1: %-------------------------DECLARATIONS-----------------------
2:
3: \documentstyle[floats,prd,aps,epsfig]{revtex}
4: \newcommand{\bce}{\begin{center}} \newcommand{\ece}{\end{center}}
5: \newcommand{\beq}{\begin{equation}} \newcommand{\eeq}{\end{equation}}
6: \newcommand{\beqy}{\begin{eqnarray}}
7: \newcommand{\eeqy}{\end{eqnarray}}
8:
9: \input epsf
10: \advance\voffset by 0.5in
11: \setlength{\oddsidemargin}{0in}
12: \setlength{\textwidth}{6.3in}
13:
14: \newcommand{\lsim}
15: {\mathrel{\raisebox{-.6ex}{$\stackrel{\textstyle<}{\sim}$}}}
16: \newcommand{\gsim}
17: {\mathrel{\raisebox{-.6ex}{$\stackrel{\textstyle>}{\sim}$}}}
18: \newcommand{\slsh}{\mbox{ \hspace{-1.1 em} $/$}}
19: \newcommand{\slst}{\mbox{ \hspace{-1.5 em} $/$}}
20: \newcommand{\ome}{\omega_B}
21:
22:
23: %------------------------END OF DECLARATIONS------------
24:
25: %------------------------TITLE AND ABSTRACT-------
26:
27: %------------------
28: \advance\voffset by -0.5in
29: \begin{document}
30: \title{Surface structure of quark stars with magnetic fields}
31: \author{Prashanth Jaikumar\footnote{jaikumar@phy.anl.gov}
32: %and Rachid Ouyed$^2$
33: }
34:
35: \address{Physics Division, Argonne National Laboratory,
36: Argonne IL 60439, USA.\\
37: %$^2$Department of Physics and Astronomy, University of Calgary,
38: % Calgary, AB T2N 1N4 Canada\\
39: }
40:
41: \maketitle
42: \begin{abstract}
43: We investigate the impact of magnetic fields on the electron
44: distribution in the electrosphere of quark stars. For moderately
45: strong magnetic fields $B\sim 10^{13}$G, quantization effects are
46: generally weak due to the large number density of electrons at
47: surface, but can nevertheless affect the spectral features of
48: quark stars. We outline the main observational characteristics of
49: quark stars as determined by their surface emission, and briefly
50: discuss their formation in explosive events termed Quark-Novae, which
51: may be connected to the $r$-process. \\
52: \bigskip
53:
54: \noindent PACS: 26.60.+c, 24.85.+p, 97.60.Jd
55:
56: \end{abstract}
57: %------------------------------END OF TITLE AND ABSTRACT------------
58:
59: %-----------------------------BEGIN INTRODUCTION-------------------
60:
61: \section{Introduction}
62: \label{sec_intro}
63: There is renewed interest in the theory and observation of strange
64: quark stars, which are believed to contain, or be entirely composed
65: of, deconfined quark matter~\cite{JSB}. An observational confirmation
66: of their existence would be conclusive evidence of quark deconfinement
67: at large baryon densities, an expected feature of Quantum
68: Chromodynamics (QCD). Furthermore, discovery of a stable bare quark star
69: affirms the Bodmer-Terazawa-Witten conjecture~\cite{BTW}, viz., that
70: at high enough density, strange quark matter, composed of up, down and
71: strange quarks, is absolutely stable with respect to nuclear
72: matter. This intriguing hypothesis is over three decades old, and bare
73: quark stars are but one possible realization put forward in the
74: intervening years. Broadly speaking, the field of quark star research
75: follows two intertwined paths: (1) efforts to describe the global
76: (mass and radius) and surface structure of such stars from
77: microscopics and (2) determination of observational features that
78: distinguish them from their close cousins, neutron stars. It is
79: possible that a few of the approximately 1500 neutron stars detected
80: thus far are really quark stars~\cite{Xu1}.
81:
82: \vskip 0.2cm
83:
84: A positive identification of a quark star would likely require a
85: complement of consistent signals. For example, physical properties
86: such as maximum mass, radius, minimum spin-periods and neutrino/photon
87: cooling rates (and hence temperature) all depend on the (unique)
88: equation of state for dense quark matter. As the latter is uncertain,
89: there is a range of allowed values for the above quantities that makes
90: a distinction based on current observational data difficult. We
91: emphasize that a bare strange star is self-bound~\cite{BTW,Chin79}
92: while a neutron star requires gravitational forces for its
93: binding. This implies a very different equation of state for the two
94: classes of stars; however, in the range of masses
95: ($1.35<M/M_{\odot}<2.4$) observed to date, the calculated radii are
96: similar ($R\sim 10$ to 15 km). Perturbative corrections to the quark
97: matter equation of state and color superconductivity effects further
98: complicate the issue~\cite{ARB}.~\footnote{We will not elaborate here
99: on the possibility of having superconducting phases of quarks. We
100: refer the interested reader to Refs.~\cite{ABR,IS,Ferrer,Blaschke,ALP}
101: for a consideration of these phases, and the consequences of their
102: existence inside compact stars in the present context.}
103:
104: \vskip 0.2cm
105:
106: It was first pointed out by Alcock et al.~\cite{AFO} that the bare
107: quark star surface would be characterized by an abrupt termination of
108: the quark density (density gradient $\sim 10^{26}$g/cm$^4$) and a less
109: rapid fall-off of a charge neutralizing skin of electrons termed the
110: electrosphere (density gradient $\sim 10^{18}$g/cm$^4$). The latter
111: extends up to 1000 Fermis above the bare quark star surface. The
112: electric field which binds these electrons to the surface is large
113: enough to pull electron-positron pairs out of the QED vacuum,
114: prompting copious pair-emission which is directed outward by the
115: radiation output by the star~\cite{AMU}. This mechanism of emission
116: from the bare surface of a strange quark star, due to both photons and
117: $e^+e^-$ pairs, can produce luminosities well above the Eddington
118: limit ($\sim 10^{38}~{\rm erg~sec}^{-1}$) up to a decade after the
119: quark star's birth~\cite{Page02}. The spectrum of emitted photons is
120: significantly harder than that from a normal cooling neutron star ($30
121: <E/{\rm keV} < 500$ instead of $0.1 < E/{\rm keV} <2.5$). After the
122: star has cooled to temperatures less than $T< 6\times 10^8$K, other
123: microscopic processes of photon emission can dominate the luminosity
124: of the star and yield a non-thermal spectrum~\cite{JPPG}. This
125: distinctive spectrum and temperature evolution, if observed, would
126: constitute an unmistakable signature of a strange quark star, but
127: would require a fortuitous detection before the star cools below the
128: observational threshold of present-generation space-based laboratories
129: such as INTEGRAL~\cite{Int}.
130:
131: \vskip 0.2cm
132:
133: Quark stars may be much more elusive if they have a suspended crust
134: of normal matter up to the neutron drip density, that is separated from
135: the quark surface by a microscopic gap but supported by the large
136: electric fields at surface~\cite{AFO,Madsen}. Alternatively, quark stars
137: may admit a mixed phase crust several meters in thickness and
138: extending up to its surface, if surface and Coulomb costs (determined
139: at the microscopic level by the current quark mass of the strange
140: quark) is small enough~\cite{JRS,ARRS}. This would obviate the need for
141: large electric fields at surface, smoothen the density gradient at
142: surface, and render the previously proffered signals
143: inapplicable. These uncertainties are not expected to be resolved
144: quickly since they involve quantifying the parameters of QCD to better
145: than 5\% in a distinctly non-perturbative regime. (NB. The density that
146: characterizes quark matter in these objects corresponds to a quark
147: chemical potential $\mu_q\sim 500$ MeV which does not greatly exceed
148: $\Lambda_{\rm QCD}\sim 200$ MeV.)
149:
150: \vskip 0.2cm
151:
152: However, we can certainly aim to quantitatively improve our current
153: understanding of the surface structure under the working assumption
154: that the star has a bare quark surface with an electrosphere, since
155: electromagnetic effects dominate over QCD effects. A clear and
156: necessary improvement is the introduction of surface magnetic fields.
157: Neutron stars are born with intense magnetic fields of order $10^{12}$
158: Gauss or higher, which may decay by over 5 orders of magnitude in the
159: star's lifetime. If a neutron star undergoes a phase conversion to a
160: bare quark star, the surface magnetic fields are expected to remain
161: intact in geometry, unless interior magnetic fields are expelled by a
162: superconducting state that displays the Meissner effect, in which case
163: rapid magnetic reconnections occur, diminishing the field at surface
164: to almost zero~\cite{OYD1}. In this work, we will superimpose a static
165: magnetic field $B$ on the electrosphere of a quark star, and calculate
166: the profile for the electrosphere as a first step toward
167: characterizing the spectrum of a magnetized quark star. We expect
168: considerable differences from the $B=0$ case since magnetized
169: atmospheres of neutron stars are already known to be very different
170: from non-magnetized ones~\cite{Pons}. Further, the inclusion of
171: magnetic fields is a step toward a more realistic treatment of the
172: quark star spectrum.
173:
174: \vskip 0.2cm
175:
176: In the next section, we reintroduce the established picture of the
177: electrosphere for the purpose of a self-contained presentation, and
178: mention the modifications due to the strong magnetic field. In section
179: III, we quantify the electrostatic potential and electron number
180: density in the magnetized electrosphere, and state qualitative impacts
181: on the electrospheric photon emission. We conclude in section IV with
182: a general discussion on the formation of Quark stars, with particular
183: attention to the attractive Quark-Nova scenario which may be connected to
184: $r$-process nucleosynthesis.
185:
186: %-----------------------END OF INTRODUCTION--------------------%
187: %-----------------------BEGIN SECTION II-----------------------%
188:
189: \section{Degenerate electron gas in a strong magnetic field}
190: \label{sec:Belectrosphere}
191: Recently, a few authors~\cite{UCH} pointed out that the deficit of
192: (massive) strange quarks due to surface effects on the star can lead
193: to a thin charged skin a few Fermis thick, which can lead to a jump in
194: the electric field at surface. The local charge neutrality condition
195: is $2/3n_u-1/3n_d-1/3n_s-n_e=0$, with the appropriate low temperature
196: ($T\ll \mu_i$) expressions in the bulk:
197:
198: \beq
199: n_i=\frac{p_{F,i}^3}{\pi^2}=\frac{(\mu_i^2-m_i^2)^{3/2}}{\pi^2};\quad m_e\ll m_{u,d}\ll m_s<\mu_q ~\forall q\quad .
200: \eeq
201:
202: Imposing the conditions of $\beta$-equilibrium and baryon number
203: conservation shows that $\mu_e\simeq m_s^2/4\mu$ to better than
204: 1\%. At surface, $s$-quarks are depleted due to surface tension
205: effects by an amount~\cite{mad}
206:
207: \beq
208: \delta n_s = -\frac{3}{4\pi}p_{F,s}^2\psi(\lambda_s); ~\psi(\lambda_s) = \left[\frac{1}{2}+\frac{\lambda_s}{\pi}-\frac{(1+\lambda_s^2)}{\pi}{\rm arctan(\lambda_s^{-1})}\right]; ~\lambda_s = \frac{m_s}{p_{F,s}}\,,
209: \eeq
210:
211: leading to a charged layer at the surface of a bare quark star
212: characterized by a surface charge density $\sigma = -(1/3)e\delta n_s$
213: and a discontinuous normal electric field with $\Delta E =
214: 4\pi\sigma$. For ``cold'' electrospheres, corresponding to the
215: temperature bound $T< 5\times 10^{11}$K, the electric field is
216: directed inward just underneath the surface and outward just above
217: it. The electrostatic potential and the electron number density are
218: substantially increased cf. the case when surface effects are
219: ignored (NB. The corresponding surface depletion of light quarks is
220: negligible.)~\cite{UCH}.
221:
222: \vskip 0.2cm
223:
224: The large electric field leads to the Schwinger instability of the QED
225: vacuum~\cite{Schwinger}, resulting in electron-positron pair-emission,
226: which can be an additional source of photons in the electrosphere,
227: which normally radiates photons via 2$\rightarrow$ 3 QED processes. A
228: model cooling calculation of bare strange stars with electrospheric
229: effects and the inclusion of color superconductivity (which can alter
230: the specific heat and thermal conductivity of quark matter) has also
231: been performed~\cite{Page02}. The conclusion is that bare strange
232: stars will display Super-Eddington photon luminosities at surface
233: temperatures $T>10^{9}$K, with a spectrum that is hard enough to
234: distinguish it from thermally radiating neutron stars. At lower
235: temperatures $6\times 10^8$K$<T<10^9$K, the bulk of the luminosity
236: comes from electron-positron pairs which subsequently imprint a wide
237: annihilation line on the non-thermal spectrum. Thermal emission is
238: much suppressed owing to plasma frequency effects of quark matter and
239: the electrosphere~\cite{CH}. Further cooling to temperatures $T<10^8$K
240: results in a non-thermal spectrum dominated by bremsstrahlung photons
241: from electron-electron collisions in the electrosphere~\cite{JPPG}.
242:
243: \vskip 0.2cm
244:
245: As mentioned in the introduction, a magnetic field is expected to
246: exist at surface, altering the electrosphere profile and its emission
247: properties. The electron motion perpendicular to the
248: magnetic field is quantized into Landau orbitals enumerated by the
249: quantum number $n_L=0,1,2..$ . For a relativistic fermion of charge
250: $q$ and mass $m$, such as the quarks and electrons in our case, the energy
251: in a $\hat{z}$-directed magnetic field $B$ is given by
252:
253:
254: \beq
255: \label{benergy}
256: \epsilon_{n} = \sqrt{m^2c^4+c^2p_z^2+2mc^2\hbar\omega_Bn}\equiv mc^2E;\quad 2n = 2(n_L+\frac{1}{2}+\sigma);\quad \omega_B=\frac{qB}{mc}\,,
257: \eeq
258:
259: where $\sigma=\pm 1/2$ is the spin-degeneracy of the fermion. The
260: $n=0$ state is non-degenerate while $n>0$ states are doubly
261: degenerate. We express the magnetic field strength in convenient units
262: '$b$' such that
263:
264: \beq
265: b\equiv \frac{\omega_B}{mc^2} = \frac{B}{4.414\times 10^{13}~{\rm G}}\quad{\rm for ~electrons}
266: \eeq
267:
268:
269: and take $B=4.414\times 10^{13}$G (corresponding to $b=1$) as an upper
270: limit on the surface magnetic field of quark stars. Although higher
271: fields may exist in magnetars and AXPs (anomalous X-ray pulsars),
272: their interior magnetic fields would likely be large enough to inhibit
273: the formation of quark matter~\cite{SC1}. For a given magnetic field,
274: both temperature and matter density determine the filling fraction of
275: the Landau levels, and hence the degree of quantization. From
276: Eqn.(\ref{benergy}), the highest Landau level occupied by a fermion
277: with energy $\epsilon_{n}$ is
278:
279: \beq n_{L,{\rm max}}: n_{{\rm max}}={\rm Int}(\nu);\quad \nu
280: =\frac{E^2-1}{2b}\quad . \eeq
281:
282: Since the quarks are highly relativistic, $E\gg 1$ and many Landau
283: levels are occupied for the light quarks, and even more for the
284: strange quark, so that quantization effects may be ignored. Further,
285: when the temperature $T\gg T_B$ where $T_B=\hbar\omega_B/k_B$ ($k_B$
286: is the Boltzmann constant), thermal effects smear out adjacent Landau
287: levels, and the magnetic field is effectively non-quantizing,
288: irrespective of the degeneracy. This effect is more important for
289: quarks, which have a larger mass, and hence smaller cyclotron
290: frequency, than electrons. We thus expect no modifications to the
291: quark distribution due to magnetic fields of typical magnitudes
292: considered in this work.
293:
294: \vskip 0.2cm
295:
296: For electrons, quantization effects become important as we move
297: outward from the electrosphere, and more electrons reside in the lower
298: Landau levels. The number density of electrons in a magnetic field is
299: given by
300:
301: \beq
302: \label{negen}
303: n_e=\frac{m_e\ome}{h^2}\int_{-\infty}^{\infty}dp_z\sum_{n,\sigma}f(\epsilon_n);\quad f(\epsilon_n)=\frac{1}{1+{\rm e}^{(\epsilon_n-\mu_e)/T}}\quad .
304: \eeq
305:
306: For a cold degenerate electron gas which satisfies $T\ll\mu_e$, we may
307: approximate
308:
309: \beq
310: \label{ne}
311: n_e(\mu_{e0})=\Theta(\mu_{e0}-1)\frac{2m_e^2\ome c}{h^2}\left[\sqrt{\mu_{e0}^2-1}+2\sum_{n=1}^{n_{\rm max}}\sqrt{\mu_{e0}^2-1-2bn}\right]; \quad \mu_{e0}=\frac{\mu_e(T=0)}{m_ec^2}\quad .
312: \eeq
313:
314: From Eqn.(\ref{ne}), the density at which the second Landau level
315: becomes populated is
316:
317: \beq
318: n_e^{\ast}=\frac{(m_e\ome)^{3/2}}{\hbar^2\pi^2\sqrt{2}}=1.24\times 10^{30}b^{3/2} ~{\rm cm}^{-3}\quad .
319: \eeq
320:
321: This is already less than the typical electron density at surface in
322: the $B=0$ case, where $n_e\sim 10^{-4}-10^{-3}/{\rm
323: fm}^{3}$~\cite{Usov}. Therefore, several Landau levels will be
324: occupied, and the $B$-field is non-quantizing. In fact, for $\nu\gg 1$
325: and $T\ll\mu_e$, it follows from Eqn.(\ref{ne}) that~\cite{Shapiro}
326:
327: \beq
328: n_e=\frac{2m_e^2\ome c}{h^2}\sqrt{\mu_{e0}^2-1}\left[\frac{2\sqrt{\mu_{e0}^2-1}}{3b}-3\right]\approx \frac{(\mu_e^2-m_e^2)^{3/2}}{3\pi^2}\quad ,
329: \eeq
330:
331: which is the usual expression relating the number density and chemical
332: potential for a relativistic degenerate Fermi gas in the $B=0$
333: case. For the range of temperatures and densities considered in this
334: work, $T\ll T_B=\hbar\omega_B/k_B$ holds, so that temperature effects
335: on electron occupation numbers is negligible. The situation is summarized
336: in Fig. 1, which shows how the number density of electrons, as
337: obtained from Eqn.(\ref{negen}) behaves as a function of electron
338: chemical potential, temperature and magnetic field. The truncation of
339: the curves at $\nu=0$ for finite $T$ represent the fact that for lower
340: (negative) chemical potentials, and hence densities, all electrons
341: reside in the $n=0$ state. We note that only for temperatures $T\geq
342: T_B$ does the number density deviate significantly from the $T=0$
343: curve (NB. This is due to the smearing of the Landau levels by
344: temperature fluctuations.). We therefore continue to use the zero
345: temperature approximation to the electron density in the rest of the
346: paper.
347: \vskip 0.5cm
348:
349:
350: \begin{figure}[!ht]
351: \bce
352: \epsfig{file=ne.eps,width=13.0cm,height=7.0cm}
353: \ece
354: %
355: \caption{The number density of electrons (scaled to $n_e^{\ast}$) from Eqn.~(\ref{negen}) as a function of the filling factor $\nu$ for various temperatures. The solid curves are for non-zero magnetic field, and display a loss of the Landau level structure with increasing temperature.}
356: %
357: \label{figps1}
358: \end{figure}
359:
360:
361:
362:
363: %-----------------------END OF SECTION II---------------------------%occurs in the
364:
365: %-----------------------BEGIN SECTION III---------------------------%
366: \section{Electrosphere profile with a magnetic field}
367: \label{sec:brem}
368: We would now like to obtain the electron density profile as a function
369: of distance in the electrosphere. This will be modified from previous
370: results~\cite{AFO,CH} which were in the absence of a magnetic
371: field. The baryon density at the surface of a strange star, where the
372: quark pressure vanishes, is expected to be in the range $(2-5)\rho_0$,
373: corresponding to quark chemical potentials in the range 300-400
374: MeV. At such densities, the strange quark current mass is not
375: ignorable, with $m_s\sim150$ MeV. Local charge neutrality then implies
376: an electron chemical potential at surface $\mu_e\approx 15-20$
377: MeV. This picture of a homogeneous surface as opposed to a crusty one
378: with quark nuggets is justified by the large value of the surface
379: tension given our parameter values~\cite{Berger}. The surface
380: depletion of strange quarks at the surface results in a positively
381: charged layer, which is two orders of magnitude thinner than the
382: electrosphere. For cold electrospheres, this charged layer results in
383: oppositely directed electric fields at the star's surface as well as
384: an enhanced electron number density $n_e\gg n_{\ast}$.
385:
386: \vskip 0.2cm
387:
388: For a general magnetic field with poloidal and toroidal components at
389: the surface, the electron distribution can, in principle, be obtained
390: using the formalism of magnetic field density functional theory or by
391: Green function techniques, both of which are complicated tasks. As a
392: first study, we adopt a simple picture of a constant magnetic field
393: directed normally to the surface and ignore Coulomb interactions and
394: exchange effects, which are justified approximations given the extreme
395: degeneracy being explored in the problem. We employ the
396: self-consistent Thomas-Fermi approximation and present a numerical
397: solution of the Poisson equation
398:
399: \beq
400: \label{eqntfermi}
401: \nabla^2\phi(\vec{r},\vec{B})=-\rho(\vec{r},\vec{B})=en_e(\vec{r},\vec{B})\,.
402: \eeq
403:
404: In the plane parallel approximation $\vec{B}=B\hat{z},~\hat{r}=\hat{z}$,
405: and equilibrium requires $\mu_e(n_e(z))-e\phi(z)=0$. For arbitrary magnetic
406: fields, the equilibrium constraint involves magnetic forces $F_{\rm mag}=
407: \vec{j}\times\vec{B}\neq 0$ where $\vec{j}$ is the electron current density.
408: With our approximations, the problem is an electrostatic one, and
409: the electrostatic potential outside ($z>0$) the star solves
410:
411: \beqy
412: \label{eqntfermi2}
413: \nabla^2\eta_+&=&\frac{2\alpha
414: bm_e^2}{\pi}\Theta(\eta_+-1)\left[\sqrt{\eta_+^2-1}+2\sum_{n=1}^{N}\sqrt{\eta_+^2-1-2bn}\right];\quad
415: N={\rm Int}\left(\frac{\eta_+^2-1}{2b}\right)\quad ,\\
416: \eta_+&=&\frac{e\phi_+}{m_ec^2}; \quad\phi_+=\phi(z>0) \quad .
417: \eeqy
418:
419: In the limit $N\gg 1$, when the electrons populate several levels, we recover
420: the expression for the non-magnetic case
421:
422: \beq
423: \nabla^2\phi_+=\frac{4\alpha}{3\pi}e^2\phi_+^3\quad ,
424: \eeq
425:
426: while for $N=0$, we obtain
427:
428: \beq
429: \nabla^2\phi_+=\frac{em\omega_Bp_F}{2\pi^2}\quad ,
430: \eeq
431:
432: the expression which has been derived previously for the
433: non-relativistic case~\cite{Fushiki} but applies equally well to the
434: relativistic case as long as only the 0$^{\rm th}$ level is
435: populated. It is interesting to note that the above equation leads to
436: an exponentially decaying electron profile at large distances from the
437: surface, contrary to the $1/z$ fall-off in the non-magnetic case. As
438: $\eta_+\rightarrow 1$, we expect interactions and exchange effects to be
439: important, but since this is only a minute region of the
440: electrosphere, we do not correct for it.
441:
442: Eq.~(\ref{eqntfermi2}) can be solved numerically once the boundary
443: conditions are determined. Clearly, $\eta_+\rightarrow 1$ as
444: $z\rightarrow\infty$ (this corresponds to vanishing electron number
445: density and is numerically implemented by the shooting method to
446: obtain the correct asymptote at sufficiently large $z$. This method is
447: also practically applicable to the case of suspended crusts above a
448: quark star; see Ref.~\cite{Madsen}.). The second boundary condition is
449: determined at the surface $z=0$, where
450: $\phi=\phi_0\Rightarrow\eta=\eta_0$ and $\eta_0$ must be
451: determined by the discontinuity in the electric field
452: $\nabla\phi_+-\nabla\phi_-=4\pi\sigma$ with $\phi_-=\phi(z<0)$. The
453: electrostatic potential $\phi_-$ inside the star is almost identical
454: to the non-magnetic case since the electron density deep inside the
455: star, given by $\tilde{\mu_e}=m_s^2/4\mu_q\sim 10~{\rm MeV}\gg
456: m_ec^2=0.5~{\rm MeV}$ so that for $b\leq 10.0$ fields, quantization
457: effects are negligible. In this case, $\phi_-\rightarrow \phi_q$ as
458: $z\rightarrow -\infty$ where the relation between $\phi_0$ and $\phi_q$
459: is given by
460:
461: \beq
462: 4\pi\sigma=e\left(\frac{2\alpha}{3\pi}\right)\left[\phi_0^2+(\phi_q-\phi_0)\left\{(\phi_0+\phi_q)^2+2\phi_q^2\right\}\right]^{1/2}\quad ,
463: \eeq
464:
465: wherein the prefactor corrects an error in equation (30) of
466: Ref.~\cite{UCH}. For our choice of $\tilde{\mu_e}=11.25$ MeV
467: (corresponding to $m_s=150$ MeV and $\mu_q=500$ MeV),
468: $\phi_q=37.4$ MeV, $\phi_0=578$ MeV and hence $\eta_0=341$ MeV. The
469: boundary conditions having been specified, the results for the
470: electrostatic potential $\phi_+,\phi_-$ as obtained from
471: Eq.(\ref{eqntfermi}) are illustrated in Fig.~\ref{figps2}, and
472: compared to the corresponding expressions for the non-magnetized
473: electrosphere with the same boundary conditions.
474:
475: \vskip 0.5cm
476:
477:
478: \begin{figure}[!ht]
479: \bce
480: \leavevmode
481: \epsfig{file=phi.eps,width=7.5cm,height=6.0cm}
482: \epsfig{file=tfermi.eps,width=7.5cm,height=6.0cm}
483: \ece
484: %
485: \caption{Left panel: The electrostatic energy (scaled to $m_ec^2$) at
486: the edge of the star, from Eqn.~(\ref{eqntfermi}) for the unmagnetized
487: ($b=0$) and magnetized ($b=1.0$) cases. Right panel: The corresponding
488: electron number density (scaled to $n_e^{\ast}$) at the surface.}
489: %
490: \label{figps2}
491: \end{figure}
492:
493:
494: We observe that while the electrostatic potential inside the star is
495: unaffected by magnetic field strengths of $b\sim{\cal O}(1)$, that
496: outside the star is increased in comparison to the unmagnetized
497: case. A similar finding for the number density reflects the fact that
498: the electrons become localized in configuration space, their wave
499: functions compressed into the Larmor radius, so that the Pauli
500: repulsion is decreased and electrons can be squeezed together in
501: greater numbers than in the unmagnetized case. This effect is most
502: pronounced when the quantization effects are large, as can be seen by
503: the wiggles that demarcate the filling of the Landau levels (right
504: panel of Fig.~\ref{figps2}). In both figures above, the curves meet at
505: the surface ($z=0$) where they satisfy the same boundary condition. At
506: distances far from the surface, the curves cross each other since the
507: total number density is the same in both cases. The number density
508: falls off exponentially for the magnetized case while it falls off as
509: $1/z$ for the unmagnetized case.
510:
511: \vskip 0.2cm
512:
513: In the absence of a thin ($10^{-5}$M$_{\odot}$) crustal layer of
514: accreted matter above the strange star, the spectral properties of a
515: strange star are linked to radiative transport in the
516: electrosphere. This is because radiation from the stellar interior is
517: cutoff by the high plasma frequency ($\omega_p\sim$ 20 MeV) of dense
518: quark matter, and Landau-Pomeranchuk-Migdal (LPM) suppression of
519: bremsstrahlung photons from quark-quark collisions~\cite{CH}. For an
520: unmagnetized electrosphere, the emission at high luminosities is
521: dominated by $e^+e^-$ pair emission so long as the temperature is high
522: enough. At lower luminosities, the emission is dominated by photon
523: bremsstrahlung from electron-electron collisions in the
524: electrosphere. The spectrum is non-thermal due to the finite plasma
525: frequency of the degenerate electron gas and large mean free path for
526: photons in the electrosphere.
527:
528: \vskip 0.2cm
529:
530: The effect of the intense magnetic field on electrospheric emission is
531: currently being investigated and will be reported elsewhere; here, we
532: only state some qualitative effects. The plasma frequency plays a
533: vital role in the emission properties of the electrosphere, since
534: lower frequency photons are rapidly damped by the complexion of the
535: refractive index
536:
537: \beq n_r^2=1-\frac{\omega_p^2}{\omega^2};\quad
538: \omega_p^2\simeq\frac{e^2}{3\pi^2}\mu_e^2\quad .
539: \eeq
540:
541: This collective effect of the plasma is important at high electron
542: density (inner regions of the electrosphere) where the Debye length is
543: smaller than the mean free path. Note that despite the large electron
544: density, the large degeneracy limits the available scattering states
545: of electrons, leading to a large mean free path for photons. In the
546: presence of a magnetic field, when quantization effects become
547: important, collective effects are determined by the Larmor radius
548: rather than the Debye length, and the reflectivity for normal
549: incidence becomes~\cite{Van}
550:
551: \beq
552: R=\frac{1}{2}\left[\left|\frac{n_{r,1}-1}{n_{r,1}+1}\right|^2+\left|\frac{n_{r,2}-1}{n_{r,2}+1}\right|^2\right];\quad
553: n_{r\{1,2\}}^2=1-{\frac{\omega_p^2}{\omega^2}}\frac{\omega}{(\omega\pm\omega_B)}\quad.
554: \eeq
555:
556: When $\omega_p\ll\omega\ll\omega_B$, the reflectivity is considerably
557: diminished from the unmagnetized case, and the electron plasma will
558: radiate more strongly. Further, the dielectric tensor will modify the
559: Poisson equation, resulting in the breakdown of the plane parallel
560: approximation employed here. Thus, our estimate of the profile is
561: simplistic and must be made self-consistent. The intense magnetic
562: field will also lead to polarization effects. Taking into account the
563: thermal conductivity of the electrons in a strong magnetic field, a
564: non-uniform surface temperature distribution is expected. The
565: inclusion of toroidal components of the magnetic field can smoothen
566: out some of these non-uniformities, and has been linked to the largely
567: uniform blackbody spectrum of some isolated neutron
568: stars~\cite{GKP1}. It would be interesting to see if the
569: electrospheric surface of the strange star displays similar spectral
570: features. In particular, the possibility of having bound positronium
571: in the magnetized electrosphere should be investigated.
572:
573: %-----------------------END OF SECTION III-----------------------------%
574: %-----------------------BEGIN SECTION IV-------------------------------%
575:
576: \section{Strange stars and Astrophysics: The Quark-Nova and the $r$-process}
577: \label{sec:quarks}
578: Strange quark stars have been linked to some of the most energetic and
579: violent phenomena in astrophysics, such as Gamma-ray bursts
580: (GRBs)~\cite{PCH}, type I X-ray superbursts~\cite{Cumming} and
581: Soft-gamma repeaters (SGRs)~\cite{Usov2} in what is being perceived as
582: a new paradigm in astrophysics - the connection of high-energy
583: astrophysics with exotic forms of dense matter. The energy released in
584: the phase conversion of a hadronic star to a quark star is typically
585: about $10^{53}$ ergs, similar in magnitude to the energy of observed
586: gamma-ray bursts, and accretion and flash heating on a strange star
587: surface is a viable model for SGR flares and superbursts. Further, the
588: emission spectra of the long bursts can also be explained by invoking
589: the presence of strange quark matter and its subsequent transition to
590: a color-superconducting state~\cite{ORV}. All of these scenarios are
591: predicated upon the existence of absolutely stable three-flavor quark
592: matter at high density.
593:
594: \vskip 0.2cm
595:
596: Here, we discuss the formation of a quark star in a Quark-Nova, a
597: scenario in which nuclear matter at high density in the core of a
598: neutron star undergoes a phase transition to two flavor (up and down)
599: quark matter, which then undergoes equilibrating weak reactions to
600: form three flavor quark matter. The advantage is that this does not
601: require the improbable simultaneous and spontaneous conversion to
602: strange matter (via weak interactions) of several neutrons within a
603: small volume. In addition, the Coulomb barrier-free absorption of
604: neutrons can enlarge the quark phase, so that the whole neutron star
605: is essentially converted to strange quark matter, thus forming a quark
606: star. The dynamics of this process is only beginning to be
607: explored~\cite{Drago}, and it is likely that some mass ejection from
608: the surface of a neutron star can take place, aided by
609: neutrinos~\cite{JRK} but powered mostly by pre-compression shock waves
610: from the rapidly advancing quark conversion front~\cite{JOOM}.
611:
612: \vskip 0.2cm
613:
614: There are various ways in which this conversion can be triggered. A
615: rapidly rotating neutron star is slowed down principally by magnetic
616: braking (and additionally by energy loss through gravitational waves),
617: thereby reducing the centrifugal forces and increasing central
618: pressures. The timescales and dynamics of this spin-down have been
619: discussed in~\cite{Staff}. The probability of triggering a first order
620: transition by forming a tiny lump of two flavor quark matter (quantum
621: nucleation) is exponentially sensitive to the central
622: pressure~\cite{Bombaci}. This makes the transition much more likely as
623: the star spins down. Material accumulated from a fallback disk around
624: a cooling neutron star or accretion from a companion star, for
625: example, in a low-mass X-ray binary system can also force the
626: conversion. Even the Bondi accretion rate of $10^{21}$ g/sec from the
627: interstellar medium can lead to a mass increase of up to 0.1$M_{\odot}$
628: within a million years which can be sufficient to compress the star
629: beyond the minimum density for the phase transition. Other
630: possibilities include the clustering of $\Lambda$-baryons at high
631: density and seeding by energetic cosmic neutrinos or
632: strangelets~\cite{AFO}. Deconfinement of quarks can also occur as
633: early as the protoneutron star stage following a core-collapse
634: supernova explosion if the central density is large enough. However,
635: in this case, neutrino trapping in the hot and dense interior can push
636: the transition density higher, delaying the collapse to the point
637: where a black hole might be formed instead~\cite{Prakash}.
638:
639: \vskip 0.2cm
640:
641: The amount of mass ejected and the dynamics of the Quark-nova make it
642: a potential candidate for $r$-process nucleosynthesis. The
643: decompressing neutron matter at the surface of the neutron star which
644: is undergoing the phase conversion to a quark star will expand rapidly
645: and chunks of neutron matter will fall apart, allowing beta-decays to
646: heat the material to close to $r$-process like temperatures of $T\sim
647: 0.1$ MeV. Following the expansion of this dense ejecta (including
648: $\beta$-decays) down to neutron drip density using network
649: calculations of Meyer~\cite{Meyer}, we find that the matter ends up
650: distributed over a wide range of elements from $Z\sim 40-70$. At this
651: stage, the full $r$-process reaction network can be coupled to obtain
652: the final abundance distribution, as well as the evolution of
653: temperature, entropy and electron fraction. The most important
654: parameter that governs the evolution of the above physical quantities
655: is the expansion timescale of the ejecta, which is the time taken to
656: evolve from a density of $\rho\simeq 10^{14}$g/cc to $\rho\simeq
657: 10^{11}$g/cc. Assuming that the ejecta is on an escape trajectory and
658: expands as a multiple of the free-fall timescale in a spherically
659: symmetric, non-interacting manner, we obtain a timescale on the order
660: of 0.01-1 milliseconds. Fig.~\ref{figps3} shows the $r$-abundance of
661: the elements (labeled ``network'') from a Quark-Nova compared to the
662: normalized solar $r$-abundance pattern. The overabundance for elements
663: with mass number $A>230$ is due to the omission of alpha-decays in the
664: network. It is also notable that not all neutrons are consumed in this
665: scenario, and that several light elements are produced with
666: significant abundance. It is likely that Quark-Novae may be an
667: additional contributor to the $r$-process material in the universe,
668: presenting another link between dense quark matter and astrophysics. A
669: detailed analysis of $r$-process nucleosynthesis from Quark-Novae,
670: including contrasts to standard or primary astrophysical sites such as
671: Type II supernovae and neutron star mergers, and comparisons to trends
672: of heavy-element abundance ratios in metal-poor stars will be
673: presented in a forthcoming article~\cite{JOOM}.
674:
675: \vskip 0.2cm
676: \begin{figure}[ht!]
677: \centerline{\includegraphics[width=0.7\textwidth,angle=0]{rp.eps}}
678: \caption{Abundance pattern of $r$-process elements from a Quark-Nova: The
679: initial electron fraction $Y_e=0.03$, entropy per nucleon $S/k_B=1$ ,
680: temperature $T_9=1.0$ and expansion timescale $\tau=1.0$
681: millisecond. The scaled comparison to solar abundance is also shown.}
682: \label{figps3}
683: \end{figure}
684: \vskip 0.2cm
685:
686: While several observable characteristics of strange stars have been
687: conjectured in the literature~\cite{Glenny,Weber}, such as period-clustering,
688: anomalies in braking index, global properties (mass and radius), and
689: small spin-periods, spectral differences between quark stars and
690: neutron stars have only been touched upon. If strange stars are made
691: in Quark-Novae, and are linked to $r$-processing of the heavy
692: elements, one can expect a most direct evidence of a quark star to
693: come from $\gamma$-decays of unstable nuclei produced in the
694: $r$-process. We expect the Quark-Nova ejecta to achieve $\gamma$-ray
695: transparency sooner than supernova ejecta since Quark star progenitors
696: (i.e., neutron stars) lack extended atmospheres, so that $r$-process
697: only gamma emitters with lifetimes of the order of years (such as
698: $^{137}$Cs, $^{144}$Ce, $^{155}$Eu and $^{194}$Os) can be used as tags
699: for the Quark-Nova.
700:
701: \vskip 0.2cm
702:
703: In conclusion, this is an exciting time for quark star
704: research. Recent developments in the theory of high density quark
705: matter have provided novel ways of explaining the most energetic
706: astrophysical events, and spurred re-examinations of old paradigms of
707: strange stars. With increasing data now coming from myriad sources
708: (egs. neutron stars that push theoretical limits on spin periods
709: and radius; GRB spectra that link to quark stars as inner engines;
710: abundance trends of heavy elements in ultra metal-poor stars that
711: point to multiple sources of the $r$-process), it appears that quarks
712: may play an important role in uncovering the secrets of many
713: astrophysical puzzles.
714:
715: %-------------------------END TEXT----------------------------%
716: \section*{Acknowledgments}
717: %
718: I thank the organizers of the IXth workshop on high-energy physics and
719: phenomenology (WHEPP-9) at the Institute of Physics, Bhubaneswar,
720: India, for hosting an exciting meeting in a most hospitable setting. I
721: acknowledge useful interactions with Kaya Mori, Rachid Ouyed, Kaori
722: Otsuki, Brad Meyer, Mandar Bhagwat and Stewart Wright, and the
723: hospitality of the Canadian Institute for Theoretical Astrophysics
724: (CITA), where discussions pertinent to this work were held. P.J. is
725: supported by the Department of Energy, Office of Nuclear Physics,
726: contract no. W-31-109-ENG-38. \rm
727: %-----------------------THE REFERENCES------------------------------------
728: \begin{flushleft}
729:
730: \begin{thebibliography}{99}
731:
732: \bibitem{JSB}
733: J. Schaffner-Bielich, J. Phys. {\bf G31}, S561 (2005).
734:
735: \bibitem{BTW} A. R. Bodmer, Phys. Rev. {\bf D4}, 1601 (1971);
736: H. Terazawa, {\it INS-Report-338}, University of Tokyo (1979);\\
737: E. Witten, Phys. Rev. {\bf D30}, 272 (1984).
738:
739: \bibitem{Xu1} R. Xu, in Proceedings of The 2005 Lake Hanas
740: International Pulsar Symposium, {\it astro-ph/0512519}.
741:
742: \bibitem{Chin79}
743: S. Chin and A. K. Kerman, Phys. Rev. Lett {\bf 43}, 1292 (1979).
744:
745: \bibitem{ARB} M. Alford, M. Braby, M. Paris and S. Reddy,
746: Astrophys. J. {\bf 629}, 969 (2005).
747:
748: \bibitem{ABR} M. Alford, J. A. Bowers and K. Rajagopal,
749: J. Phys. {\bf G27}, 541 (2001).
750:
751: \bibitem{IS}
752: I. A. Shovkovy, M. Hanauske and M. Huang, {\it hep-ph/0310286}.
753:
754: \bibitem{Ferrer}
755: E. J. Ferrer, V. de la Incera and C. Manuel, {\it hep-ph/0601179}.
756:
757: \bibitem{Blaschke}
758: D. M. Sedrakian and D. Blaschke, Astrofiz. {\bf 45}, 203 (2002).
759:
760: \bibitem{ALP} A. Drago, A. Lavagno and G. Pagliara, Phys. Rev. {\bf
761: D69}, 057505 (2005).
762:
763: \bibitem{AFO}
764: C. Alcock, E. Farhi and A. V. Olinto, Astrophys. J. {\bf 310},
765: 261 (1986).
766:
767: \bibitem{AMU} A. G. Aksenov, M. Milgrom and V. V. Usov,
768: Astrophys. J. {\bf 609}, 363 (2004).
769:
770: \bibitem{Page02}
771: D. Page and V. V. Usov, Phys. Rev. Lett. {\bf 89}, 131101 (2002).
772:
773: \bibitem{JPPG}
774: P. Jaikumar, C. Gale, D. Page and M. Prakash, Phys. Rev.
775: {\bf D70}, 023004 (2004).
776:
777: \bibitem{Int}
778: J. Knodlseder, astro-ph/0207527.
779:
780: \bibitem{Madsen}
781: M. Stejner and J. Madsen, Phys. Rev. {\bf D72}, 123005 (2005).
782:
783: \bibitem{JRS} P. Jaikumar, S. Reddy and A. Steiner,
784: Phys. Rev. Lett. {\bf 96}, 041101 (2006).
785:
786: \bibitem{ARRS}
787: M. Alford, K. Rajagopal, S. Reddy and A. Steiner, {\it hep-ph/0604134}.
788:
789: \bibitem{OYD1}
790: R. Ouyed, B. Niebergal, W. Dobbler and D. Leahy, {\it astro-ph/0510691}.
791:
792: \bibitem{Pons} J. F. P\'{e}rez-Azor\'{i}n, J. A. Miralles and
793: J. A. Pons, {\it astro-ph/0510684}.
794:
795: \bibitem{UCH} V. V. Usov, K. S. Cheng and T. Harko, Astrophys. J. {\bf
796: 620}, 915 (2005).
797:
798: \bibitem{mad}
799: J. Madsen, Phys. Rev. Lett. {\bf 87}, 172003 (2001).
800:
801: \bibitem{Schwinger}
802: J. Schwinger, Phys. Rev. {\bf 82}, 664 (1951).
803:
804: \bibitem{CH}
805: K. S. Cheng and T. S. Harko, Astrophys. J. {\bf 622}, 1033 (2005).
806:
807: \bibitem{SC1}
808: S. Ghosh and S. Chakrabarty, Pramana {\bf 60}, 901 (2002).
809:
810: \bibitem{Usov}
811: V. V. Usov, Astrophys.J. {\bf 550}, L179 (2001).
812:
813: \bibitem{Shapiro} A. M. Abrahams and S. L. Shapiro, Astrophys. J. {\bf
814: 374}, 652 (1991).
815:
816: \bibitem{Berger}
817: M. S. Berger and R. L. Jaffe, Phys. Rev. {\bf C35}, 213 (1987).
818:
819: \bibitem{Fushiki} I. Fushiki, E. H. Gudmundsson and C. J. Pethick,
820: Astrophys. J. {\bf 342}, 958 (1989).
821:
822: \bibitem{Van} M. van Adelsberg, D. Lai, A. Y. Potekhin and P. Arras,
823: Astrophys. J. {\bf 628}, 902 (2005).
824:
825: \bibitem{GKP1} U. Geppert, M. K\"{u}ker and D. Page, Astron. \&
826: Astrophys. {\bf 426}, 267G (2004); see also {\it astro-ph/0512530}.
827:
828: \bibitem{PCH}
829: B. Paczynski and P. Haensel, MNRAS {\bf 362}, L4 (2005).
830:
831: \bibitem{Cumming}
832: D. Page and A. Cumming, Astrophys.J. {\bf 635}, L157 (2005).
833:
834: \bibitem{Usov2}
835: V. V. Usov, Phys. Rev. Lett. {\bf 87}, 021101 (2001).
836:
837: \bibitem{ORV}
838: R. Ouyed, R. Rapp and C.Vogt, Astrophys. J. {\bf 632}, 1001 (2005).
839:
840: \bibitem{Drago}
841: A. Drago, A. Lavagno and I. Parenti; astro-ph/0512652
842:
843: \bibitem{JRK}
844: P. Jaikumar, R. Ouyed and P. Ker\"anen, Astrophys. J. {\bf 618}, 485 (2004).
845:
846: \bibitem{JOOM}
847: P. Jaikumar, R. Ouyed, K. Otsuki and B. S. Meyer, {\it in preparation}.
848:
849: \bibitem{Staff}
850: J. E. Staff, R. Ouyed and P. Jaikumar, {\it astro-ph/0603743}.
851:
852: \bibitem{Bombaci}
853: I. Bombaci, I. Parenti, and I. Vidana, Astrophys. J {\bf 614}, 314 (2004).
854:
855: \bibitem{Prakash}
856: M. Prakash, I. Bombaci, M. Prakash, P. J. Ellis, J. M. Lattimer and R. Knorren, Phys. Rep. {\bf 280}, 1 (1997).
857:
858: \bibitem{Meyer}
859: B. S. Meyer, Astrophys. J. {\bf 343}, 254 (1989).
860:
861: \bibitem{Glenny}
862: N. K. Glendenning, {\it Compact Stars}, 2$^{\rm nd}$ ed. (Springer-Verlag, New York).
863:
864: \bibitem{Weber}
865: F. Weber, Prog. Part. Nucl. Phys. {\bf 54}, 193 (2005).
866:
867: \end{thebibliography}
868: \end{flushleft}
869: \end{document}
870: %%%%%%%%%%%%%%%%%%%%%%%%%%% THE END %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
871: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
872: 9 pages, 3 figures. Contribution to the proceedings of the IXth Workshop on High Energy Physics Phenomenology (WHEPP-9), Bhubaneswar, India, 3-14 Jan. 2006
873:
874: