a}\frac{\mathrm{d}\mathbb{S}}{\mathrm{d}t}
1653: -\rho\frac{1}{\delta a}\frac{\mathrm{d}E_\mathrm{C}}{\mathrm{d}
1654: t}\right]_{a=\mathrm{constant}}\\
1655: &=&-e^\Phi\left[T\frac{\mathrm{d}s}{\mathrm{d}t}+\sum_k\mu_k\frac{\mathrm{d}Y_k}{\mathrm{d}t}\right]_{a=\mathrm{constant}}
1656: \; .
1657: \end{eqnarray}
1658:
1659: In this discussion we have used baryon number as the
1660: independent variable to emphasize the importance of a Lagrangian
1661: description for stars with variable structure. This is of course
1662: not required for all applications and for comparison with other
1663: discussions we note that our expressions can be converted to use radial
1664: coordinate as the independent coordinate through the standard relation
1665: \begin{equation}
1666: \mathrm{d} a=4\pi r^2\rho\left(1-\frac{2m}{r}\right)^{-\frac{1}{2}}\mathrm{d}r
1667: \end{equation}
1668:
1669: %\bibliographystyle{aa}
1670: %\bibliography{/Users/Morten/Documents/Bibliography/bibliography}
1671: \begin{thebibliography}{69}
1672: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
1673:
1674: \bibitem[{{Aguilera} {et~al.}(2008){Aguilera}, {Pons}, \&
1675: {Miralles}}]{Aguilera:2007}
1676: {Aguilera}, D.~N., {Pons}, J.~A., \& {Miralles}, J.~A. 2008,
1677: \aap, 486, 255
1678:
1679: \bibitem[{{Aguilera} {et~al.}(2008){Aguilera}, {Pons}, \&
1680: {Miralles}}]{Aguilera:2008}
1681: {Aguilera}, D.~N., {Pons}, J.~A., \& {Miralles}, J.~A. 2008, \apjl, 673, L167
1682:
1683: \bibitem[{{Alford}(2001)}]{Alford:2001}
1684: {Alford}, M. 2001, Annual Review of Nuclear and Particle Science, 51, 131
1685:
1686: \bibitem[{{Alford} {et~al.}(1999){Alford}, {Rajagopal}, \&
1687: {Wilczek}}]{Alford:1999}
1688: {Alford}, M., {Rajagopal}, K., \& {Wilczek}, F. 1999, Nuclear Physics B, 537,
1689: 443
1690:
1691: \bibitem[{{Alford} {et~al.}(2008){Alford}, {Rajagopal}, {Schaefer}, \&
1692: {Schmitt}}]{Alford:2007a}
1693: {Alford}, M.~G., {Rajagopal}, K., {Schaefer}, T., \& {Schmitt}, A. 2008,
1694: Reviews of Modern Physics, 80, 1455
1695:
1696: \bibitem[{{Andersson} {et~al.}(2005){Andersson}, {Comer}, \&
1697: {Glampedakis}}]{Andersson:2005}
1698: {Andersson}, N., {Comer}, G.~L., \& {Glampedakis}, K. 2005, Nuclear Physics A,
1699: 763, 212
1700:
1701: \bibitem[{{Baldo} {et~al.}(2003){Baldo}, {Burgio}, \& {Schulze}}]{Baldo:2003}
1702: {Baldo}, M., {Burgio}, G.~F., \& {Schulze}, H.~. 2003, [ArXiv:astro-ph/0312446]
1703:
1704: \bibitem[{{Baym}(2007)}]{Baym:2006}
1705: {Baym}, G. 2007, AIP Conference Proceedings, 892, 8
1706:
1707: \bibitem[{{Berger} \& {Jaffe}(1987)}]{Berger:1987}
1708: {Berger}, M.~S. \& {Jaffe}, R.~L. 1987, \prc, 35, 213
1709:
1710: \bibitem[{{Blaschke} {et~al.}(2000){Blaschke}, {Kl{\"a}hn}, \&
1711: {Voskresensky}}]{Blaschke:2000}
1712: {Blaschke}, D., {Kl{\"a}hn}, T., \& {Voskresensky}, D.~N. 2000, \apj, 533, 406
1713:
1714: \bibitem[{{Drago} {et~al.}(2008){Drago}, {Pagliara}, \& {Parenti}}]{Drago:2008}
1715: {Drago}, A., {Pagliara}, G., \& {Parenti}, I. 2008, \apjl, 678, L117
1716:
1717: \bibitem[{{Endo} {et~al.}(2006){Endo}, {Maruyama}, {Chiba}, \&
1718: {Tatsumi}}]{Endo:2006}
1719: {Endo}, T., {Maruyama}, T., {Chiba}, S., \& {Tatsumi}, T. 2006, Progress of
1720: Theoretical Physics, 115, 337
1721:
1722: \bibitem[{{Fern{\'a}ndez} \& {Reisenegger}(2005)}]{Fernandez:2005}
1723: {Fern{\'a}ndez}, R. \& {Reisenegger}, A. 2005, \apj, 625, 291
1724:
1725: \bibitem[{{Freire}(2008)}]{Freire:2007a}
1726: {Freire}, P.~C.~C. 2008, AIP Conference Proceedings, 983, 459
1727:
1728: \bibitem[{{Freire} {et~al.}(2008{\natexlab{a}}){Freire}, {Ransom}, {B{\'e}gin},
1729: {Stairs}, {Hessels}, {Frey}, \& {Camilo}}]{Freire:2008a}
1730: {Freire}, P.~C.~C., {Ransom}, S.~M., {B{\'e}gin}, S., {et~al.}
1731: 2008{\natexlab{a}}, \apj, 675, 670
1732:
1733: \bibitem[{{Freire} {et~al.}(2008{\natexlab{b}}){Freire}, {Ransom}, {B{\'e}gin},
1734: {Stairs}, {Hessels}, {Frey}, \& {Camilo}}]{Freire:2008}
1735: {Freire}, P.~C.~C., {Ransom}, S.~M., {B{\'e}gin}, S., {et~al.}
1736: 2008{\natexlab{b}}, in American Institute of Physics Conference Series, Vol.
1737: 983, 40 Years of Pulsars: Millisecond Pulsars, Magnetars and
1738: More, ed. C. Bassa, Z. Wang, A. Cumming, \& V. M. Kaspi (Melville: AIP), 604
1739:
1740: \bibitem[{{Geppert}(2006)}]{Geppert:2006}
1741: {Geppert}, U. 2006, [ArXiv:astro-ph/0611708]
1742:
1743: \bibitem[{{Geppert} {et~al.}(2004){Geppert}, {K{\"u}ker}, \&
1744: {Page}}]{Geppert:2004}
1745: {Geppert}, U., {K{\"u}ker}, M., \& {Page}, D. 2004, \aap, 426, 267
1746:
1747: \bibitem[{{Glendenning}(1990)}]{Glendenning:1990}
1748: {Glendenning}, N.~K. 1990, Nuclear Physics A, 512, 737
1749:
1750: \bibitem[{{Glendenning}(1992)}]{Glendenning:1992}
1751: {Glendenning}, N.~K. 1992, \prd, 46, 1274
1752:
1753: \bibitem[{{Glendenning}(2000)}]{Glendenning:2000}
1754: {Glendenning}, N.~K. 2000, {Compact stars : nuclear physics, particle physics,
1755: and general relativity} (Berlin: Springer)
1756:
1757: \bibitem[{{Glendenning} \& {Weber}(2001)}]{Glendenning:2001a}
1758: {Glendenning}, N.~K. \& {Weber}, F. 2001, \apjl, 559, L119
1759:
1760: \bibitem[{{Gudmundsson} {et~al.}(1983){Gudmundsson}, {Pethick}, \&
1761: {Epstein}}]{Gudmundsson:1983}
1762: {Gudmundsson}, E.~H., {Pethick}, C.~J., \& {Epstein}, R.~I. 1983, \apj, 272,
1763: 286
1764:
1765: \bibitem[{{Hartle}(1967)}]{Hartle:1967}
1766: {Hartle}, J.~B. 1967, \apj, 150, 1005
1767:
1768: \bibitem[{{Hartle} \& {Thorne}(1968)}]{Hartle:1968}
1769: {Hartle}, J.~B. \& {Thorne}, K.~S. 1968, \apj, 153, 807
1770:
1771: \bibitem[{{Heiselberg} {et~al.}(1993){Heiselberg}, {Pethick}, \&
1772: {Staubo}}]{Heiselberg:1993}
1773: {Heiselberg}, H., {Pethick}, C.~J., \& {Staubo}, E.~F. 1993, Physical Review
1774: Letters, 70, 1355
1775:
1776: \bibitem[{{Jaikumar} {et~al.}(2002){Jaikumar}, {Prakash}, \&
1777: {Sch{\"a}fer}}]{Jaikumar:2002}
1778: {Jaikumar}, P., {Prakash}, M., \& {Sch{\"a}fer}, T. 2002, \prd, 66, 063003
1779:
1780: \bibitem[{{Jaikumar} \& {Prakash}(2001)}]{Jaikumar:2001}
1781: {Jaikumar}, P. \& {Prakash}, S. 2001, Physics Letters B, 516, 345
1782:
1783: \bibitem[{{Kaaret} {et~al.}(2007){Kaaret}, {Prieskorn}, {Zand}, {Brandt},
1784: {Lund}, {Mereghetti}, {G{\"o}tz}, {Kuulkers}, \& {Tomsick}}]{Kaaret:2007}
1785: {Kaaret}, P., {Prieskorn}, Z., {Zand}, J.~J.~M.~i., {et~al.} 2007, \apjl, 657,
1786: L97
1787:
1788: \bibitem[{{Kaminker} {et~al.}(2001){Kaminker}, {Haensel}, \&
1789: {Yakovlev}}]{Kaminker:2001}
1790: {Kaminker}, A.~D., {Haensel}, P., \& {Yakovlev}, D.~G. 2001, \aap, 373, L17
1791:
1792: \bibitem[{{Landau} \& {Lifshitz}(1980)}]{Landau:1980}
1793: {Landau}, L.~D. \& {Lifshitz}, E.~M. 1980, {Statistical physics. Pt.1, Pt.2},
1794: 3rd edn. (Oxford: Pergamon Press)
1795:
1796: \bibitem[{{Lattimer} \& {Prakash}(2007)}]{Lattimer:2007}
1797: {Lattimer}, J.~M. \& {Prakash}, M. 2007, \physrep, 442, 109
1798:
1799: \bibitem[{{Lattimer} {et~al.}(1991){Lattimer}, {Prakash}, {Pethick}, \&
1800: {Haensel}}]{Lattimer:1991}
1801: {Lattimer}, J.~M., {Prakash}, M., {Pethick}, C.~J., \& {Haensel}, P. 1991,
1802: Physical Review Letters, 66, 2701
1803:
1804: \bibitem[{{Leinson} \& {P{\'e}rez}(2006)}]{Leinson:2006}
1805: {Leinson}, L.~B. \& {P{\'e}rez}, A. 2006, Physics Letters B, 638, 114
1806:
1807: \bibitem[{{Lombardo} \& {Schulze}(2001)}]{Lombardo:2001}
1808: {Lombardo}, U. \& {Schulze}, H.-J. 2001, LNP, 578, 30
1809:
1810: \bibitem[{{Miao} \& {Xiao-Ping}(2007)}]{Miao:2007}
1811: {Miao}, K. \& {Xiao-Ping}, Z. 2007, \mnras, 375, 1503
1812:
1813: \bibitem[{{Miao} {et~al.}(2007){Miao}, {Xiao-Ping}, \& {Na-Na}}]{Miao:2007a}
1814: {Miao}, K., {Xiao-Ping}, Z., \& {Na-Na}, P. 2007, [ArXiv:astro-ph/0708.0900]
1815:
1816: \bibitem[{{Miralles} \& {van Riper}(1996)}]{Miralles:1996}
1817: {Miralles}, J.~A. \& {van Riper}, K.~A. 1996, \apjs, 105, 407
1818:
1819: \bibitem[{{Miralles} {et~al.}(1993){Miralles}, {van Riper}, \&
1820: {Lattimer}}]{Miralles:1993}
1821: {Miralles}, J.~A., {van Riper}, K.~A., \& {Lattimer}, J.~M. 1993, \apj, 407,
1822: 687
1823:
1824: \bibitem[{{M{\"u}hlschlegel}(1959)}]{Muhlschlegel:1959}
1825: {M{\"u}hlschlegel}, B. 1959, Zeitschrift f{\"u}r Physik, 155, 313
1826:
1827: \bibitem[{{Niebergal} {et~al.}(2007){Niebergal}, {Ouyed}, \&
1828: {Leahy}}]{Niebergal:2007}
1829: {Niebergal}, B., {Ouyed}, R., \& {Leahy}, D. 2007, \aap, 476, L5
1830:
1831: \bibitem[{{\"O}zel(2006)}]{Ozel:2006}
1832: {\"O}zel, F. 2006, Nature, 441, 1115
1833:
1834: \bibitem[{{Page} {et~al.}(2006){Page}, {Geppert}, \& {Weber}}]{Page:2006a}
1835: {Page}, D., {Geppert}, U., \& {Weber}, F. 2006, Nuclear Physics A, 777, 497
1836:
1837: \bibitem[{{Page} {et~al.}(2004){Page}, {Lattimer}, {Prakash}, \&
1838: {Steiner}}]{Page:2004}
1839: {Page}, D., {Lattimer}, J.~M., {Prakash}, M., \& {Steiner}, A.~W. 2004, \apjs,
1840: 155, 623
1841:
1842: \bibitem[{{Page} \& {Reddy}(2006)}]{Page:2006}
1843: {Page}, D. \& {Reddy}, S. 2006, Annual Review of Nuclear and Particle Science,
1844: 56, 327
1845:
1846: \bibitem[{{Pons} {et~al.}(2007){Pons}, {Link}, {Miralles}, \&
1847: {Geppert}}]{Pons:2007}
1848: {Pons}, J.~A., {Link}, B., {Miralles}, J.~A., \& {Geppert}, U. 2007, Physical
1849: Review Letters, 98, 071101
1850:
1851: \bibitem[{{Potekhin} {et~al.}(1997){Potekhin}, {Chabrier}, \&
1852: {Yakovlev}}]{Potekhin:1997}
1853: {Potekhin}, A.~Y., {Chabrier}, G., \& {Yakovlev}, D.~G. 1997, \aap, 323, 415
1854:
1855: \bibitem[{{Potekhin} \& {Yakovlev}(2001)}]{Potekhin:2001}
1856: {Potekhin}, A.~Y. \& {Yakovlev}, D.~G. 2001, \aap, 374, 213
1857:
1858: \bibitem[{{Reisenegger}(1995)}]{Reisenegger:1995}
1859: {Reisenegger}, A. 1995, \apj, 442, 749
1860:
1861: \bibitem[{{Schaab} {et~al.}(1999){Schaab}, {Sedrakian}, {Weber}, \&
1862: {Weigel}}]{Schaab:1999}
1863: {Schaab}, C., {Sedrakian}, A., {Weber}, F., \& {Weigel}, M.~K. 1999, \aap, 346,
1864: 465
1865:
1866: \bibitem[{{Schaab} {et~al.}(1996){Schaab}, {Weber}, {Weigel}, \&
1867: {Glendenning}}]{Schaab:1996}
1868: {Schaab}, C., {Weber}, F., {Weigel}, M.~K., \& {Glendenning}, N.~K. 1996,
1869: Nuclear Physics A, 605, 531
1870:
1871: \bibitem[{{Schaab} \& {Weigel}(1998)}]{Schaab:1998a}
1872: {Schaab}, C. \& {Weigel}, M.~K. 1998, \aap, 336, L13
1873:
1874: \bibitem[{{Schaffner-Bielich}(2007)}]{Schaffner-Bielich:2007}
1875: {Schaffner-Bielich}, J. 2007, [ArXiv:astro-ph/0703113]
1876:
1877: \bibitem[{{Schmitt} {et~al.}(2002){Schmitt}, {Wang}, \&
1878: {Rischke}}]{Schmitt:2002}
1879: {Schmitt}, A., {Wang}, Q., \& {Rischke}, D.~H. 2002, \prd, 66, 114010
1880:
1881: \bibitem[{{Schulze} {et~al.}(2006){Schulze}, {Polls}, {Ramos}, \&
1882: {Vida{\~n}a}}]{Schulze:2006}
1883: {Schulze}, H.-J., {Polls}, A., {Ramos}, A., \& {Vida{\~n}a}, I. 2006, \prc, 73,
1884: 058801
1885:
1886: \bibitem[{{Steiner} \& {Reddy}(2009)}]{Steiner:2008}
1887: {Steiner}, A.~W. \& {Reddy}, S. 2009, \prc, 79, 015802
1888:
1889: \bibitem[{{Steiner} {et~al.}(2002){Steiner}, {Reddy}, \&
1890: {Prakash}}]{Steiner:2002}
1891: {Steiner}, A.~W., {Reddy}, S., \& {Prakash}, M. 2002, \prd, 66, 094007
1892:
1893: \bibitem[{{Thorne}(1966)}]{Thorne:1966}
1894: {Thorne}, K.~S. 1966, in Proc. Int School of Physics
1895: Enrico Fermi XXXV, ed. L.~{Gratton}, (Academic: New York), 166
1896:
1897: \bibitem[{{Van Riper}(1991)}]{van-Riper:1991}
1898: {Van Riper}, K.~A. 1991, \apjs, 75, 449
1899:
1900: \bibitem[{{Voskresensky} {et~al.}(2003){Voskresensky}, {Yasuhira}, \&
1901: {Tatsumi}}]{Voskresensky:2003}
1902: {Voskresensky}, D.~N., {Yasuhira}, M., \& {Tatsumi}, T. 2003, Nuclear Physics
1903: A, 723, 291
1904:
1905: \bibitem[{{Weber}(1999)}]{Weber:1999}
1906: {Weber}, F. 1999, {Pulsars as astrophysical laboratories for nuclear and
1907: particle physics} (IOP Publishing)
1908:
1909: \bibitem[{{Weber}(2005)}]{Weber:2005}
1910: {Weber}, F. 2005, Progress in Particle and Nuclear Physics, 54, 193
1911:
1912: \bibitem[{{Weber} \& {Glendenning}(1992)}]{Weber:1992}
1913: {Weber}, F. \& {Glendenning}, N.~K. 1992, \apj, 390, 541
1914:
1915: \bibitem[{{Weber} {et~al.}(1991){Weber}, {Glendenning}, \&
1916: {Weigel}}]{Weber:1991}
1917: {Weber}, F., {Glendenning}, N.~K., \& {Weigel}, M.~K. 1991, \apj, 373, 579
1918:
1919: \bibitem[{{Xiaoping} {et~al.}(2008){Xiaoping}, {Li}, {Xia}, \&
1920: {Miao}}]{Xiaoping:2008}
1921: {Xiaoping}, Z., {Li}, Z., {Xia}, Z., \& {Miao}, K. 2008, ArXiv e-prints, 0808.1587
1922:
1923: \bibitem[{{Yakovlev} {et~al.}(2001){Yakovlev}, {Kaminker}, {Gnedin}, \&
1924: {Haensel}}]{Yakovlev:2001}
1925: {Yakovlev}, D.~G., {Kaminker}, A.~D., {Gnedin}, O.~Y., \& {Haensel}, P. 2001,
1926: \physrep, 354, 1
1927:
1928: \bibitem[{{Yakovlev} {et~al.}(1999){Yakovlev}, {Levenfish}, \&
1929: {Shibanov}}]{Yakovlev:1999}
1930: {Yakovlev}, D.~G., {Levenfish}, K.~P., \& {Shibanov}, Y.~A. 1999, Soviet
1931: Physics Uspekhi, 42, 737
1932:
1933: \bibitem[{{Yakovlev} \& {Pethick}(2004)}]{Yakovlev:2004a}
1934: {Yakovlev}, D.~G. \& {Pethick}, C.~J. 2004, \araa, 42, 169
1935:
1936: \bibitem[{{Zdunik} {et~al.}(2006){Zdunik}, {Bejger}, {Haensel}, \&
1937: {Gourgoulhon}}]{Zdunik:2006}
1938: {Zdunik}, J.~L., {Bejger}, M., {Haensel}, P., \& {Gourgoulhon}, E. 2006, \aap,
1939: 450, 747
1940:
1941: \end{thebibliography}
1942:
1943:
1944: \end{document}
1945: