1: \documentclass{article}
2:
3: \bibliographystyle{livrevrel}
4:
5: \usepackage{epubtk}
6: \usepackage{graphicx}
7: \usepackage[square,comma,sort]{natbib}
8:
9: %\showlistoftablesfalse
10:
11: \begin{document}
12:
13: \title{Probes and Tests of Strong-Field Gravity with Observations
14: in the Electromagnetic Spectrum}
15:
16: \author{%
17: \epubtkAuthorData{Dimitrios Psaltis}{%
18: Physics and Astronomy Departments\\
19: University of Arizona}{%
20: dpsaltis@physics.arizona.edu}{%
21: http//www.physics.arizona.edu/\~dpsaltis}%
22: }
23:
24: \date{}
25: \maketitle
26:
27: \begin{abstract}
28: Neutron stars and black holes are the astrophysical systems with the
29: strongest gravitational fields in the universe. In this article, I
30: review the prospect of probing with observations of such compact
31: objects some of the most intriguing General Relativistic predictions
32: in the strong-field regime: the absence of stable circular orbits near
33: a compact object and the presence of event horizons around black-hole
34: singularities. I discuss the need for a theoretical framework within
35: which future experiments will provide detailed, quantitative tests of
36: gravity theories. Finally, I summarize the constraints imposed by
37: current observations of neutron stars on potential deviations from
38: General Relativity.
39: \end{abstract}
40:
41: \epubtkKeywords{}
42:
43:
44: \newpage
45:
46: \section{Introduction}
47: \label{sec:int}
48:
49: Over the past 90 years, the basic ingredients of General Relativity
50: have been tested in many different ways and in many different
51: settings. From the solar eclipse expedition of 1917 to the modern
52: observations of double neutron stars, General Relativity has passed all
53: tests with flying colors~\cite{Will06}. Yet, our inability to devise a
54: renormalizable quantum gravity theory as well as the mathematical
55: singularities found in many solutions of Einstein's equations suggest
56: that we should look harder for gravitational phenomena not described
57: by General Relativity.
58:
59: The search for such deviations has been very fruitful in the regime of
60: very weak fields. Observations of high-redshift
61: supernovae~\cite{Perl97, Riess98} and of the cosmic microwave
62: background with WMAP~\cite{Spergel03} have measured a non-zero
63: cosmological constant (or a slowly rolling field that behaves as such
64: at late times). This discovery can be incorporated within the
65: framework of General Relativity, if interpreted simply as a constant
66: in the Einstein--Hilbert action. It nevertheless brought to the surface
67: a major problem in trying to connect gravity to basic ideas of quantum
68: vacuum fluctuations~\cite{Weinberg89,Carroll01}.
69:
70: In the strong-field regime, on the other hand, which is relevant for
71: the evolution of the very early universe and for determining the
72: properties of black holes and neutron stars, little progress has been
73: made in testing the predictions of general
74: relativity~\cite{Stairs03}. There are two reasons that have been
75: responsible for this lag. First, phenomena that occur in strong
76: gravitational fields are complex and often explosive, making it very
77: difficult to find observable properties that depend cleanly on the
78: gravitational field and that allow for quantitative tests of gravity
79: theories. Second, there exists no general theoretical framework within
80: which to quantify deviations from general relativistic predictions in
81: the strong-field regime.
82:
83: During the current decade, technological advances and increased
84: theoretical activity have led to developments that promise to make
85: strong-field gravity tests a routine in the near future. The first
86: generation of earth-based gravitational wave observatories (such as
87: LIGO~\cite{LIGO}, GEO600~\cite{GEO}, TAMA300~\cite{TAMA}, and
88: VIRGO~\cite{VIRGO}) as well as the Beyond Einstein Missions (such as
89: Constellation-X, LISA, and the Black Hole Imager~\cite{beyondEins})
90: will offer an unprecedented look into the near fields of black holes
91: and neutron stars. Moreover, recent ideas on quantum
92: gravity~\cite{Burgess04}, brane-world gravity~\cite{Maartens04}, or
93: other Lagrangian extensions of general
94: relativity~\cite{Woodard06,Sotiriou08} will provide the means with
95: which the experimental results will be interpreted.
96:
97: In this article, I review the theoretical and experimental prospects
98: of testing strong-field General Relativity with observations in the
99: electromagnetic spectrum. In the first few sections, I discuss the
100: motivation for performing such tests and then describe the
101: astrophysical settings in which strong-field effects can be
102: measured. In Section~\ref{sec:need}, I elaborate on the need for a
103: theoretical framework within which strong-field gravity tests can be
104: performed and in Section~\ref{sec:tests} I review the current
105: quantitative tests of General Relativity in the strong-field regime
106: that use neutron stars. Finally, in Section~\ref{section:beyond} I
107: discuss the prospect of probing and testing strong gravitational
108: fields with upcoming experiments and observatories.
109:
110:
111: \newpage
112:
113: \section{The motivation for strong-field tests}
114: \label{sec:motiv}
115:
116: Most physical scientists would agree that there is very little need to
117: motivate testing one of the fundamental theories of physics in a
118: regime that experiments have probed only marginally, so far. However,
119: in the particular case of testing the strong-field predictions of
120: General Relativity, there exist at least three arguments that provide
121: additional strong support to such an endeavor. First, there is no
122: fundamental reason to choose Einstein's equations over other
123: alternatives. Second, gravitational tests to date seldom probe strong
124: gravitational fields. Finally, it is known that General Relativity
125: breaks down at the strong-field regime. I will now elaborate on each
126: of these arguments.
127:
128: \smallskip
129:
130: \noindent $\bullet$ {\em There is no fundamental reason to choose
131: Einstein's equations over other alternatives.~--\/} All theories of
132: gravity, including Newton's theory and General Relativity, have two
133: distinct ingredients. The first describes how matter moves in the
134: presence of a gravitational field. The second describes how the
135: gravitational field is generated in the presence of matter. For
136: Newtonian dynamics, the first ingredient is Newton's second law
137: together with the assertion that the gravitational and inertial masses
138: of an object are the same; the second ingredient is Poisson's
139: equation. For General Relativity, the first ingredient arises from the
140: equivalence principle, whereas the second is Einstein's field
141: equation.
142:
143: The equivalence principle, in its various formulations, dictates the
144: geometric aspect of the theory~\cite{Will06}: it is impossible to tell
145: the difference between a reference frame at rest and one free-falling
146: in a gravitational field, by performing local, non-gravitational (for
147: the Einstein Equivalence Principle) or even gravitational (for the
148: Strong Equivalence Principle) experiments. Moreover, the equivalence
149: principle encompasses the Lorentz symmetry, as well as our belief that
150: there is no preferred frame and position anywhere in the
151: spacetime. Because of its central importance in any gravity theory,
152: there have been many attempts during the last century at testing the
153: validity of the equivalence principle. These were performed mostly in
154: the weak-field regime and have resulted in upper limits on possible
155: violations of this principle that are as stringent as one part in
156: $10^{12}$~\cite{Will06}.
157:
158: Contrary to the case of the equivalence principle, there are no
159: compelling arguments one can make that lead uniquely to Einstein's
160: field equation. In fact, Einstein reached the field equation, more or
161: less, by reverse engineering (see the informative discussion
162: in~\cite{Misner73, Pais82}) and, soon afterwards, Hilbert
163: constructed a Lagrangian action that leads to the same equation. The
164: Einstein--Hilbert action is directly proportional to the Ricci scalar,
165: $R$,
166: %
167: \begin{equation}
168: S= \frac{c^4}{16\pi G}\int d^4 x \, \sqrt{-g} \, (R-2\Lambda)\;,
169: \label{EHaction}
170: \end{equation}
171: %
172: where $g\equiv\det\vert g_{\mu\nu}\vert$, $g_{\mu\nu}$ is the
173: spacetime metric, $c$ is the speed of light, $G$ is the gravitational
174: constant, and $\Lambda$ is the cosmological constant. While such a
175: theory is entirely self-consistent at the classical level, it may
176: represent only an approximation that is valid at the scales of
177: curvature that are found in terrestrial, solar, and stellar-system
178: tests.
179:
180: Indeed, a self-consistent theory of gravity can also be constructed
181: for any other action that obeys the following four simple
182: requirements~\cite{Misner73}. It has to: {\em (i)\/} reproduce the
183: Minkowski spacetime in the absence of matter and of the cosmological
184: constant, {\em (ii)\/} be constructed from only the Riemann curvature
185: tensor and the metric, {\em (iii)\/} follow the symmetries and
186: conservation laws of the stress-energy tensor of matter, and {\em
187: (iv)\/} reproduce Poisson's equation in the Newtonian limit. Of all
188: the possibilities that meet these requirements, the field equations
189: that are derived from the Einstein--Hilbert action are the only ones
190: that are also linear in the Riemann tensor. Albeit simple and elegant,
191: a more general classical action of the form~\cite{Sotiriou08}
192: %
193: \begin{equation}
194: S =\frac{c^4}{16\pi G}\
195: \int d^4 x \, \sqrt{-g} \,
196: f(R)\;,
197: \label{EHaction2}
198: \end{equation}
199: %
200: also obeys the same requirements. Indeed, the
201: action~(\ref{EHaction2}) results in a field equation that allows for
202: the Minkowski solution in the absence of matter, is constructed only
203: from the Riemann tensor, obeys the usual symmetries and conservation
204: laws~\cite{Wald84}, and can be made to produce negligible corrections
205: at the small curvatures probed by weak-field gravitational
206: experiments. On the other hand, the predictions of the theory may be
207: significantly different at the strong curvatures probed by
208: gravitational tests involving compact objects.
209:
210: The single, rank-2 tensor field $g_{\mu\nu}$ (i.e., the metric) of the
211: Einstein--Hilbert action may also not be adequate to describe
212: completely the gravitational force (although, if additional fields are
213: introduced, then the strong equivalence principle is violated, with
214: important implications for the frame- and time- dependence of
215: gravitational experiments). In fact, a variant of such theories with
216: an additional scalar field, the Brans--Dicke theory~\cite{Brans61},
217: has been the most widely used alternative to General Relativity to be
218: tested against experiments. Today, scalar-tensor theories are among
219: the prime candidates for explaining the acceleration of the universe
220: at late times (the ``dark energy''~\cite{Peebles03}). Depending on the
221: coupling between the metric, the scalar field, and matter, the
222: relative contribution of such additional fields may become significant
223: only at the high curvatures found in early universe or in the vicinity
224: of compact objects.
225:
226: Although the above discussion has considered only the classical action
227: of the gravitational field in a phenomenological manner, it is
228: important to also note that corrections to the Einstein--Hilbert action
229: occur naturally in quantum gravity theories and in string theory. For
230: example, if we choose to interpret the metric $g_{\mu\nu}$ as a
231: quantum field, we can take Equation~(\ref{EHaction}) as a quantum
232: field-theoretic action defined at an ultraviolet scale (such as the
233: Planck scale), and proceed to perform quantum-mechanical calculations
234: in the usual way~\cite{Donoghue94}. However, radiative corrections
235: will induce an infinite series of counterterms as we flow to lower
236: energies and such counterterms will not be reabsorbed into the
237: original Lagrangian by adjusting its bare parameters. Instead, such
238: terms will appear as new, higher-derivative correction terms in the
239: Einstein--Hilbert action~(\ref{EHaction}).
240:
241: Finally, it is worth emphasizing that the previous discussion focuses
242: on Lagrangian gravity in a four-dimensional spacetime. In the context
243: of string theory, General Relativity emerges only as a leading
244: approximation. String theory also predicts an infinite set of
245: non-linear terms in the scalar curvature, all suppressed by powers of
246: the Planck scale. Moreover, the low-energy effective action of string
247: theory contains additional scalar (dilatonic) and vector gravitational
248: fields~\cite{Green88}. Motivated by ideas of string theory, brane-world
249: gravity~\cite{Maartens04,DGP00a,DGP00b,DGP00c} also provides a
250: self-consistent theory that is consistent with all current tests of
251: gravity.
252:
253: All the above strongly support the notion that the field equation that
254: arises from the Einstein--Hilbert action may be appropriate only at the
255: scales that have been probed by current gravitational tests. But how deep
256: have we looked?
257:
258: \smallskip
259:
260: \noindent $\bullet$ {\em Gravitational tests to date seldom probe
261: strong gravitational fields.~--\/} All historical tests of general
262: relativity have been performed in our solar system. The strongest
263: gravitational field they can, therefore, probe is that at the surface
264: of the Sun, which corresponds to a gravitational redshift of
265: %
266: \begin{equation}
267: z_\odot\simeq \frac{GM_\odot}{R_\odot c^2}\simeq 2\times 10^{-6}\;,
268: \end{equation}
269: %
270: and to a spacetime curvature of
271: %
272: \begin{equation}
273: \frac{GM_\odot}{R_\odot^3 c^2}\simeq 4\times 10^{-28}~\mbox{cm}^{-2}\;.
274: \end{equation}
275: %
276: Coincidentally, the gravitational fields that have been probed in
277: tests using double neutron stars are of the same magnitude, since the
278: masses and separation of the two neutron stars in the systems under
279: consideration are comparable to the mass and radius of the Sun,
280: respectively\footnote{Note, however, that some of the phenomena
281: observed in double neutron stars depend on the coupling of matter and
282: gravity in the strong-field regime~\cite{Damour93}.}. These are
283: substantially weaker fields than those found in the vicinities of
284: neutron stars and stellar-mass black holes, which correspond to a
285: redshift of $\sim 1$ and a spacetime curvature of $\simeq
286: 2\times 10^{-13}$~cm$^{-2}$.
287:
288: It is instructive to compare the degree to which current tests verify
289: the predictions of General Relativity to the increase in the strength
290: of the gravitational field going from the solar system to the vicinity
291: of a compact object. Current constraints on the deviation of the PPN
292: parameters from the General Relativistic predictions are of order
293: $\simeq 10^{-5}$~\cite{Will06}. It is conceivable, therefore, that
294: deviations consistent with these constraints can grow and become of order
295: unity when the redshift of the gravitational field probed is increased
296: by six orders of magnitude and the spacetime curvature by fifteen!
297:
298: Is it possible, however, that General Relativity still describes
299: accurately phenomena that occur in the strong gravitational fields
300: found in the vicinity of stellar-mass black holes and neutron stars?
301:
302: \smallskip
303:
304: \noindent $\bullet$ {\em General Relativity breaks down at the
305: strong-field regime.~--\/} Our current understanding of the physical
306: world leaves very little doubt that the theory of General Relativity
307: itself breaks down at the limit of very strong gravitational
308: fields. Considering the theory simply as a classical, geometric
309: description of the spacetime leads to predictions of infinite matter
310: densities and curvatures in two different settings. Integrating
311: forward in time the Oppenheimer--Snyder equations, which describe the
312: collapse of a cloud of dust~\cite{Oppenheimer39}, leads to the
313: formation of a black hole with a singularity at its
314: center. Integrating backwards in time the Friedmann equation, which
315: describes the evolution of a uniform and isotropic universe, always
316: results in a singularity at the beginning of time, the big
317: bang. Clearly, the outcome in both of these settings is unphysical.
318:
319: It is widely believed that quantum gravity prohibits these unphysical
320: situations that occur at the limit of infinitely strong gravitational
321: fields. Even though none of the observable astrophysical objects
322: offer the possibility of testing gravity at the Planck scale, they
323: will nevertheless allow placing constraints on deviations from general
324: relativity that are as large as $\sim 10$ orders of magnitude more
325: stringent compared to all other current tests. This is the best result
326: we can expect in the near future to come out of the detection of
327: gravitational waves and the observation of the innermost regions of
328: neutron stars and black holes with NASA's Beyond Einstein missions. If
329: the history of the recent detection of a minute yet non-zero
330: cosmological constant is any measure of our inability to predict even
331: the order of magnitude of gravitational effects that we have not
332: directly probed, then we might be up for a pleasant surprise!
333:
334:
335: \newpage
336:
337: \section{Astrophysical and Cosmological Settings of Strong Gravitational
338: Fields}
339: \label{sec:strong}
340:
341:
342: \subsection{When is a gravitational field strong?}
343: \label{subsec:definition}
344:
345: Looking at the Schwarzschild spacetime, it is natural to measure the
346: ``strength'' of the gravitational field at a distance $r$ away from an
347: object of mass $M$ by the parameter
348: %
349: \begin{equation}
350: \epsilon\equiv \frac{GM}{rc^2}\;,
351: \end{equation}
352: %
353: which is proportional to the Newtonian gravitational potential and is
354: also directly related to the redshift. Infinitesimal gravitational
355: fields correspond to the limit $\epsilon\rightarrow 0$, which leads to
356: the Minkowski spacetime of special relativity. Weak gravitational
357: fields correspond to $\epsilon\ll 1$, which leads to Newtonian
358: gravity. Finally, the strongest gravitational fields accessible to an
359: observer are characterized by $\epsilon\rightarrow 1$, at which point
360: the black-hole horizon of an object of mass $M$ is approached. (Note
361: that formally the horizon of a Schwarzschild black hole occurs at
362: $\epsilon=2$; I drop here the factor of 2, as I am mostly interested
363: in dimensional arguments).
364:
365: Albeit useful in defining post-Newtonian expansions, the parameter
366: $\epsilon$ is not fundamental in characterizing a gravitational field
367: in Einstein's theory. Indeed, the geodesic equation and the Einstein
368: field equation (or equivalently, the Einstein--Hilbert
369: action~[\ref{EHaction}]) are written in terms of the Ricci scalar, the
370: Ricci tensor, and the Riemann tensor, all of which measure the
371: curvature of the field and not its potential. As a result, when we
372: consider deviations from General Relativity that arise by adding
373: linearly terms to the Einstein--Hilbert action, the critical strength
374: of the gravitational field beyond which these additional terms become
375: important is typically given in terms of the spacetime curvature.
376:
377: For example, in the presence of a cosmological
378: constant, the metric of a spherically symmetric object becomes
379: %
380: \begin{equation}
381: ds^2=-\left(1-\frac{2GM}{rc^2}-\frac{\Lambda r^2}{3}\right)dt^2
382: +\left(1-\frac{2GM}{rc^2}-\frac{\Lambda r^2}{3}\right)^{-1}dr^2
383: +r^2(d\theta^2+\sin^2\theta d\phi^2)
384: \end{equation}
385: %
386: and the Newtonian approximation becomes invalid when
387: %
388: \begin{equation}
389: \frac{GM}{c^2r^3}\ll \frac{1}{6}\Lambda\;.
390: \label{eq:curv_lambda}
391: \end{equation}
392: %
393: In this case, a gravitational field is ``weak'' if the spacetime
394: curvature is smaller than $\Lambda/3$, independent of the value of the
395: parameter $\epsilon$. In the opposite extreme, if there are additional
396: terms in the action of the gravitational field beyond the
397: Einstein--Hilbert term, such as
398: %
399: \begin{equation}
400: S =\frac{c^4}{16\pi G}\
401: \int d^4 x \, \sqrt{-g} \,
402: (R+\alpha R^2)\;,
403: \label{eq:EH3}
404: \end{equation}
405: %
406: then the General Relativistic predictions become inaccurate at
407: strong gravitational fields defined by the condition
408: %
409: \begin{equation}
410: \frac{GM}{c^2r^3}\gg \frac{1}{\alpha}\;,
411: \label{eq:curv_nonlinear}
412: \end{equation}
413: %
414: even if the parameter $\epsilon$ is much smaller than unity. Note
415: that, in Equation~(\ref{eq:EH3}), $\alpha$ is an appropriate constant
416: with units of (length)$^2$ and I have set the Ricci scalar to $R\sim
417: GM/r^3c^2$ (I use this here as an order of magnitude estimate and do
418: not consider the fact that, if the distance $r$ is larger than the
419: radius of the object, then the Ricci scalar in General Relativity
420: vanishes).
421:
422: Similar considerations lead to a condition on curvature when we add to
423: the Einstein--Hilbert action terms that invoke additional scalar,
424: vector, and tensor fields. In all these cases, a strong gravitational
425: field is characterized not by a large gravitational potential (i.e., a
426: high value of the parameter $\epsilon$) but rather by a large
427: curvature
428: %
429: \begin{equation}
430: \xi\equiv \frac{GM}{r^3c^2}\;.
431: \label{eq:xi}
432: \end{equation}
433: %
434: Because the condition that the curvature needs to satisfy in order for
435: a gravitational field to be considered ``strong'' depends on the
436: particular deviation from General Relativity under study (cf.\
437: Equations~[\ref{eq:curv_lambda}] and [\ref{eq:curv_nonlinear}]) I will
438: not normalize the parameter $\xi$ to any particular energy density but
439: rather leave it, hereafter, as a dimensional quantity.
440:
441: This is an appropriate parameter with which to measure the strength of
442: a gravitational field in a geometric theory of gravity, such as
443: General Relativity, because the curvature is the lowest order quantity
444: of the gravitational field that cannot be set to zero by a coordinate
445: transformation. Moreover, because the curvature measures energy
446: density, a limit on curvature will correspond to an energy scale
447: beyond which additional gravitational degrees of freedom may become
448: important.
449:
450:
451: \subsection{A parameter space for tests of gravity}
452: \label{subsec:parameters}
453:
454: The two parameters, $\epsilon$ and $\xi$, define a parameter space on
455: which we can quantify the strengths of the gravitational fields probed
456: by different tests of gravity (see Figure~\ref{fig:fields}). Only a
457: fraction of this parameter space is accessible to experiments. Regions
458: of the parameter space with potential $\epsilon>1$ correspond to
459: distances from a gravitating object that are smaller than the horizon
460: radius and are, therefore, inaccessible to observers. (I neglect here,
461: for simplicity, the small numerical factor in the horizon radius that
462: depends on the spin of the black hole.) In Figure~\ref{fig:fields},
463: this region is outlined by the vertical red line.
464:
465: \epubtkImage{strong_fields.png}{
466: \begin{figure}[htbp]
467: \centerline{\includegraphics[height=18.0cm]{strong_fields}}
468: \caption{A parameter space for quantifying the strength
469: of a gravitational field. The $x$-axis measures the potential
470: $\epsilon\equiv GM/rc^2$ and the $y$-axis measures the spacetime
471: curvature $\xi\equiv GM/r^3c^2$ of the gravitational field at a radius
472: $r$ away from a central object of mass $M$. These two parameters
473: provide two different quantitative measures of the strength of the
474: gravitational fields. The various curves, points, and legends are
475: described in the text.}
476: \label{fig:fields}
477: \end{figure}}
478:
479: Quantifying deviations from General Relativity for part of the
480: parameter space requires a detailed understanding of the properties of
481: dark matter and dark energy, which is beyond current capabilities. In
482: the limit of very small values of the curvature, the presence of a
483: non-zero cosmological constant affects the outcome of gravitational
484: experiments when (see eq.~[\ref{eq:curv_lambda}])
485: %
486: \begin{equation}
487: \xi\le \frac{3G\Omega_\Lambda H_0^2}{8\pi c^2}\simeq 5\times
488: 10^{-58}\left(\frac{\Omega_\Lambda}{0.73}\right)
489: \left(\frac{H_0}{73~\mbox{km/s/Mpc}}\right)^2~\mbox{cm}^{-2}\;,
490: \end{equation}
491: %
492: where $\Omega_\Lambda$ is the current density of dark energy in units
493: of the critical density and $H_0$ is the current value of the Hubble
494: constant. Phenomena that probe such low values of curvature (i.e.,
495: below the horizontal green line in Figure~\ref{fig:fields}) can lead to
496: quantitative tests of General Relativity only if a specific model of
497: dark energy (e.g., a cosmological constant) is assumed.
498:
499: The ability to perform a quantitative test of a gravity theory also
500: relies on an independent measurement of the mass that generates the
501: gravitational field. This is not always possible, especially in
502: various cosmological settings, where gravitational phenomena are used
503: mostly to infer the presence of dark matter and not to test General
504: Relativistic predictions. Dark matter is typically required in systems
505: for which the acceleration drops below the so-called MOND acceleration
506: scale $a_0\simeq 10^{-8}$~cm~s$^{-2}$~\cite{Milgrom83, Sanders02,
507: Bekenstein07}. (This is an observed fact, independent of whether the
508: inability of Newtonian gravity to account for observations is due to
509: the presence of dark matter or to the breakdown of the theory itself.)
510: This acceleration scale is also comparable to $a_0\simeq c
511: H_0$. Systems for which dark matter is necessary to account for their
512: gravitational fields are characterized by
513: %
514: \begin{equation}
515: \xi \le \left(\frac{a_0}{c^2}\right)^2 \frac{1}{\epsilon}
516: \simeq \left(\frac{H_0}{c}\right)^2 \frac{1}{\epsilon}\;.
517: \end{equation}
518: %
519: This region of the parameter space is outlined by the purple line in
520: Figure~\ref{fig:fields}. The fact that the three lines that correspond
521: to the Schwarzschild horizon, the MOND acceleration scale, and the
522: dark energy all seem to intersect roughly in one point in the
523: parameter space is directly related to the cosmic coincidence problem,
524: i.e., that the universe is flat, with comparable amounts of (mostly
525: dark) matter and dark energy.
526:
527: In the opposite limit of very strong gravitational fields, General
528: Relativity is expected to break down when quantum effects become
529: impossible to neglect. This is expected to happen if a gravitational
530: test probes a distance from an object of mass $M$ that is comparable
531: to the Compton wavelength $\lambda_{\rm C}\equiv h/Mc$, where $h$
532: is Planck's constant. Quantum effects are, therefore, expected to
533: dominate when
534: %
535: \begin{equation}
536: \xi\ge \frac{1}{L_{\rm P}^2}\epsilon^2\;,
537: \end{equation}
538: %
539: where $L_{\rm P}\equiv (Gh/c^3)^{1/2}\simeq 4\times 10^{-33}$~cm is
540: the Planck length. This part of the parameter space is not shown in
541: Figure~\ref{fig:fields}, as it is many orders of magnitude away from
542: the values of the parameters that correspond to astrophysical systems.
543:
544: Having defined the parameter space and outlined the various limiting
545: cases, I can now identify the various astrophysical systems that
546: probe its various regimes. In general, systems of constant central
547: mass $M$ will follow curves of the form
548: %
549: \begin{equation}
550: \xi=\frac{c^4}{G^2M^2}\epsilon^3\;,
551: \end{equation}
552: %
553: whereas probes at a constant distance $r$ away from the central object
554: will follow curves of the form
555: %
556: \begin{equation}
557: \xi=\frac{1}{r^2}\epsilon\;.
558: \end{equation}
559: %
560: Figure~\ref{fig:fields} shows a number of representative contours of
561: constant mass and distance.
562:
563: The strongest gravitational fields around astrophysical systems can be
564: found in the vicinities of neutron stars (NS in
565: Figure~\ref{fig:fields}) and black holes in X-ray binaries
566: (XRB). Large gravitational potentials but smaller curvatures can be
567: found around the horizons of intermediate mass black holes ($\sim
568: 10^2\mbox{\,--\,}10^4\,M_\odot$; IMBHs) and in active galactic nuclei
569: ($10^6\mbox{\,--\,}10^{10}\,M_\odot$; AGN). Weaker gravitational fields exist near
570: the surfaces of white dwarfs (WD), main-sequence stars (MS), or at the
571: distances of the various planets in our solar system (SS). Finally,
572: even weaker gravitational fields are probed by observations of the
573: motions of stars in the vicinity of the black hole in the center of
574: the Milky Way (Sgr~A$^*$), and by studies of the rotational curve of
575: the Milky Way (MW) and other galaxies. In placing the various systems
576: on the parameter space shown in Figure~\ref{fig:fields}, I have used a
577: typical mass-radius relation for neutron stars and white
578: dwarfs~\cite{Shapiro84}, the calculated mass-radius relation of
579: main-sequence stars~\cite{Clayton83}, and the inferred mass-radius
580: profile of the inner region around Sgr~A$^*$~\cite{Schodel02}, which
581: smoothly approaches the mass profile inferred from the rotation curve
582: of the Milky Way~\cite{Dehnen98}.
583:
584: \epubtkImage{tests.png}{
585: \begin{figure}[t]
586: \centerline{\includegraphics[width=12.0cm]{tests}}
587: \caption{Tests of General Relativity placed on an appropriate
588: parameter space. The long-dashed line represents the event horizon
589: of Schwarzschild black holes.}
590: \label{fig:tests}
591: \end{figure}}
592:
593: Current tests of General Relativity with astrophysical objects probe a
594: wide range of gravitational potentials and curvatures (see
595: Figure~\ref{fig:tests}). However, they fall short of probing the most
596: extreme phenomena that are predicted by the theory to occur in the
597: vicinities of compact objects. For example, tests during solar
598: eclipses, with double neutron stars (such as the Hulse--Taylor pulsar),
599: or with Grav Prob B probe curvatures that are the same as those found
600: near the horizons of supermassive black holes, but potentials that are
601: smaller by six to ten orders of magnitude. Moreover, all these tests
602: probe curvatures that are smaller by thirteen or more orders of
603: magnitude from those found near the surfaces of neutron stars and the
604: horizons of stellar-mass black holes. Future experiments, such as the
605: gravitational wave detectors and the Beyond Einstein missions, will
606: offer for the first time the opportunity to probe directly such strong
607: gravitational fields.
608:
609: The whole range of gravitational fields, from the weakest to the
610: strongest, can also be found during various epochs of the evolution of
611: the universe. As a result, observations of cosmological phenomena may
612: also probe very strong gravitational fields. The scalar curvature of
613: a flat universe is given by
614: %
615: \begin{equation}
616: R=\frac{6}{\alpha^2}\left(\alpha \ddot{\alpha}+\dot{\alpha}^2\right)\;,
617: \end{equation}
618: %
619: where $\alpha$ is the scale factor. Using the Friedmann equation,
620: the scalar curvature becomes
621: %
622: \begin{equation}
623: R=3\left(\frac{\Omega_{\rm m}^0}{\alpha^3}+4\Omega_{\rm d}^0\right)
624: \left(\frac{H_0}{c}\right)^2\;,
625: \label{eq:Rcosmo}
626: \end{equation}
627: %
628: where $\Omega_{\rm m}^0$ and $\Omega_{\rm d}^0$ are the
629: (non-relativistic) matter and dark energy densities in the present
630: universe, respectively, in units of the critical
631: density. Equation~(\ref{eq:Rcosmo}) shows that, at late times, the
632: radius of curvature of the universe is comparable to the Hubble
633: distance.
634:
635: The evolution of the scalar curvature with redshift for a flat
636: universe and for the best-fit cosmological parameters obtained by the
637: WMAP mission~\cite{Spergel03} is shown in Figure~\ref{fig:cosmo}.
638: Identified on this figure are several characteristic epochs that have
639: been used in testing General Relativistic predictions: the $z\simeq 1$
640: epoch of type I supernovae that are used to measure the value of the
641: cosmological constant~\cite{Perl97, Riess98}; the $z\simeq 1000$ epoch
642: at which the acoustic peaks of the cosmic microwave background
643: observed by WMAP are produced; and the period of nucleosynthesis
644: during which the temperature of the universe was in the range 60~keV
645: -- 1~MeV~\cite{Santiago97, Carroll02}. The period of big-bang
646: nucleosynthesis is the earliest epoch for which quantitative tests
647: have been performed. The corresponding scalar curvature of the
648: universe at that time, however, is still small and comparable to the
649: curvatures of gravitational fields probed by current tests of General
650: Relativity in the solar system. It was only when the temperature of
651: the Universe was $\sim 100$~GeV that its curvature was $\simeq
652: 10^{-12}$~cm$^{-2}$, i.e., comparable to that found around a neutron
653: star or stellar-mass black hole. This is the period of electroweak
654: baryogenesis, for which no detailed theoretical models or data exist
655: to date.
656:
657: \epubtkImage{cosmology.png}{
658: \begin{figure}[htbp]
659: \centerline{\includegraphics[width=12.0cm]{cosmology}}
660: \caption{The scalar curvature of our universe, as a function
661: of redshift. The curve corresponds to a flat universe with the
662: best-fit values of the cosmological parameters obtained by the WMAP
663: mission~\cite{Spergel03}. The arrows point to the curvature and
664: redshift of the universe during various epochs.}
665: \label{fig:cosmo}
666: \end{figure}}
667:
668:
669: \subsection{Probing versus testing strong-field gravity}
670: \label{subsec:probvstest}
671:
672: The parameter space shown in Figure~\ref{fig:fields} is useful in
673: identifying the strength of the gravitational field probed by a
674: particular test of gravity. However, it is important to emphasize that
675: probing a gravitational field of a given strength is not necessarily
676: the same as testing General Relativity in that regime. I discuss
677: bellow the difference with two examples from scalar-tensor gravity
678: that illustrate the two opposite extremes.
679:
680: First, a phenomenon that occurs in a weak gravitational field may
681: actually be testing the strong-field regime of gravity. In General
682: Relativity, Birkhoff's theorem states that the external spacetime of a
683: spherically symmetric object is described by the Schwarzschild metric,
684: independent of the properties of the object itself. Birkhoff's
685: theorem, however, does not apply to a variety of gravity theories,
686: such as scalar-tensor or non-linear (e.g., $R+R^2$) theories. In fact,
687: in these theories, the spacetime at any point around a spherically
688: symmetric object depends on the mass distribution that generates the
689: spacetime, which may itself lie in a strong gravitational field and,
690: therefore, probe that regime of the theory. For example, in
691: Brans--Dicke gravity, which is a special case of scalar-tensor
692: theories, the evolution of the binary orbit in a system with two
693: neutron stars due to the emission of gravitational waves depends on
694: the coupling of matter to the scalar field, which occurs in the strong
695: gravitational field of each neutron
696: star~\cite{Eardley75, Will89, Damour96}. As a result, even though the
697: gravitational field that corresponds to a double-neutron star orbit is
698: rather weak (see Figure~\ref{fig:tests}), observations of the orbital
699: decay of the binary actually test General Relativity against
700: scalar-tensor theories in the strong-field regime~\cite{Damour96}.
701:
702: In the opposite extreme, phenomena that probe strong gravitational
703: fields cannot necessarily be used in testing General Relativity in
704: this regime. Analytical and numerical studies strongly suggest that
705: the end state of the collapse of a star in Brans--Dicke gravity is a
706: black hole described by the Kerr spacetime of General
707: Relativity~\cite{Thorne71, Bekenstein72, Hawking72, Scheel95,
708: Psaltis07b}. Therefore, the observation of a phenomenon that occurs
709: even just above the horizon of a black hole cannot be used in testing
710: General Relativity against Brans--Dicke gravity in the strong-field
711: regime, because both theories make the exact same prediction for that
712: phenomenon.
713:
714: In the following, I will distinguish attempts to probe phenomena that
715: occur exclusively in the strong-field regime of General Relativity
716: from those that aim to test the strong-field predictions of the theory
717: against various alternatives.
718:
719:
720: \newpage
721:
722: \section{Probing Strong Gravitational Fields with Astrophysical Objects}
723: \label{sec:probes}
724:
725: A number of astrophysical objects offer the possibility of detecting
726: directly the observable consequences of two strong-field predictions
727: of General Relativity that have no weak-field or Newtonian
728: counterparts: the presence of a horizon around a collapsed object and
729: the lack of stable circular orbits in the vicinity of a neutron star
730: or black hole. As in most other areas of astrophysics research, we
731: have to rely on imaging, spectral, or timing observations in order to
732: reveal the information of the strong-field effects that is encoded in
733: the detected photons. The construction of gravitational wave
734: observatories will offer, for the first time in the near future, a
735: wealth of additional probes into the inner workings of gravitational
736: fields in the vicinities of compact objects.
737:
738: In the following, I review a number of recent attempts to probe
739: strong-field phenomena that have used a variety of techniques and were
740: applied to different astrophysical objects. I will only discuss
741: phenomena that are observable in the electromagnetic spectrum and
742: refer to a number of excellent reviews on the gravitational phenomena
743: that are anticipated to be detected by gravitational wave
744: observatories~\cite{Hugh00, Flanagan05}.
745:
746:
747: \subsection{Black hole images}
748: \label{subsec:image}
749:
750: To paraphrase the common proverb, {\em a picture is worth a thousand
751: spectra.} Directly imaging the vicinity of a black hole promises to
752: provide a direct evidence for the existence of a horizon. However,
753: black holes are notoriously small, and the resolution required for
754: imaging their horizons is, for most cases, beyond current
755: capabilities. For a stellar-mass black hole in the galaxy, the opening
756: angle of the horizon, as viewed by an observer on Earth, is only
757: %
758: \begin{equation}
759: \theta=2\times 10^{-4}\left(\frac{M}{10\,M_\odot}\right)
760: \left(\frac{1~\mbox{kpc}}{D}\right)~\mu\mbox{arcsec}\;,
761: \end{equation}
762: %
763: where $M$ is the mass of the black hole and $D$ is its distance.
764:
765: For a supermassive black hole in a distant galaxy, the opening angle is
766: %
767: \begin{equation}
768: \theta=20\left(\frac{M}{10^9\,M_\odot}\right)
769: \left(\frac{1~\mbox{Mpc}}{D}\right)~\mu\mbox{arcsec}\;.
770: \end{equation}
771: %
772: This is shown in Figure~\ref{fig:images} for a number of supermassive
773: black holes with secure mass determinations. The angular size of the
774: horizons of some of the sources are barely resolvable today with
775: interferometric observations in the sub-mm/infrared wavelengths and
776: will be resolvable in the X-rays in near future with the Black Hole
777: Imager~\cite{Ozel01}.
778:
779: \epubtkImage{images.png}{
780: \begin{figure}[htbp]
781: \centerline{\includegraphics[width=12.0cm]{images}}
782: \caption{The opening angles, as viewed by an observer on
783: Earth, of the horizons of a number of supermassive black holes in
784: distant galaxies with a secure dynamical mass measurement (sample
785: of~\cite{Tremaine02}). The opening angle of the black hole horizon in
786: the center of the Milky Way (Sgr~A$^*$) is also shown for comparison.}
787: \label{fig:images}
788: \end{figure}}
789:
790: The black hole that combines the highest brightness with the largest
791: angular size of the horizon is the one that powers the source
792: Sgr~A$^*$, in the center of the Milky Way. Since the first
793: measurements of the size of the source at 7~mm~\cite{Lo98} and at
794: 1.4~mm~\cite{Krichbaum98} demonstrated that the emitting region is only
795: a few times larger than the radius of the horizon (see
796: Figure~\ref{fig:sgrsize}), a number of observational and theoretical
797: investigations have aimed to probe deeper into the gravitational field
798: of the black hole and constrain its properties.
799:
800: \epubtkImage{size.png}{
801: \begin{figure}[htbp]
802: \centerline{\includegraphics[width=12.0cm]{size}}
803: \caption{The major axis of the accretion flow around
804: the black hole in the center of the Milky Way, as measured at
805: different wavelengths, in units of the Schwarzschild radius (left axis) and
806: in milliarcsec (right axis; adapted from~\cite{Shen05}). Even with current
807: technology, the innermost radii of the accretion flow can be readily
808: observed.}
809: \label{fig:sgrsize}
810: \end{figure}}
811:
812: The long-wavelength spectrum of Sgr~A$^*$ peaks at a frequency of
813: $\simeq 10^{12}$~Hz, suggesting that the emission changes from
814: optically thick (probably synchrotron emission) to optically thin at a
815: comparable frequency (see, e.g., \cite{Narayan95}). As a
816: result, observations at frequencies comparable to or higher than the
817: transition frequency can, in principle, probe the accretion flow at
818: regions very close to the horizon of the black hole.
819:
820: Even though the exact shape and size of the image of Sgr~A$^*$ at long
821: wavelengths depends on the detailed structure of the underlying
822: accretion flow (cf.~\cite{Ozel00} and \cite{Yuan03}), there
823: exist two generic observable signatures of its strong gravitational
824: field. First, the horizon leaves a `shadow' on the image of the
825: source, which is equal to $\simeq\sqrt{27}GM/c^2$ and roughly
826: independent of the spin of the black hole~\cite{Bardeen73, Falcke00,
827: Takahashi04, Broderick06,Noble07}. Second, the brightness of the image of the
828: accretion flow is highly non-uniform because of the high velocity of
829: the accreting plasma and the effects of the strong gravitational
830: lensing. Simultaenously fitting the size, shape, polarization map, and
831: centroid of the image observed at different wavelengths with future
832: telescopes, will offer the unique possibility of removing the
833: complications introduced by the unknown nature of the accretion flow,
834: imaging directly the black hole shadow, and measuring the spin of the
835: black hole~\cite{Broderick06b}.
836:
837:
838: \subsection{Continuum spectroscopy of accreting black holes}
839: \label{subsec:contspec}
840:
841: There have been at least three different efforts published in the
842: literature that use the luminosities and the continuum spectra of
843: accreting black holes to look for evidence of strong-field phenomena.
844:
845: \epubtkImage{bh_vs_ns.png}{
846: \begin{figure}[htbp]
847: \centerline{\includegraphics[width=10.0cm]{bh_vs_ns}}
848: \caption{The 2--20~keV quiescent luminosities of black hole
849: candidates (filled circles) and neutron stars (open circles) in units
850: of the Eddington luminosity for different galactic binary systems, as
851: a function of their orbital periods, which are thought to determine
852: the mass transfer rate between the two stars. The systematically lower
853: luminisoties of the black hole systems have been attributed to the
854: presence of the event horizon~\cite{Narayan97, McClintock04}.}
855: \label{fig:bhvsns}
856: \end{figure}}
857:
858: \subsubsection{Luminosities of black holes in quiescence and the absence
859: of a hard surface}
860: \label{subsub:horizon}
861:
862: More low-mass X-ray binaries are stellar systems in which the primary
863: star is a compact object and the secondary star is filling its Roche
864: lobe. Matter is transferred from the companion star to the compact
865: object and releases its gravitational potential energy mostly as
866: high-energy radiation, making these systems the brightest sources in
867: the X-ray sky~\cite{Psaltis06, McClintock06}.
868:
869: The rate with which mass is transfered from the companion star to the
870: compact object is determined by the ratio of masses of the two stars,
871: the evolutionary state of the companion star, and the orbital
872: separation~\cite{Verbunt93}. On the other hand, the
873: rate with which energy is released in the form of high-energy
874: radiation depends on the rate of mass transfer, the state of the
875: accretion flow (i.e., whether it is via a geometrically thin disk or a
876: geometrically thick but radiatively inefficient flow), and on whether the
877: compact object has a hard surface or an event horizon. Indeed, for a
878: neutron-star system in steady state, most of the released
879: gravitational potential energy has to be radiated away (only a small
880: fraction heats the stellar core~\cite{Brown98}), whereas
881: for a black hole system, a significant amount of the potential energy
882: may be advected inwards past the event horizon, and hence may be forever
883: lost from the observable universe. For similar systems, in the same
884: accretion state, one would therefore expect black holes to be
885: systematically underluminous than neutron stars~\cite{Narayan97}.
886:
887: The luminosities of transient black holes and neutron stars in their
888: quiescent states most clearly show this trend. When plotted against
889: the orbital periods of the binary systems, which are used here as
890: observable proxies to the mass transfer rates, sources that are
891: believed to be black holes, based on their large masses, are
892: systematically underluminous (Figure~\ref{fig:bhvsns}
893: and~\cite{Narayan97, Garcia01, McClintock04}). Although the physical
894: mechanism behind the difference in luminosities is still a matter of
895: debate~\cite{Narayan97, Bildsten00, Lasota00}, the trend shown in
896: Figure~\ref{fig:bhvsns} appears to be a strong, albeit indirect,
897: evidence for the presence of an event horizon in compact objects with
898: masses larger than the highest possible mass of a neutron star.
899:
900: \subsubsection{Hard X-ray spectra of luminous black holes
901: and the presence of an event horizon}
902: \label{subsub:hard}
903:
904: Galactic black holes in some of their most luminous states (the
905: so-called very high states) have mostly thermal spectra in the soft
906: X-rays with power-law tails that extend well into the soft
907: $\gamma$-rays~\cite{Grove98}. It has been hypothesized that these
908: power-law tails are the result of Compton upscattering of soft X-ray
909: photons off the relativistic electrons that flow into the black hole
910: event horizon with speeds that approach the speed of light and,
911: therefore, constitute an observational signature of the presence of an
912: event horizon (e.g., see~\cite{Titarchuk98, Laurent99})
913:
914: A relativistic converging flow has indeed the potential of producing
915: power-law spectral tails (e.g., see~\cite{Payne81, Titarchuk97,
916: Psaltis01}). However, this mechanism is identical to a second order
917: Fermi acceleration and hence the power-law tail is a result of
918: multiple scatterings away from the horizon with small energy exchange
919: per scattering rather than the result of very few scatterings of
920: photons with ultrarelativistic electrons near the black hole
921: horizon~\cite{Psaltis97, Papathanassiou00}. Moreover, the model
922: spectra always cut-off at energies smaller than the electron rest
923: mass~\cite{Laurent99, Niedzwiecki06} whereas the observed spectra
924: extend into the MeV range~\cite{Grove98}. Successful theoretical
925: models of the power-law spectra of black holes that are based on
926: Comptonization of soft photons by non-thermal
927: electrons~\cite{Gierlinski99} as well as the discovery of similar
928: power-law tails in the spectra of accreting neutron stars that extend
929: to $\sim 100-200$~keV~\cite{DiSalvo01, DiSalvo06} have shown
930: conclusively that the observed power-law tails do not constitute
931: evidence for black hole event horizons.
932:
933:
934: \subsubsection{Measuring the radii of the innermost stable
935: circular orbits of black holes using continuum spectra}
936: \label{subsub:isco}
937:
938: The thermal spectrum of a black hole source in some of its most
939: luminous states is believed to originate in a geometrically thin
940: accretion disk. The temperature profile of such an accretion disk away
941: from the black hole is determined entirely by energy conservation and
942: is independent of the magnitude and properties of the mechanism that
943: transports angular momentum and allows for matter to accrete (as long
944: as this mechanism is local; see~\cite{Shakura73, Balbus99}). The
945: situation is very different, however, near the radius of the innermost
946: stable circular orbit (hereafter ISCO).
947:
948: Inside the ISCO, fluid elements cannot stay in circular orbits but
949: instead quickly loose centrifugal support and rapidly fall into the
950: black hole. The density of the accretion disk inside the ISCO is very
951: small and the viscous heating is believed to be strongly
952: diminished. It is, therefore, expected that only material outside the
953: ISCO contributes to the observed thermal spectrum. The temperature
954: profile of the accretion flow just outside the ISCO depends rather
955: strongly on the mechanism that transports angular momentum outwards
956: and in particular on the magnitude of the torque at the
957: ISCO~\cite{Krolik99, Gammie99, Agol00}. To lowest order, however, if
958: the entire accretion disk spectrum can be decomposed as a sum of
959: blackbodies, each at the local temperature of every radial annulus,
960: then the highest temperature will be that of the plasma near the ISCO
961: and the corresponding flux of radiation will be directly proportional
962: to the square of the ISCO radius.
963:
964: \epubtkImage{spin.png}{
965: \begin{figure}[htbp]
966: \centerline{\includegraphics[width=13.0cm]{spin}}
967: \caption{The spectra emerging from geometrically thin
968: accretion disks around black holes with different spins, but with the
969: same accretion luminosity~\cite{Davis05}.From left to right, the
970: curves correspond to spins ($a/M$) of 0, 0.2, 0.4, 0.6, 0.78, 0.881,
971: 0.936, 0.966, and 0.99. The spin values were chosen to give roughly
972: equal variation in the position of the spectral peak for spins $>0.8$.
973: The other parameters which determine the model are the viscosity
974: parameter, $\alpha=0.01$, the inclination of the observer, $\cos i=0.5$,
975: the mass of the black hole, $M=10\,M_\odot$, and the accretion
976: luminosity, $L=0.1 L_{\rm Edd}$. The peak energies of the spectra increase
977: with increasing spin, as a consequence of the fact that the ISCO
978: radius decreases with spin.}
979: \label{fig:bhspin}
980: \end{figure}}
981:
982: Phenomenological fits of multi-temperature blackbody models to the
983: observed spectra of black holes provide strong support to the above
984: interpretation. When model spectra are fit to observations of any
985: given black hole in luminosity states that differ by more than one
986: order of magnitude, the inferred ISCO radius remains approximately
987: constant~\cite{Tanaka95}. For systems with a dynamically measured mass
988: and with a known distance, such an observation can lead to a
989: measurement of the physical size of the ISCO and hence of the spin of
990: the black hole~\cite{Zhang97, Gierlinski01} (see
991: Figure~\ref{fig:bhspin}).
992:
993: There are a number of complications associated with producing the
994: model spectra of multitemperature blackbody disks that are required in
995: measuring spectroscopically the ISCO radius around a black
996: hole. First, as discussed above, the temperature profile of an
997: accretion disk at the region around the ISCO depends very strongly on
998: the details of the mechanism of angular momentum transport, which are
999: poorly understood~\cite{Krolik99, Gammie99, Agol00}. Second, the
1000: vertical structure of the disk at each annulus, which determines the
1001: emerging spectrum, may or may not be in hydrostatic equilibrium near
1002: the ISCO, as it is often assumed, and its structure depends strongly
1003: on the external irradiation of the disk plasma by photons that
1004: originate in other parts of the disk. Finally, material in the inner
1005: accretion disk is highly ionized and often far from local
1006: thermodynamic equilibrium, generating spectra that can be
1007: significantly different from blackbodies~\cite{Hubeny97}.
1008:
1009: There have been a number of approximate models of multi-temperature
1010: accretion disks that take into account some of these effects, in a
1011: phenomenological or in an {\em ab initio} way. The models of Li et
1012: al.~\cite{Li05} are based on the alpha-model for angular momentum
1013: transport, assume that the local emission from each annulus is a
1014: blackbody at the local temperature, and take into account the strong
1015: lensing of the emitted photons by the central black hole. On the other
1016: hand, the models of Davis et al.~\cite{Davis05} are the result of
1017: ionization-equilibrium and radiative transfer calculations at each
1018: annulus, they are based on the alpha model for angular momentum but
1019: allow for non-zero torques at the ISCO, and take into account the
1020: strong lensing of photons by the black hole.
1021:
1022: Given the flux $F$ of the accretion disk measured by an observer on
1023: Earth, the color temperature $T_{\rm col}$ that corresponds to the
1024: innermost region in the disk that is emitting (which presumably is
1025: near the ISCO), the distance $D$ to the source, and the mass $M$ of
1026: the black hole, the spin $\alpha$ of the black hole can be
1027: inferred~\cite{Zhang97} by equating the radius of the ISCO, i.e.,
1028: %
1029: \begin{equation}
1030: r_{\rm ISCO}=
1031: \sqrt{\frac{2GM}{c^2}}\left\{3+A_2\pm [(3-A_1)(3+A_1+2A_2)]^{1/2}\right\}
1032: \label{eq:fisco}
1033: \end{equation}
1034: %
1035: to the one inferred spectroscopically (since $F\sim T^4 R^2$) by
1036: %
1037: \begin{equation}
1038: r_{\rm spec}=
1039: D \left[\frac{F}{2\sigma g(\theta, \alpha)}\right]^{1/2}
1040: \left[\frac{f_{\rm col} f_{\rm GR}(\theta, \alpha)}{T_{\rm col}}\right]^2\;.
1041: \label{eq:rspec}
1042: \end{equation}
1043: %
1044: Here $A_1=1+(1-a^2)^{1/3}[(1+a)^{1/3}+(1-a)^{1/3}]$,
1045: $A_2=(3a^2+A_1^2)^{1/2}$, $a$ is the specific angular momentum per
1046: unit mass for the black hole, and the positive (negative) sign is
1047: taken for prograde (retrograde) disks. In these equations, $\sigma$ is
1048: the Stefan--Boltzmann constant and $\theta$ is the inclination of the
1049: observer with respect to the symmetry axis of the accretion disk. The
1050: functions $g(\theta, \alpha$) and $f_{\rm GR}(\theta, \alpha)$ are
1051: correction factors for the flux and the temperature, respectively,
1052: that need to be calculated when going from an accretion disk annulus
1053: to a distant observer and incorporate the combined effects of
1054: gravitational lensing, gravitational redshift, and Doppler boosting of
1055: the disk photons. Given a thickness of the accretion disk, both these
1056: transfer functions can be computed to any desired degree of
1057: accuracy. Finally, the factor $f_{\rm col}$ measures the ratio of the
1058: color temperature of the spectrum (as measured by fitting a blackbody
1059: to the observed spectrum) to the effective temperature in that annulus
1060: in the accretion disk (which is a measure of the total radiation flux
1061: emerging from that annulus). Computing the value of the factor $f_{\rm
1062: col}$ is the goal of the recent calculations of the ionization
1063: equilibrium and radiative transfer in accretion disks~\cite{Davis05}.
1064:
1065: Fitting these spectral models to a number of observations of black hole
1066: candidates with dynamically measured masses has resulted in
1067: approximate measurements of their spins: $a>0.7$ for
1068: GRS~1915$+$105~\cite{Midleton05, McClintock06b}; $a=0.75-0.85$ for
1069: 4U~1543-44~\cite{Shafee06}; $a=0.65-0.75$ for
1070: GRO~J1655$-$40~\cite{Shafee06}. It is remarkable that all inferred
1071: values of the black hole spins are high, comparable to the maximum
1072: allowed by the Kerr solution.
1073:
1074: Equations~(\ref{eq:fisco}) and (\ref{eq:rspec}) demonstrate the strong
1075: dependence of the inferred values of black hole spins on various
1076: observable quantities (the mass of, distance to, and inclination of
1077: the black hole, as well as the flux, and temperature of its disk
1078: spectrum) and on a model parameter (the color correction factor
1079: $f_{\rm col}$). Numerical simulations of magnetohydrodynamic flows
1080: onto black holes are finely tuned to resolve the length- and
1081: timescales of phenomena that occur in the vicinity of the horizon of a
1082: black hole (see, e.g.,~\cite{Gammie03, deVilliers03}). When such
1083: models incorporate accurate multi-dimensional radiative transfer, they
1084: will provide the best theoretical spectra to be compared directly to
1085: observations (see, e.g.,~\cite{Blaes06}). Moreover, monitoring of
1086: the same sources at long wavelengths will improve the measurements of
1087: their masses and distances. Finally, combination of this with other
1088: methods based on line spectra and the rapid variability properties of
1089: accreting black holes will enable us to tighten the uncertainties in
1090: the various model parameters and observed quantities that enter
1091: Equation~(\ref{eq:fisco}) and measure with high precision the spins of
1092: galactic black holes.
1093:
1094:
1095: \subsection{Line spectroscopy of accreting compact objects}
1096: \label{subsec:lines}
1097:
1098: Heavy elements on the surface layers of neutron stars or in the
1099: accretion flows around black holes that are not fully ionized generate
1100: atomic emission and absorption lines that can be detected by a distant
1101: observer with a large gravitational redshift. The value of the
1102: gravitational redshift can be used to uniquely identify the region in
1103: the spacetime of the compact object in which the observed photons are
1104: produced.
1105:
1106:
1107: \subsubsection{Atomic lines from the surfaces of neutron
1108: stars}
1109: \label{subsub:linesns}
1110:
1111: The gravitational redshift of an atomic line from the surface layer of
1112: a neutron star leads to a unique determination of the relation between
1113: its mass and radius. The detection of a rotationally broadened atomic
1114: line from a rapidly spinning neutron star offers the additional
1115: possibility of measuring directly the stellar radius~\cite{Ozel03,
1116: Chang06} and, therefore, of determining its mass, as well. The
1117: profile of a rotationally broadened atomic line can be used to study
1118: frame-dragging effects in the strong-field
1119: regime~\cite{Bhattacharya05}. Moreover, detecting a gravitationally
1120: redshifted and rotationally broadened atomic line can lead to a
1121: measurement of the oblateness of the spinning star~\cite{Cadeau07},
1122: which is determined by the strong-field coupling of matter with the
1123: gravitationally field. Unfortunately, this is one of the very few
1124: astrophysical settings discussed in this review in which observations
1125: significantly trail behind theoretical investigations.
1126:
1127: Despite many optimistic expectations and early claims (see,
1128: e.g.,~\cite{Lewin93}), the observed spectra of almost all
1129: weakly-magnetic neutron stars are remarkably featureless. The best
1130: studied case is that of the nearby isolated neutron star
1131: RX~J1856$-$3754, which was observed for 450~ks with the Chandra X-ray
1132: Observatory and showed no evidence for any atomic lines from heavy
1133: elements~\cite{Braje02}. This is in fact not surprising, given that
1134: heavy elements drift inwards of the photosphere in timescales of
1135: minutes~\cite{Bildsten92} and it takes only $\simeq 10^{-7}\,M_\odot$
1136: of light elements to blanket a heavy element surface.
1137:
1138: There are two types of neutron stars, however, in the atmospheres of
1139: which heavy elements may abound: young cooling neutron stars and
1140: accreting X-ray bursters~\cite{Ozel03}. On the one hand, the escaping
1141: latent heat of the supernova explosion makes young neutron stars
1142: relatively bright sources of X-rays. Their strong magnetic fields can
1143: inhibit accretion of light elements either from the supernova fallback
1144: or from the interstellar medium, leaving the surface heavy elements
1145: exposed. On the other hand, in the atmospheres of accreting, weakly
1146: magnetic neutron stars, heavy elements are continuously replenished.
1147: Moreover, large radiation fluxes pass through their atmospheres during
1148: thermonuclear bursts~\cite{Strohmayer06} making them very bright and
1149: easily detectable.
1150:
1151: The most promising detection to date of gravitationally redshifted
1152: lines from the surface of a neutron star came from an observation with
1153: XMM-Newton of the source EXO~0748$-$676, which showed redshifted
1154: atomic lines during thermonuclear flashes~\cite{Cottam02}. This is a
1155: slowly spinning neutron star (47~Hz~\cite{Villarreal04}) and hence its
1156: external spacetime can be accurately described by the Schwarzschild
1157: metric. In this case, the measurement of a gravitational redshift of
1158: $z=0.35$ leads to a unique determination of the relation between the
1159: mass and the radius of the neutron star, i.e., $M\simeq 1.4
1160: (R/10$~km$)M_\odot$. Combination of this result with the spectral
1161: properties of thermonuclear bursts during periods of photospheric
1162: radius expansion and in the cooling tales also allowed for an
1163: independent determination of the mass and radius of the neutron
1164: star~\cite{Ozel06}.
1165:
1166: Future observations of bursting or young neutron stars with upcoming
1167: X-ray missions such as Constellation-X~\cite{ConX} and
1168: XEUS~\cite{XEUS} have the potential of detecting many gravitationally
1169: redshifted atomic lines and, hence, of probing the coupling of matter
1170: to the strong gravitational fields found in the interiors of neutron
1171: stars.
1172:
1173:
1174: \subsubsection{Relativistically broadened iron lines in accreting
1175: black holes}
1176: \label{subsub:ironbh}
1177:
1178: Astrophysical black holes in active galactic nuclei accreting at
1179: moderate rates offer another possibility of probing strong
1180: gravitational fields using atomic spectroscopy (for an extensive
1181: review on the subject see~\cite{Reynolds03}; see also
1182: \cite{Miller07} for a review of iron line observations from
1183: stellar-mass black holes). The relatively cool accretion disks in
1184: these systems act as large mirrors, reflecting the high-energy
1185: radiation that is believed to be produced in the disk coronae by
1186: magnetic flaring~\cite{Guilbert88}. The spectrum of reflected
1187: radiation in hard X-rays is determined by electron scattering, whereas
1188: the spectrum in the soft X-rays is characterized by a large number of
1189: fluorescent lines caused by bound-bound transitions of the partially
1190: ionized material. The combination of the high yield and relatively
1191: high abundance of iron atoms in the accreting material make the iron
1192: K$\alpha$ line, with a rest energy of 6.4~keV for a neutral atom, the
1193: most prominent feature of the spectrum.
1194:
1195:
1196: \epubtkImage{iron_lines_a.png}{
1197: \begin{figure}[htbp]
1198: \centerline{\includegraphics[width=7.5cm]{iron_lines_alpha}
1199: \includegraphics[width=7.5cm]{iron_lines_a}}
1200: \caption{Theoretical models of relativistically broadened
1201: iron line profiles from accretion flows around black holes. The left
1202: panel shows the dependence of the line profile on the spin parameter
1203: of the black hole, whereas the right panel shows its dependence on the
1204: emissivity index (see text). All calculations were performed for an
1205: inclination angle of 40$^\circ$~\cite{Brenneman06}.}
1206: \label{fig:ironlines_th}
1207: \end{figure}}
1208:
1209: The profile of the fluorescent iron line as observed at infinity is
1210: determined mainly by general and special relativistic effects that
1211: influence the propagation of photons from the point of reflection to the
1212: observer~\cite{Laor91}. Dividing an accretion disk into a series of
1213: concentric rings orbiting at the local Keplerian frequency, special
1214: relativistic effects produce a rotational splitting of the line
1215: emerging from each ring, whereas general relativistic effects generate
1216: an overall redshift~\cite{Fabian89}. The combination of these effects
1217: integrated over the entire surface of the accretion disk leads to a
1218: characteristic profile for the iron reflection line, which is broad
1219: with a shallow and extended red wing (Figure~\ref{fig:ironlines_th}).
1220:
1221: The magnitude of the relativistic effects depends on the specifics of
1222: the spacetime of the black hole, on the position and orientation of
1223: the observer, on the position and properties of the source of X-rays
1224: above the accretion disk, and on the dependence of fluorescence yield
1225: on position on the accretion disk through its dependence on the
1226: ionization states of the elements~\cite{George91}. Given a model for
1227: the source of X-rays and the accretion disk, fitting the profile of an
1228: iron line from an accreting black hole can lead, in principle, to a
1229: direct mapping of its spacetime. Unfortunately, the source of X-ray
1230: illumination and the physical properties of the accretion flows
1231: themselves are poorly understood.
1232:
1233: If we make assumptions regarding these astrophysical complications
1234: that are largely model independent, a general property of the
1235: spacetime, such as the spin of the black hole, can be measured. The
1236: accretion disk is typically modeled as a geometrically thin reflecting
1237: surface at the rotational equator of the black hole that is extending
1238: inwards until the radius of the innermost stable circular orbit. Even
1239: though the density of the material inside this radius is significant
1240: and might reflect the illuminating X-rays, its ionization state
1241: changes rapidly, leading to small changes in the resulting iron line
1242: profile~\cite{Reynolds97, Brenneman06}. The extent of the iron line
1243: towards lower energies is a measure of the innermost radius of the
1244: accretion disk. By assumption, this radius is set as the radius of the
1245: innermost stable circular orbit, which depends on the spin of the
1246: black hole. Fitting theoretical models to observations can, therefore,
1247: lead to a measurement of the black hole spin.
1248:
1249: The uncertainties in the position of the illuminating source and in
1250: the disk structure are often modeled by a single function for the
1251: ``emissivity'' of the iron line, which measures the flux in the iron
1252: line that emerges locally from each patch on the accretion disk. This
1253: is typically taken to be axisymmetric and to have a power-law
1254: dependence on radius, i.e., $r^{-a}$. Increasing the emissivity index
1255: $a$ results in iron line profiles with more extended red wings, which
1256: is degenerate with increasing the spin of the black hole (see
1257: Figure~\ref{fig:ironlines_th} and~\cite{Beckwith04}). This uncertainty
1258: can introduce significant systematics in modeling iron-line profiles
1259: from slowly spinning black holes. For rapidly spinning black holes,
1260: however, masking the effect of the black hole spin by steepening the
1261: emissivity function requires an unphysically high value for the
1262: emissivity index~\cite{Brenneman06}.
1263:
1264: Since the original observation of broadened iron lines from the
1265: supermasive black hole MCG-6-15-30 with ASCA~\cite{Tanaka95b},
1266: observations of other active galactic nuclei with
1267: ASCA~\cite{Nandra97}, XMM-Newton~\cite{Nandra06}, and more recently
1268: with Suzaku~\cite{Reeves06} as well as of stellar mass black
1269: holes~\cite{Miller06} have revealed many more examples of such
1270: redshifted atomic lines. The best studied case remains MCG-6-15-30
1271: (see Figure~\ref{fig:ironlines_obs}), in which the extended red wing of
1272: the line has been discussed as evidence for a rapidly spinning black
1273: hole ($\alpha\ge 0.98$~\cite{Brenneman06}).
1274:
1275: \epubtkImage{iron_line_obs.pbg}{
1276: \begin{figure}[htbp]
1277: \centerline{\includegraphics[width=10.0cm]{iron_line_obs}}
1278: \caption{The 0.5-10~keV spectrum of the supermassive black hole
1279: in the center of the galaxy MCG-6-15-30 as observed with
1280: XMM-EPIC. Panel (a) shows the ratio of the observed spectrum to a
1281: power-law model and reveals the complicated structure of the
1282: residuals. Panel (b) shows the ratio of the observed spectrum to a
1283: model of the warm absorber, which accounts for the low-energy
1284: residuals. Panel (c) shows the 2-9~keV spectrum of the source
1285: together with a model of the relativistically broadened iron
1286: line~\cite{Wilms01}}
1287: \label{fig:ironlines_obs}
1288: \end{figure}}
1289:
1290: Perhaps the most challenging, although most rewarding to understand,
1291: property of iron lines is their time variability. Current observations
1292: of iron lines from accreting black holes (e.g., the one shown in
1293: Figure~\ref{fig:ironlines_obs}) are integrated over a time that is
1294: equal to many hundred times the dynamical timescale in the
1295: accretion-disk region where the lines are formed. As a result, an
1296: observed line profile is not the result of reflection from an
1297: accretion disk of a single flaring event, but rather the convolution
1298: of many such events that occurred over the duration of the
1299: observation. Moreover, the continuum spectrum of the black hole, which
1300: is presumably reflected off the accretion disk to produce the
1301: fluorescent iron line, changes over longer timescales, implying a
1302: correlated variability of the line itself.
1303:
1304: Observations with current instruments can only investigate the
1305: correlated variability of the iron line with the continuum spectrum
1306: (see, however,~\cite{Iwasawa04}). They have shown that the flux in the
1307: line remains remarkably constant even though the continuum flux
1308: changes by almost an order of magnitude~\cite{Fabian03}. General
1309: relativistic light bending, which leads to focusing of the
1310: photon rays towards the innermost regions of the accretion disk may
1311: be responsible for this puzzling effect~\cite{Miniutti04}.
1312:
1313: Future observations with upcoming X-ray missions such as
1314: Constellation~X~\cite{ConX} and XEUS~\cite{XEUS} will resolve the time
1315: evolution of the reflected iron line from a single magnetic
1316: flare~\cite{Reynolds99}. Because density inhomogeneities in the
1317: turbulent accretion flow move, roughly, in test-particle
1318: orbits~\cite{Armitage03}, the time evolution of the redshift of the
1319: iron line from a single flare reflected mainly off a localized density
1320: inhomogeneity will allow for a direct mapping of the spacetime around
1321: the black hole.
1322:
1323:
1324: \subsection{The fast variability of accreting compact objects}
1325: \label{subsec:var}
1326:
1327: The strongest gravitational fields in astrophysics can be probed only
1328: with rapidly variable phenomena around neutron stars and galactic
1329: black holes (see Figure~\ref{fig:tests_future}). Such phenomena have
1330: been discovered in almost all known accreting compact objects in the
1331: galaxy. They are quasi-periodic oscillations (QPOs) with frequencies
1332: in the range $\sim 1$~Hz$-1$~kHz that remain coherent for tens to
1333: hundreds of cycles and follow a rich and often complicated
1334: phenomenology (for an extensive review of the observations
1335: see~\cite{vanderKlis06}).
1336:
1337: \epubtkImage{qpo1820.pbg}{
1338: \begin{figure}[htbp]
1339: \centerline{\includegraphics[width=10.0cm,angle=-90]{qpo1820}}
1340: \caption{The dependence of the twin QPO frequencies
1341: on the X-ray countrate observed by the PCA instrument onboard RXTE,
1342: for the neutron-star source 4U~1820$-$30~\cite{Zhang98}. The
1343: flattening of the correlation at high frequencies has been discussed
1344: as a signature of the innermost stable circular orbit.}
1345: \label{fig:qpo_isco}
1346: \end{figure}}
1347:
1348:
1349: \subsubsection{Quasi-periodic oscillations in neutron stars}
1350:
1351: The fastest oscillations detected from accreting, weakly magnetic
1352: neutron stars are pairs of QPOs with variable frequencies that reach
1353: up to $\sim 1300$~Hz and with frequency separations of order $\sim
1354: 300$~Hz~\cite{vanderKlis06}. The origin of these oscillations is still
1355: a matter of debate. However, all current models associate at least one
1356: of the oscillation frequencies with a characteristic dynamical
1357: frequency in a geometrically thin accretion disk (see discussion
1358: in~\cite{Psaltis01a} and~\cite{Miller98, Stella99, Psaltis00b}).
1359:
1360: The highest dynamical frequency excited at any radius in an equatorial
1361: accretion disk around a compact object is the one associated with the
1362: circular orbit of a test particle at that radius~\cite{Bardeen73b};
1363: this is often referred to as the azimuthal, orbital, or Keplerian
1364: frequency. A mode in the accretion disk associated with this frequency
1365: can give rise to a long-lived quasi-periodic oscillation only if it
1366: lives outside the innermost stable circular orbit. The azimuthal
1367: frequency at this radius provides, therefore, an upper limit on the
1368: frequency of any observed oscillation~\cite{Kluzniak85, Miller98}. As
1369: a result, detecting such rapid oscillations offers the possibility of
1370: measuring the location and of understanding the properties of the
1371: region near the innermost stable circular orbit around a neutron star.
1372:
1373: \begin{figure}[t]
1374: \centerline{
1375: \includegraphics[width=10.0cm, height=10.5cm]{qpo1636}}
1376: \caption{
1377: The dependence of the amplitude and quality factor of the lower kHz QPO on
1378: its frequency for the neutron-star source
1379: 4U~1636$-$56~\cite{Barret06}. The drop of the QPO amplitude and
1380: coherence at high frequencies have been discussed as signatures of
1381: the innermost stable circular orbit.}
1382: \label{fig:qpo_isco2}
1383: \end{figure}
1384:
1385: The signature of the ISCO on the amplitudes and characteristics of the
1386: observed oscillations is hard to predict without a firm model for the
1387: generation of the oscillations in the X-ray flux. Two potential
1388: signatures have been discussed, however, based on the phenomenology of
1389: the oscillations. The first one is associated with the fact that the
1390: frequencies of the oscillations appear to increase roughly with
1391: accretion rate. When an oscillation frequency reaches that of the
1392: innermost stable circular orbit, one would expect its frequency to
1393: remain constant over a wide range of accretion
1394: rates~\cite{Miller98}. Such a trend has been observed in the
1395: quasi-periodic oscillations of the globular cluster source
1396: 4U~1820$-$30~\cite{Zhang98, Kaaret99}; Figure~\ref{fig:qpo_isco}). When
1397: observations of the source obtained over different epochs are
1398: combined, the dependence of the frequency of the fastest oscillation
1399: on the observed accretion rate appears to flatten at a value $\simeq
1400: 1050$~Hz. This is comparable to the azimuthal frequency at the
1401: innermost stable circular orbit for a $\simeq 2.1\,M_\odot$ neutron
1402: star~\cite{Zhang98}.
1403:
1404: Albeit suggestive, the interpretation of the 4U~1820$-$30 data relies
1405: on the assumption that the oscillatory frequencies in an accretion
1406: disk depend monotonically on the accretion rate and, furthermore, that
1407: the X-ray countrate is a good measure of the accretion rate. This
1408: assumption is probably justified for short timescales (of order one
1409: day) but is known to break down on longer timescales, such as those
1410: used in Figure~\ref{fig:qpo_isco}~\cite{vanderKlis01}. Indeed, in a
1411: given source, the same oscillation frequencies have been observed over
1412: a wide range of X-ray countrates and vice
1413: versa~\cite{vanderKlis01}. The hard X-ray color of a source and not
1414: the countrate appears to be a more unique measure of the accretion rate,
1415: which is presumably the physical parameter that determines the
1416: oscillation frequencies~\cite{Mendez99}. When the data of 4U~1820$-$30
1417: are plotted against hard color, the characteristic flattening seen in
1418: Figure~\ref{fig:qpo_isco} disappears~\cite{Mendez99}.
1419:
1420: A second signature of the innermost stable circular orbit is a
1421: potential decrease in the amplitude and coherence of the oscillations
1422: when the region where they are excited approaches the ISCO. Such a
1423: trend has been observed in a number of accreting neutron
1424: stars~(Figure~\ref{fig:qpo_isco2} and~\cite{Barret06, Barret07})
1425: and has been questioned on similar grounds as the study of
1426: 4U~1820$-$30~\cite{Mendez06}. The most significant criticism comes from
1427: the fact that the drop in amplitude and coherence is rather gradual
1428: and occurs over a $\sim 150$~Hz range of frequencies. Even assuming
1429: that this drop is a signature of the ISCO, measuring its location will
1430: be possible only within a detailed model of the frequencies of
1431: quasi-periodic oscillations.
1432:
1433: Among more model-dependent ideas, perhaps the most exciting prospect
1434: of probing strong-field gravity effects in neutron stars with
1435: quasi-periodic oscillations comes from applying the relativistic model
1436: of QPOs~\cite{Stella99} to the observed correlations between various
1437: pairs of QPO frequencies~\cite{Psaltis99}. In the relativistic model,
1438: the highest-frequency QPO is identified with the azimuthal frequency
1439: of a test particle in orbit at a given radius. The peak separation of
1440: this QPO with the second higher frequency QPO is identified as the
1441: radial epicyclic frequency of the test particle in the same orbit. A
1442: variant of this model can account for the observed correlations
1443: between oscillations frequencies, when hydrodynamic effects are taken
1444: into account~\cite{Psaltis00b}. Because the two observed frequencies
1445: are directly related to the azimuthal and radial frequencies at
1446: various radii in the accretion flow, interpretation of the data with
1447: this model can provide a direct map of the exterior spacetime of the
1448: neutron stars, to within the $\simeq 10$\% uncertainty introduced by
1449: the hydrodynamic corrections to the oscillation frequencies.
1450:
1451:
1452: \subsubsection{Quasi-periodic oscillations in black holes}
1453: \label{subsub:bhqpo}
1454:
1455: Pairs of rapid quasi-periodic oscillations have also been detected
1456: from a number of accreting systems that harbor black hole
1457: candidates~\cite{McClintock06}. The phenomenology of these
1458: oscillations is very different from the one discussed above for
1459: accreting neutron stars. The frequencies of the rapid oscillations
1460: observed in each source vary at most by a percent over a wide range of
1461: luminosities and their ratios are practically equal to ratios of small
1462: integers (2:3 for XTE~J1550$-$564 and GRO~J1655$-$40, 3:5 for
1463: GRS~1915$+$105, etc.).
1464:
1465: The high frequencies of the oscillations observed from black hole
1466: sources with dynamically measured masses demonstrate that they
1467: originate in regions very close to the black hole horizons. In fact,
1468: requiring the frequency of the 450~Hz oscillation observed from
1469: GRO~J1655$-$40 to be limited by the azimuthal frequency at the ISCO
1470: necessitates a spining black hole with a Kerr spin parameter $a/M\ge
1471: 0.25$~\cite{Strohmayer01}. Moreover, the frequencies of the observed
1472: oscillations are roughly inversely proportional to the black holes
1473: masses, as one would expect if they were associated to dynamical
1474: frequencies near the innermost stable circular
1475: orbit~\cite{Abramowicz04}.
1476:
1477: \epubtkImage{bhqpo_lin.png}{
1478: \begin{figure}[htbp]
1479: \centerline
1480: {\includegraphics[width=7.0cm]{bhqpo_lin}
1481: \includegraphics[width=7.0cm]{bhqpo_res}}
1482: \caption{
1483: (Left Panel) The intersection of the two solid lines shows the
1484: black hole mass and spin for the source GRO~J1655$-$40 for which the
1485: observed 300~Hz and the 450~Hz oscillations can be explained as the
1486: lowest order $c$- and $g$-modes, respectively. The intersection of the
1487: dotted lines makes the opposite identification of disk modes to the
1488: observed oscillatory frequencies (after~\cite{Wagoner01}). (Right
1489: Panel) Each solid line traces pairs of black hole mass and spin for
1490: which the observed frequencies correspond to different resonances
1491: between the Keplerian and periastron precession frequencies
1492: (after~\cite{Abramowicz02}). In both panels, the horizontal dashed
1493: lines show the uncertainty in the dynamically measured mass of the
1494: black hole.}
1495: \label{fig:bhqpo}
1496: \end{figure}}
1497:
1498: As in the case of neutron stars, using black hole quasi-periodic
1499: oscillations to probe directly strong gravitational fields is hampered
1500: by the lack of a firm understanding of the physical mechanism that is
1501: producing them. In one interpretation, they are associated with linear
1502: oscillatory modes that are trapped just outside the radius of the
1503: innermost stable circular orbit (for reviews see~\cite{Wagoner99,
1504: Kato01, Nowak01}). The frequencies of these modes depend primarily on
1505: the mass and spin of the black hole. Identifying the two observed
1506: oscillations with the lowest order linear modes, therefore, leads to
1507: two pairs of values for the mass and spin of the black hole (depending
1508: on which oscillation is identified with which mode). For example, for
1509: the case of the black hole GRO~J1655$-$40, one of the inferred pairs
1510: of values agrees with the dynamically measured mass of the black hole
1511: of $6.9\pm 1.0\,M_\odot$ and results in an estimated value of the
1512: black hole spin of $a/M\sim 0.9$ (Figure~\ref{fig:bhqpo}
1513: and~\cite{Wagoner01}). Although compelling, this interpretation leaves
1514: to coincidence the fact that the ratios of the oscillation frequencies
1515: are approximately equal to ratios of small integers.
1516:
1517: In an alternate model, the oscillations are assumed to be excited in
1518: regions of the accretion disks where two of the dynamical frequencies
1519: are in parametric resonance, i.e., their ratios are equal to ratios of
1520: small integers~\cite{Abramowicz02}. In this case, the
1521: frequencies of the oscillations depend on the mass and spin of the
1522: black hole as well as on the radius at which the resonance occurs. As
1523: a result, the observation of two oscillations from any given source
1524: does not lead to a unique measurement of its mass and spin, but rather
1525: to a families of solutions. For example, identifying the frequencies of
1526: the two oscillations observed from GRO~J1655$-$40 as a 3:2, a 3:1, or
1527: a 2:1 resonance between the Keplerian and the periastron precession
1528: frequencies at any radius in the accretion disk leads to three family
1529: of solutions, as shown in Figure~\ref{fig:bhqpo}. The dynamically
1530: measured mass of the black hole then picks only two of the possible
1531: families of solutions and leads to a smaller value for the inferred
1532: spin.
1533:
1534: Future observations of accreting neutron stars and black holes with
1535: upcoming missions that will have fast timing capabilities, such as
1536: XEUS~\cite{XEUS}, will be able to discover a large spectrum of
1537: quasi-periodic oscillations from each source. Such observations will
1538: constrain significantly the underlying physical model for these
1539: oscillations, which remains the most important source of uncertainty
1540: in using fast variability phenomena in probing strong gravitational
1541: fields.
1542:
1543:
1544: \newpage
1545:
1546: \section{The Need for a Theoretical Framework for Strong-Field Gravity
1547: Tests}
1548: \label{sec:need}
1549:
1550: Modern observations of black holes and neutron stars in the galaxy
1551: provide ample opportunities for testing the predictions of general
1552: relativity in the strong field regime, as discussed in the previous
1553: section. In several cases, astrophysical complications make such
1554: studies strongly dependent on model assumptions. This will be remedied
1555: in the near future, with the anticipated advances in the observational
1556: techniques and in the theoretical modeling of the various
1557: astrophysical phenomena. A second difficult hurdle, however, in
1558: performing quantitative tests of gravity with compact objects will be
1559: the lack of a parametric extension to General Relativity, i.e., the
1560: equivalent of the PPN formalism, that is suitable for calculations in
1561: the strong-field regime.
1562:
1563: In the past, {\em bona fide\/} tests of strong-field general
1564: relativity have been performed using particular parametric extensions
1565: to the Einstein--Hilbert action. This appears a priori to be a
1566: reasonable approach for a number of reasons. First, deriving the
1567: parametric field equations from a Lagrangian action ensures that
1568: fundamental symmetries and conservation laws are obeyed. Second, the
1569: parametric Lagrangian action can be used over the entire range of
1570: field strengths available to an observer and, therefore, even tests of
1571: General Relativity in the weak-field limit (i.e., with the PPN
1572: formalism) can be translated into constraints on the parameters of the
1573: action. This is often important, when strong-field tests lead to
1574: degenerate constraints between different parameters. Finally,
1575: phenomenological Lagrangian extensions can be motivated by ideas of
1576: quantum gravity and string theory and, potentially, help constrain the
1577: fundamental scales of such theories. There are, however, several
1578: issues that need to be settled before any such parametric extension of
1579: the Einstein--Hilbert action can become a useful theoretical framework
1580: for strong-field gravity tests (see also~\cite{Sotiriou08} and
1581: references therein).
1582:
1583: {\em First, gravity is highly non-linear and strong-field phenomena
1584: often show a non-perturbative dependence on small changes to the
1585: theory.~--\/} I will illustrate this with scalar-tensor theories that
1586: result from adding a minimally coupled scalar field to the Ricci
1587: curvature in the action. Such fields have been studied for more than
1588: 40 years in the form of Brans--Dicke gravity~\cite{Will93} and have
1589: been recently invoked as alternatives to a cosmological constant for
1590: modeling the acceleration of the universe~\cite{Peebles03}. In the
1591: context of compact-object astrophysics, constraints on the relative
1592: contribution of scalar fields coupled in different ways to the metric
1593: have been obtained from observations of the orbital decay of double
1594: neutron stars~\cite{Will89, Damour93} and compact X-ray
1595: binaries~\cite{Will89, Psaltis07a}. More recently, similar constraints
1596: on scalar extensions to General Relativity have been placed using the
1597: observation of redshifted lines from an X-ray burster~\cite{DeDeo03}
1598: and of quasi-periodic oscillations observed in accreting neutron
1599: stars~\cite{DeDeo07}. The oscillatory modes of neutron stars in such
1600: theories and the prospect of constraining them using gravitational
1601: wave signatures have also been studied~\cite{Sotani04, Sotani05}.
1602:
1603: The general form of the Lagrangian of a scalar-tensor theory is given,
1604: in the appropriate frame, by the Bregmann-Wagoner action
1605: (see~\cite{Will93} for details)
1606: %
1607: \begin{equation}
1608: S=\frac{1}{16\pi}\int d^4x \sqrt{-g_*}\left[R_* \pm
1609: g_*^{\mu\nu}\partial_\mu \phi \partial_\nu\phi +
1610: 2\lambda(\phi)\right]+S_{\rm m}[\phi_m, A^2(\phi)g_{*\mu\nu}]\;,
1611: \label{eq:st}
1612: \end{equation}
1613: %
1614: where $A(\phi)$ and $\lambda(\phi)$ are two arbitrary functions, and
1615: $S_m$ is the action for the matter field $\phi_m$. In the
1616: strong-field regime, the potential term $\lambda(\phi)$ in the
1617: action~(\ref{eq:st}) is typically negligible and is set to zero. On the
1618: other hand, the functional form of the coupling function $A(\phi)$ can
1619: be parametrized to measure deviations from General Relativity.
1620:
1621: Damour and Esposito-Farese~\cite{Damour93} considered a second-order
1622: parametric form
1623: %
1624: \begin{equation}
1625: A(\phi)= \exp\left[\alpha_0(\phi-\phi_0)+
1626: \frac{1}{2}\beta_0(\phi-\phi_0)^2+...\right]\;,
1627: \label{eq:stcoupling}
1628: \end{equation}
1629: %
1630: with $\phi_0\rightarrow 0$ a background cosmological value for the
1631: scalar field and $\alpha_0$ and $\beta_0$ the two parameters of the
1632: theory to be constrained by observations. The linear term,
1633: parametrized by $\alpha_0$, can be best constrained with weak-field
1634: tests. On the other hand, constraining significantly the non-linear
1635: term, parametrized by $\beta_0$, requires strong-field phenomena, such
1636: as those found around neutron stars. Indeed, the two main PPN
1637: parameters for such a scalar-tensor theory are
1638: %
1639: \begin{eqnarray}
1640: \gamma^{PPN}-1&=&-2\frac{\alpha_0^2}{1+\alpha_0^2}\nonumber\\
1641: \beta^{PPN}-1&=&\frac{\beta_0\alpha_0^2}{2(1+\alpha_0^2)^2}\;.
1642: \end{eqnarray}
1643: %
1644: The deviation of the PPN parameters from the general relativistic
1645: values is of second order in $\alpha_0$ and of third order in the
1646: product $\alpha_0^2\beta_0$. As a result, a very good limit on
1647: the parameter $\alpha_0$ renders the parameter $\beta_0$ practically
1648: unconstrainable by weak-field tests.
1649:
1650: The study of Damour and Esposito-Farese~\cite{Damour93} revealed one of
1651: the main reasons that necessitate careful theoretical studies of
1652: possible extensions of General Relativity that are suitable for
1653: strong-field tests. The order of a term added to the Lagrangian action
1654: of the gravitational field is not necessarily a good estimate of the
1655: expected magnitude of the observable effects introduced by this
1656: additional term. For example, because of the non-linear coupling
1657: between the scalar field and matter introduced by the coupling
1658: function~(\ref{eq:stcoupling}), the deviation from general
1659: relativistic predictions is not perturbative. For values of $\beta_0$
1660: less that about $-6$, it becomes energetically favorable for neutron
1661: stars to become ``scalarized'', with properties that differ
1662: significantly from their general relativistic
1663: counterparts~\cite{Damour93}. Such non-perturbative effects make
1664: quantitative tests of strong-field gravity possible even when the
1665: astrophysical complications are only marginally understood.
1666:
1667: A similar situation, albeit in the opposite regime, arises in an
1668: extended gravity theory in which a term proportional to the inverse of
1669: the Ricci scalar curvature, $R^{-1}$, is added to the Einstein--Hilbert
1670: action in order to explain the accelerated expansion of the
1671: universe~\cite{Carroll04}. Although one would expect that such an
1672: addition can only affect gravitational fields that are extremely
1673: weak, it turns out that it also alters to zeroth order the
1674: post-Newtonian parameter $\gamma$ and can, therefore, be excluded by
1675: simple solar-system tests~\cite{Chiba03}.
1676:
1677: {\em Second, Lagrangian extensions of General Relativity often suffer
1678: from serious problems with instabilities.~--\/} This issue can be
1679: understood by considering a Lagrangian action that includes terms of
1680: second order in the Ricci scalar, i.e., $R^2$, as well as the terms of
1681: similar order that can be constructed with the Ricci and Riemann
1682: tensors. For the sake of the argument, I will consider here the
1683: parametric Lagrangian
1684: %
1685: \begin{equation}
1686: {\cal S}= \frac{c^4}{16\pi G}\int d^4x
1687: \sqrt{-g}\left(R+\alpha_2 R^2 + \beta_2 R_{\sigma \tau}R^{\sigma \tau}
1688: + \gamma_2 R^{\alpha\beta\gamma\delta} R_{\alpha\beta\gamma\delta}\right)\;,
1689: \label{eq:PL}
1690: \end{equation}
1691: %
1692: with $\alpha_2$, $\beta_2$, and $\gamma_2$ the three parameters of the
1693: theory. Such terms arise naturally as high-order corrections in
1694: quantum gravity and string theory and their relative importance
1695: increases with the curvature of the metric~\cite{Donoghue94,
1696: Burgess04}. They have also been invoked as alternatives to the
1697: inflation paradigm for the early expansion of the
1698: universe~\cite{Starobinski80}. The predictions for astrophysical
1699: objects of extended gravity theories that incorporate high-order terms
1700: have been reported only for a few limited cases in the literature. The
1701: dependence of the stellar properties on $R^2$ terms in the action has
1702: been studied by Parker and Simon~\cite{Parker93}, who simply derived
1703: the generalized Tolmann--Oppenheimer--Volkoff equation without solving
1704: it, and by Barraco and Hamity~\cite{Barraco98} who attempted to solve
1705: the problem using a perturbation analysis (unfortunately, this last
1706: study suffers from a large number of errors).
1707:
1708: This second-order gravity theory has a number of unappealing
1709: properties (see discussion in~\cite{Simon90, Simon91}). Classically, a
1710: high-order gravity theory requires more than two boundary conditions,
1711: which is a fact that appears to be incompatible with all other
1712: physical theories. Quantum mechanically, second-order gravity theories
1713: lead to unstable vacuum solutions. Both these phenomena could be
1714: artifacts of the possibility that the action~(\ref{eq:PL}) may arise
1715: as a low-energy expansion of a non-local Lagrangian that is
1716: fundamentally of second order~\cite{Simon90, Simon91}.
1717: Phenomenologically speaking, these problems can be overcome by
1718: requiring the field equations to be of second order, when extremizing
1719: the action. This procedure leads to a generalized, high-order gravity
1720: theory that remains consistent with classical expectations and is
1721: stable quantum-mechanically (according to the procedure outlined
1722: in~\cite{Simon90, Simon91}), but requires a different than usual
1723: derivation of the field equations~\cite{DeDeo08}.
1724:
1725: Even when these issues are being taken into account, the terms
1726: proportional to $\beta_2$ and $\gamma_2$ lead to field equations with
1727: solutions that suffer from the Ostrogradski
1728: instability~\cite{Woodard06}. And even if these terms are dropped and
1729: only actions that are generic function of only the Ricci scalar are
1730: considered, then the resulting solutions for the expansion of the
1731: universe~\cite{Dolgov03} and for spherically symmetric
1732: stars~\cite{Seifert07} can be violently unstable, depending on the sign
1733: of the second-order term.
1734:
1735: A potential resolution of several of these problems in theories with
1736: high-order terms in the action appears to be offered by the Palatini
1737: formalism. In this approach, the field equations are derived by
1738: extremizing the action under variations in the metric and the
1739: connection, which is considered as an independent
1740: field~\cite{Sotiriou07}. For the simple Einstein--Hilbert action, both
1741: approaches are equivalent and give rise to the equations of general
1742: relativity; when the action has non-linear terms in $R$, the two
1743: approaches diverge. Unfortunately, the Palatini formalism leads to
1744: equations that cannot handle in general the transition across the
1745: surface layer of a star to the matter-free space outside it, and is
1746: therefore not a viable alternative~\cite{Barausse07}.
1747:
1748: {\em Finally, it is crucial that we identify the astrophysical
1749: phenomena that can be used in testing particular aspects of
1750: strong-field gravity.\/} For example, in the case of the
1751: classical tests of General Relativity, it is easy to show that the
1752: deflection of light during a solar eclipse and the Shapiro time delay
1753: depend on one (and the same) component of the metric of the Sun (i.e.,
1754: on the PPN parameter $\gamma$). Therefore, they do not provide
1755: independent tests of General Relativity (as long as we accept the
1756: validity of the equivalence principle). On the other hand, the
1757: perihelion precession of Mercury and the gravitational redshift depend
1758: on the other component of the metric (i.e., on the PPN parameter
1759: $\beta$) and, therefore, provide complementary tests of the
1760: theory. Understanding such degeneracies is an important component of
1761: performing tests of gravity theories.
1762:
1763: In the case of strong gravitational fields, this issue can be
1764: illustrated again by studying the high-order Lagrangian
1765: action~(\ref{eq:PL}) in the metric formalism (see
1766: also~\cite{Burgess04}). In principle, as the strength of the
1767: gravitational field increases, the terms that are of second-order in
1768: the Ricci scalar become more important and, therefore, affect the
1769: observable properties of neutron stars and black holes. However,
1770: because of the Gauss-Bonnet identity,
1771: %
1772: \begin{equation}
1773: \frac{\delta}{\delta g_{\mu\nu}}\int d^4x \left(
1774: R^2 -4 R_{\sigma \tau}R^{\sigma \tau}
1775: + R^{\alpha\beta\gamma\delta} R_{\alpha\beta\gamma\delta}\right)
1776: =0\;,
1777: \end{equation}
1778: %
1779: variations, with respect to the metric, of the term proportional to
1780: $\gamma_2$ in Equation~(\ref{eq:PL}) can be expressed as variations of
1781: the terms proportional to $\alpha_2$ and $\beta_2$. Therefore, for all
1782: non-quantum gravity tests, the predictions of the theory described by
1783: the Lagrangian action~(\ref{eq:PL}) are identical to those of the
1784: Lagrangian
1785: %
1786: \begin{equation}
1787: {\cal S}= \frac{c^4}{16\pi G}\int d^4x
1788: \sqrt{-g}\left[R+(\alpha_2-\gamma_2) R^2 + (\beta_2+4\gamma_2)
1789: R_{\sigma \tau}R^{\sigma \tau}\right]\;.
1790: \label{eq:r2_clas}
1791: \end{equation}
1792: %
1793: As a result, {\em astrophysical tests that do not invoke
1794: quantum-gravity effects can only constrain a particular combination of
1795: the parameters, i.e., $\alpha_2-\gamma_2$ and $\beta_2+4\gamma_2$.} It
1796: is only through phenomena related to quantum gravity, such as the
1797: evaporation of black holes, that the parameter $\gamma$ may
1798: be constrained.
1799:
1800: When the spacetime is isotropic and homogeneous, as in the case of
1801: tests using the cosmic evolution of the scale factor, an additional
1802: identity is satisfied, i.e.,
1803: %
1804: \begin{equation}
1805: \frac{\delta}{\delta g_{\mu\nu}}\int d^4x \left(
1806: R^2 -3 R_{\sigma \tau}R^{\sigma \tau}\right)=0\;.
1807: \end{equation}
1808: %
1809: This implies that, for cosmological tests, the predictions of the theory
1810: described by the Lagrangian action~(\ref{eq:PL}) are identical to those
1811: of the Lagrangian
1812: %
1813: \begin{equation}
1814: {\cal S}= \frac{c^4}{16\pi G}\int d^4x
1815: \sqrt{-g}\left[R+(\alpha_2+\frac{1}{3}\beta_2
1816: +\frac{1}{3}\gamma_2) R^2\right]\;.
1817: \label{eq:r2_cosm}
1818: \end{equation}
1819: %
1820: As a result, {\em such cosmological tests of gravity can only constrain a
1821: particular combination of the parameters, i.e.,
1822: $\alpha_2+\beta_2/3+\gamma_2/3$.}
1823:
1824: The parameters $\alpha_2$ and $\beta_2$ can be independently
1825: constrained using observations of spacetimes that are strongly curved
1826: but are not isotropic and homogeneous, such as those found in the
1827: vicinities of black holes and neutron stars. Measuring the properties
1828: of neutron stars, such as their radii, maximum masses and maximum
1829: spins, which require the solution of the field equations in the
1830: presence of matter, will provide independent constraints on the
1831: combination of parameters $\alpha_2-\gamma_2$ and $\beta_2+4\gamma_2$.
1832: However, one can show that in absence of matter, the external
1833: spacetime of a black hole, as given by the solution to Einstein's
1834: field equation, is also one (but not necessarily the only) solution of
1835: the parametric field equation that arises from the Lagrangian
1836: action~(\ref{eq:PL}). As a result, tests that involve black holes will
1837: probably be inadequate in distinguishing between the particular theory
1838: described by Equation~(\ref{eq:PL}) and general
1839: relativity~\cite{Psaltis07b}.
1840:
1841: This is, in fact, a general problem of using astrophysical
1842: observations of black holes to test General Relativity in the
1843: strong-field regime. The Kerr solution is not unique to general
1844: relativity~\cite{Psaltis07b}. For example, there is strong
1845: analytical~\cite{Thorne71, Bekenstein72, Hawking72} and numerical
1846: evidence~\cite{Scheel95} that, in Brans--Dicke scalar-tensor gravity
1847: theories, the end product of the collapse of a stellar configuration
1848: is a black hole described by the same Kerr solution as in Einstein's
1849: theory. The same appears to be true in several other theories
1850: generated by adding additional degrees of freedom to Einstein's
1851: gravity; the only vacuum solutions that are astrophysically relevant
1852: are those described by the Kerr metric~\cite{Psaltis07b}. Until a
1853: counter-example is discovered, studies of the strong gravitational
1854: fields found in the vicinities of black holes can be performed only
1855: within phenomenological frameworks, such as those involving multipole
1856: expansions of the Schwarzschild and Kerr metrics~\cite{Ryan95,
1857: Collins04, Glampedakis06}.
1858:
1859: To date, it has only been possible to test quantitatively the
1860: predictions of General Relativity in the strong-field regime using
1861: observations of neutron stars, as I will discuss in the following
1862: section. In all cases, the general relativistic predictions were
1863: contrasted to those of scalar-tensor gravity, with Einstein's theory
1864: passing all the tests with flying colors.
1865:
1866:
1867: \newpage
1868:
1869: \section{Current Tests of Strong-Field Gravity with Neutron Stars}
1870: \label{sec:tests}
1871:
1872: Performing tests of strong-field gravity with neutron stars requires
1873: knowledge of the equation of state of neutron-star matter to a degree
1874: better than the required precision of the gravitational test. This
1875: appears {\em a priori\/} to be a serious hurdle given the wide range
1876: of predictions of equally plausible theories of neutron-star matter
1877: (see~\cite{Lattimer01} for a recent compilation). It is easy to
1878: show, however, that current uncertainties in our modeling of the
1879: properties of ultra-dense matter do not preclude significant
1880: constraints on the strong-field behavior of gravity~\cite{DeDeo03}.
1881:
1882: \epubtkImage{ns.png}{
1883: \begin{figure}[htbp]
1884: \centerline{\includegraphics[height=10.0cm]{ns}}
1885: \caption{Mass-radius relations of neutron stars
1886: in General Relativity (GR), scalar-tensor (ST), and Rosen's bimetric
1887: theory of gravity~\cite{DeDeo03}. The shaded areas represent the range
1888: of mass-radius relations predicted in each case by neutron-star
1889: equations of state without unconfined quarks or condensates. All
1890: gravity theories shown in the figure are consistent with solar-system
1891: tests but introduce variations in the predicted sizes of neutron stars
1892: that are significantly larger than the uncertainty caused by the
1893: unknown equation of state.}
1894: \label{fig:mr}
1895: \end{figure}}
1896:
1897: During the last three decades, neutron-star models have been
1898: calculated for a variety of gravity theories (see~\cite{Will93} and
1899: references therein) and were invariably different, both in size and in
1900: allowed mass, than their general relativistic counterparts. As an
1901: example, Figure~\ref{fig:mr} shows neutron-star models calculated in
1902: three representative theories that cannot be excluded by current tests
1903: that do not involve neutron stars. In the figure, the shaded areas
1904: represent the uncertainty introduced by the unknown equation of state
1905: of neutron-star matter (not including quark stars or large neutron
1906: stars with condensates). Clearly, the deviations in neutron-star
1907: properties from the predictions of General Relativity for these
1908: theories (that are still consistent with weak-field tests) are larger
1909: than the uncertainty introduced by the unknown equation of state of
1910: neutron-star matter.
1911:
1912: This is a direct consequence of the fact that the curvature around a
1913: neutron star is larger by $\sim 13$ orders of magnitude compared to
1914: the curvature probed by solar-system tests, whereas the density inside
1915: the neutron star is larger by only an order of magnitude compared to
1916: the densities probed by nuclear scattering data that are used to
1917: constrain the equation of state. Given that the current values of the
1918: post-Newtonian parameters are known from weak-field tests to within
1919: $\sim 10^{-4}$, it is reasonable that deviations from general
1920: relativity can be hidden in the weak-field limit but may become
1921: dominant as the curvature is increased by more than ten orders of
1922: magnitude. Neutron stars can indeed be used in testing the
1923: strong-field behavior of a gravity theory.
1924:
1925: \epubtkImage{xpsr_bd.png}{
1926: \begin{figure}[htbp]
1927: \centerline{\includegraphics[height=10.0cm]{xpsr_bd}}
1928: \caption{The limiting rate for the evolution of the
1929: orbital periods ($\tau_{\rm P}^{-1}\equiv \dot{P}/P$) of five known
1930: millisecond accreting pulsars as a function of the Brans--Dicke
1931: parameter $\omega_{\rm BD}$. The lower half of the plot corresponds
1932: to an orbital period that decreases with time ($\dot{P}/P<0$), whereas
1933: the upper half corresponds to an orbital period that increases with
1934: time ($\dot{P}/P>0$). Only the area outside the two curves for each
1935: system is physically allowed~\cite{Psaltis07a}.}
1936: \label{fig:xpsr_bd}
1937: \end{figure}}
1938:
1939:
1940: \subsection{Brans--Dicke gravity and the orbital decay of binary systems
1941: with neutron stars}
1942:
1943: Binary stellar systems that are currently known to harbor at least one
1944: neutron star have orbital separations that are too large to be used in
1945: probing directly strong gravitational fields. Even at that separation,
1946: however, the orbital evolution of the binary system caused by the
1947: emission of gravitational waves is affected, in a scalar-tensor
1948: theory, by the coupling of matter to the scalar field, which occurs in
1949: a strong gravitational field. This manifests itself as a violation of
1950: the strong equivalence principle, with many observable consequences
1951: such as the rapid decay of the orbit due to emission of dipole
1952: radiation~\cite{Eardley75, Will89}. The various quantitative tests of
1953: strong-field gravity using binary systems with radio pulsars have been
1954: reviewed in detail elsewhere~\cite{Stairs03}. Here, I will focus only
1955: on tests that involve the orbital period evolution of the binary
1956: systems.
1957:
1958: The best studied binaries with compact objects are the double neutron
1959: stars, with the Hulse--Taylor pulsar (PSR 1913$+$16) as the
1960: prototypical case. Unfortunately, in all double neutron-star systems,
1961: the masses of the two members of the binary are surprisingly
1962: similar~\cite{Thorsett99} and this severely limits the prospects of
1963: placing strong constraints on the dipole radiation from them. Indeed,
1964: the magnitude of dipole radiation depends on the difference of the
1965: sensitivities between the two members of the binaries, and for neutron
1966: stars the sensitivities depend primarily on their masses. The
1967: resulting constraint imposed on the Brans--Dicke parameter $\omega$ by
1968: the Hulse--Taylor pulsar is significantly smaller than the limit
1969: $\omega>40000$ set by the Cassini mission~\cite{Bertotti03}.
1970:
1971: The constraint is significantly improved when studying binary systems
1972: in which only one of the two stars is a neutron star. There are
1973: several known neutron star-white dwarf binaries that are suitable for
1974: this purpose, in which the neutron stars appear as radio pulsars
1975: (e.g., PSR B0655$+$64~\cite{Damour96};
1976: PSR~J0437$-$4715~\cite{vanStraten01}), as millisecond accreting X-ray
1977: pulsars (e.g., XTE~J1808$-$456~\cite{Psaltis07a}), or as non-pulsing
1978: X-ray sources (e.g., 4U~1820$-$30~\cite{Will89}). In the last two
1979: cases, the evolution of the binary orbit is also affected
1980: significantly by mass transfer from the companion star to the neutron
1981: star. However, for each value of the Brans--Dicke parameter
1982: $\omega_{\rm BD}$, there is a minimum absolute value for the rate of
1983: evolution of the orbital period (see Figure~\ref{fig:xpsr_bd}
1984: and~\cite{Psaltis07a}). An accurate measurement of the orbital period
1985: derivative in any of these systems offers, therefore, the potential of
1986: placing a lower limit on the Brans--Dicke parameter. Because of the
1987: astrophysical complications introduced by mass transfer, the optimal
1988: constraint on $\omega_{\rm BD}$ is of order $10^4$ in this method,
1989: which is comparable to the Cassini limit.
1990:
1991: \epubtkImage{st2_psr.png}{
1992: \begin{figure}[htbp]
1993: \centerline{\includegraphics[height=10.0cm]{st2_psr}}
1994: \caption{Constraints on the two parameters of a second-order
1995: scalar-tensor theory placed by the timing properties of a number of
1996: binary stellar systems that harbor neutron stars. For two of the
1997: systems, current constraints are contrasted to those expected in the
1998: near future when a measurement of the orbital period derivative is
1999: possible to an accuracy of 1\%. General relativity corresponds to the
2000: origin of the parameter space; the constraint imposed by the Cassini
2001: mission is also shown for comparison~\cite{Damour07}. In all cases,
2002: the allowed part of the parameter space is under the corresponding
2003: curve.}
2004: \label{fig:st_psr}
2005: \end{figure}}
2006:
2007:
2008: \subsection{Second-order scalar-tensor gravity and radio pulsars}
2009:
2010: As discussed in the previous section, observations of strong-field
2011: phenomena provide constraints on Brans--Dicke scalar-tensor gravity,
2012: which are, however, at most comparable to those of solar system tests.
2013: This is true because the fractional deviation of a Brans--Dicke theory
2014: from General Relativity is of order $\omega_{\rm BD}^{-1}$, both for
2015: weak and strong gravitational fields, and the solar-system tests have
2016: superb accuracy. On the other hand, a scalar-tensor theory with a
2017: second-order coupling (e.g., the one arising from the
2018: action~(\ref{eq:st}) with the coupling~(\ref{eq:stcoupling})) allows
2019: for large deviations in the strong-field regime while being consistent
2020: with the weak-field limits~\cite{Damour93, Damour96}.
2021:
2022: In the case of neutron stars, the second-order scalar-tensor theory
2023: described by Damour and Esposito-Farese~\cite{Damour93} leads to a
2024: non-perturbative effect known as spontaneous scalarization (similar to
2025: the spontaneous magnetization in ferromagnetism). For significantly
2026: large negative values of the parameter $\beta_0$, there is a range of
2027: neutron-star masses for which it becomes energetically favorable for
2028: the scalar field to acquire high values inside the neutron star and
2029: affect significantly its structure compared to the general
2030: relativistic predictions. An example of the mass-radius relation for
2031: neutron stars in a second-order scalar-tensor theory with $\beta_0=-8$
2032: is shown in Figure~\ref{fig:mr}.
2033:
2034: The properties and stability of scalarized neutron stars have been
2035: studied extensively in the literature~\cite{Damour93, Salgado98,
2036: Harada98}. For the purposes of tests of strong-field gravity, the
2037: coupling of matter with the gravitational field and the external
2038: spacetimes of scalar stars are so different compared to their general
2039: relativistic counterparts that large negative values of $\beta_0$ can
2040: be firmly excluded with current observations of binary stellar systems
2041: that harbor radio pulsars. Figure~{\ref{fig:st_psr}} shows the current
2042: constraints on the two parameters $\alpha_0$ and $\beta_0$ of the
2043: theory imposed by the timing observations of the Hulse--Taylor pulsar
2044: (PSR~J1913$+$16), of a pulsar in an asymmetric binary with a white
2045: dwarf (PSR~J1141$-$6545), and of two other pulsars (PSR~J0737$-$3079
2046: and PSR~B1534$+$12). The best weak-field limits, including those
2047: imposed by the Cassini mission, are also shown for
2048: comparison~\cite{Damour07}.
2049:
2050: As expected, weak-field tests bound significantly the value of the
2051: parameter $\alpha_0$, leaving the parameter $\beta_0$ largely
2052: unconstrained. Between the binary systems with radio pulsars, the one
2053: with the white-dwarf companion provides the most stringent
2054: constraints because the large asymmetry between the two compact object
2055: leads to the prediction of strong dipole gravitational radiation that
2056: can be excluded observationally. Finally, for large negative values
2057: of the parameter $\beta_0$, the scalarization of the neutron stars makes
2058: the predictions of the theory incompatible with observations.
2059:
2060: \epubtkImage{redshift.png}{
2061: \begin{figure}[htbp]
2062: \centerline{\includegraphics[height=10.0cm]{redshift}}
2063: \caption{Contours of constant
2064: gravitational redshift measured at infinity for an atomic line
2065: originating at the surface of a neutron star in a scalar-tensor
2066: gravity theory, for different values of the parameter $\beta_0$ that
2067: measures the relative contribution of the scalar field. The thick
2068: curve separates the scalarized stars from the general relativistic
2069: counterparts. The measurements of a redshift of $z=0.35$ from a
2070: burster~\cite{Cottam02} and the astrophysical constraint of a baryonic
2071: mass of at least $1.4\,M_\odot$ (dashed lines) result in a bound on the
2072: parameter $\beta$ of $-\beta<9$~\cite{DeDeo03}. }
2073: \label{fig:z}
2074: \end{figure}}
2075:
2076:
2077: \subsection{Second-order scalar-tensor gravity and X-ray observations
2078: of accreting neutron stars}
2079:
2080: The quantitative features of a number of phenomena observed in the
2081: X-rays from accreting neutron stars depend strongly on their masses
2082: and radii, as discussed in~\ref{sec:strong}. The constraints imposed by two of
2083: these phenomena on the parameters of the second-order scalar-tensor
2084: gravity of Damour and Esposito-Farese~\cite{Damour93} have been studied
2085: recently~\cite{DeDeo03, DeDeo07}.
2086:
2087: The first phenomenon is the observation of gravitationally redshifted
2088: atomic lines during X-ray bursts from the source
2089: EXO~0748$-$56~\cite{Cottam02}. Figure~\ref{fig:z} shows the values of
2090: the gravitational redshift from the surface of neutron stars with
2091: different masses, in second-order scalar-tensor theories with
2092: different values of the parameter $\beta_0$~\cite{DeDeo03}. In this
2093: calculation, the parameter $\alpha_0$ was set to zero and the
2094: neutron-star structure was calculated using the equation of state
2095: U~\cite{Cook94}. The hatch-filled area corresponds to neutron-star
2096: masses that are unacceptable for each value of the parameter
2097: $\beta_0$, while the thick curve separates the scalarized stars from
2098: the general relativistic counterparts.
2099:
2100: A dynamical measurement of the mass of EXO~0748$-$56 can rule out the
2101: possibility that the neutron star in this source is scalarized,
2102: because scalarized stars have very different surface redshifts
2103: compared to the general relativistic stars of the same mass. The
2104: source EXO~0748$-$56 lies in an eclipsing binary system which makes it
2105: a prime candidate for a dynamical mass measurement. In the absence of
2106: such a measurement, however, a limit on the parameter $\beta_0$ can be
2107: placed under the astrophysical constraint that the baryonic mass of
2108: the neutron stars is larger than $\simeq 1.4\,M_\odot$. This is a
2109: reasonable assumption, given that a progenitor core of a lower mass
2110: would not have collapsed to form a neutron star. Combining this
2111: constraint with the measured redshift of $z=0.35$ leads to a limit on
2112: the parameter $-\beta_0<9$, which depends only weakly on the assumed
2113: equation of state of neutron-star matter~\cite{DeDeo03}.
2114:
2115: \epubtkImage{maxorb.png}{
2116: \begin{figure}[htbp]
2117: \centerline{\includegraphics[height=10.0cm]{maxorb}}
2118: \caption{The maximum orbital frequency outside a neutron star
2119: of mass $M_{\rm ADM}$ for different scalar-tensor theories identified
2120: by the parameter $\beta$. The dashed line shows the maximum observed
2121: frequency of a quasi-periodic oscillation from an accreting neutron
2122: star~\cite{DeDeo07}.}
2123: \label{fig:QPO}
2124: \end{figure}}
2125:
2126: A second set of phenomena that can lead to strong-field tests of
2127: gravity are the fast quasi-periodic oscillations observed from many
2128: bright accreting neutron stars~\cite{vanderKlis06}. The highest known
2129: frequency of such an oscillations is 1330~Hz, observed from the source
2130: 4U~1636$-$53 and corresponds to the Keplerian frequency of the
2131: innermost stable circular orbit of a $1.6\,M_\odot$ slowly spinning
2132: neutron star. Figure~\ref{fig:QPO} shows the maximum Keplerian
2133: frequency outside a neutron star in the second-order scalar tensor
2134: theory, for different values of the parameter $\beta_0$. For small
2135: stellar masses, the limiting frequency is achieved at the surface of
2136: the star, whereas for large stellar masses, the limiting frequency is
2137: reached at the innermost stable circular orbit. This figure shows that
2138: scalarized stars allow for higher frequencies than their general
2139: relativistic counterparts. Requiring, therefore, the observed
2140: oscillation frequency to be smaller than the highest Keplerian
2141: frequency of a stable orbit outside the compact object cannot be used
2142: to constrain the parameters of this theory. On the other hand, the
2143: correlations between the various dynamical frequencies outside the
2144: compact object depend strongly on the parameter $\beta$ and hence the
2145: gravity theory can be constrained given a particular model for the
2146: oscillations~\cite{DeDeo07}.
2147:
2148:
2149: \newpage
2150:
2151: \section{Going Beyond Einstein}
2152: \label{section:beyond}
2153:
2154: \epubtkImage{tests_future.png}{
2155: \begin{figure}[htbp]
2156: \centerline{\includegraphics[width=12.0cm]{tests_future}}
2157: \caption{The spectral (redshift) and timing capabilities
2158: required for an observatory to probe different strengths of
2159: gravitational fields. Phenomena that occur in the vicinities
2160: of neutron stars and stellar-mass black holes experience large
2161: redshift and occur over sub-millisecond timescales.}
2162: \label{fig:tests_future}
2163: \end{figure}}
2164:
2165: Testing General Relativity in the strong-field regime with neutron
2166: stars and black holes will require advanced observatories that will be
2167: able to resolve various phenomena in the characteristic energy and
2168: time-scales in which they occur. The two parameters used to quantify
2169: the strength of a gravitational field in
2170: Section~\ref{subsec:definition} are also useful in discussion the
2171: specifications required by such future observatories.
2172:
2173: The potential and the curvature in a gravitational field are related
2174: directly to the characteristic energy- and time-scales, respectively,
2175: that need to be resolved in order for an observation to be able to
2176: probe a particular region of the parameter space. The potential
2177: $\epsilon$ gives directly the gravitational redshift $z$ according to
2178: %
2179: \begin{equation}
2180: z=1-(1-2\epsilon)^{-1/2}\;,
2181: \end{equation}
2182: %
2183: the measurement of which is the goal of spectroscopic observations;
2184: for weak gravitational fields $z\simeq\epsilon$. At the same time, the
2185: curvature $\xi$ is directly related to the dynamical timescale $\tau$
2186: in the same region of a gravitational field by
2187: %
2188: \begin{equation}
2189: \tau=\frac{2\pi}{c}\xi^{-1/2}\;.
2190: \end{equation}
2191: %
2192: As shown in Figure~\ref{fig:tests_future}, only observatories with
2193: excellent spectroscopic and millisecond timing capabilities will be
2194: able to resolve phenomena that occur in the strongest gravitational
2195: fields found in astrophysics, i.e., those in the vicinities of neutron
2196: stars and stellar-mass black holes.
2197:
2198: \epubtkImage{gwaves.png}{
2199: \begin{figure}[htbp]
2200: \centerline{\includegraphics[width=12.0cm]{gwaves}}
2201: \caption{The parameter space that will be probed by
2202: an experiment based on a gravitational wave detection with LIGO and
2203: LISA, for an assumed source at a distance of 1~Mpc.}
2204: \label{fig:gwaves}
2205: \end{figure}}
2206:
2207: One of the most promising avenues towards testing strong-field general
2208: relativity is via the detection of the gravitational waves emitted
2209: during the coalescence of compact objects. In the simple case in which
2210: two compact objects of mass $M$ are orbiting each other in circular
2211: orbits with separation $a$, slowly approaching because of the emission
2212: of gravitational waves, the characteristic period $P$ of the
2213: gravitational wave is half of the orbital period and, therefore, is
2214: related to the spacetime curvature by
2215: %
2216: \begin{equation}
2217: P=\frac{\pi}{c}\xi^{-1/2}\;.
2218: \label{eq:Pgw}
2219: \end{equation}
2220: %
2221: At the same time, the strain $h$ detected by an observatory on Earth
2222: for a gravitational wave emitted by such a source placed at a distance
2223: $D$, is~\cite{Flanagan05}
2224: %
2225: \begin{equation}
2226: h=\left(\frac{GM}{ac^2}\right)\left(\frac{GM}{Dc^2}\right)\;.
2227: \end{equation}
2228: %
2229: Given the distance to a source and the measurement of a strain, the
2230: curvature of the gravitational field probed is
2231: %
2232: \begin{equation}
2233: \xi=\frac{\epsilon^5}{h^2 D^2}=10^{-3}\epsilon^5
2234: \left(\frac{h}{10^{-23}}\right)^{-2}
2235: \left(\frac{D}{1~\mbox{Mpc}}\right)^{-2}~\mbox{cm}^{-2}\;.
2236: \label{eq:xigw}
2237: \end{equation}
2238: %
2239: The sensitivity of each detector of gravitational waves depends
2240: strongly on the period of the wave. Using equations~(\ref{eq:Pgw})
2241: and~(\ref{eq:xigw}), the sensitivity curve of a detector can be
2242: converted into a region of the parameter space that can be probed,
2243: given the distance to the source. This is shown in
2244: Figure~\ref{fig:gwaves} for the advanced LIGO and LISA, for an assumed
2245: source distance of 1~Mpc. Gravitational waves detected by LISA will
2246: probe the same curvatures as current tests of General Relativity but
2247: significantly larger potentials. On the other hand, gravitational
2248: waves detected by the advanced LIGO have the potential of probing
2249: directly the strongest gravitational fields found around astrophysical
2250: objects.
2251:
2252: In the near future, a number of observatories will exploit new
2253: techniques and open new horizons in gravitational physics by exploring
2254: the strong-field region of the parameter space shown in
2255: Figure~\ref{fig:tests_future}. Observations with the Square Kilometre
2256: Array~\cite{SKA} may lead to the discovery of the most optimal binary
2257: systems for strong-field gravity tests with pulsar timing, in which a
2258: pulsar is orbiting a black hole~\cite{Kramer05}. High energy
2259: observations of black holes and neutron stars with
2260: Constellation-X~\cite{ConX} and XEUS~\cite{XEUS} will detect highly
2261: redshifted atomic lines and measure their rapid variability
2262: properties. Finally, gravitational wave observatories, either from the
2263: ground (such as LIGO~\cite{LIGO}, GEO600~\cite{GEO},
2264: TAMA300~\cite{TAMA}, and VIRGO~\cite{VIRGO}) or from space (such as
2265: LISA~\cite{LISA}) will detect directly for the first time one of the
2266: most remarkable predictions of General Relativity, the generation of
2267: gravitational waves from orbiting compact objects and black hole
2268: ringing.
2269:
2270:
2271: \newpage
2272:
2273: \section{Acknowledgements}
2274: \label{acknowledgements}
2275:
2276: It is my great pleasure to acknowledge the many fruitful discussions
2277: and collaborations with a number of people that have shaped my ideas
2278: on astrophysical tests of strong-field gravity. In particular, I thank
2279: T.\ Belloni, D.\ Chakrabarty, S.\ DeDeo, F.\ Lamb, C.\ Miller, R.\
2280: Narayan, J.\ McClintock, F.\ \"Ozel, and M.\ van der Klis. I am
2281: indebted to S.\ DeDeo and F.\ \"Ozel for helping me settle on and
2282: understand the defition of strong-field gravity. I am also grateful to
2283: G.\ Esposito-Far\'ese, T.\ Johanssen, J.\ McClintock, F.\ \"Ozel, and
2284: C.\ Reynolds for their detailed comments that helped me greatly
2285: improve the presentation of this review.
2286:
2287:
2288: \newpage
2289:
2290: \bibliography{ms}
2291:
2292: \end{document}
2293: