0704.2612/ms.tex
1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: %%%%%%%%%%%%%%%%%%%%  FORMAT FOR LATEX PAPERS  %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4: %\documentclass[12pt,preprint]{aastex}
5: \documentclass[]{emulateapj}
6: 
7: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
8: %%%%%%%%%%%%%%%%%%%%  PACKAGES TO BE LOADED  %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
9: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
10: \usepackage{natbib}
11: \usepackage{amsfonts}
12: \usepackage{amsmath}
13: \usepackage{amssymb}
14: \usepackage{bm}
15: \usepackage{graphicx}
16: \usepackage[latin1]{inputenc}
17: \usepackage{latexsym}
18: \usepackage{rotating}
19: \usepackage{bm}             % bold math
20: \usepackage{amsfonts}       % AMS fonts
21: 
22: 
23: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
24: %%%%%%%%%%%%%%%%%%%%  USER DEFINED COMMANDS  %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
25: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
26: \newcommand{\mb}[1]{\mbox{\boldmath $#1$}}
27: \def \nn   {\nonumber}
28: \def \met  {\mbox{g}}
29: \def \metb {\mbox{\bf g}}
30: 
31: \newcommand\be{\begin{equation}}
32: \newcommand\ba{\begin{eqnarray}}
33: \newcommand\ee{\end{equation}}
34: \newcommand\ea{\end{eqnarray}}
35: 
36: \newcommand{\note}[1]{[$\blacktriangleright$~\textbf{#1}~$\blacktriangleleft$]}
37: \newcommand{\cut}{{\textrm{cut}}}
38: \newcommand{\orb}{{\textrm{orb}}}
39: \newcommand{\rr}{{\textrm{rr}}}
40: \newcommand{\Msun}{\textrm{M}_{\odot}}                   
41: \newcommand{\Mbh}{M_{\bullet}}
42: \newcommand{\abh}{a_{\bullet}}
43: \newcommand{\Sbh}{S_{\bullet}}
44: \newcommand{\Mmw}{M_{\mbox{\tiny MW}}}
45: \newcommand{\Rbh}{R_{\bullet}}
46: \newcommand{\Mco}{m}
47: 
48: 
49: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
50: %%%%%%%%%%%%%%%%%%%%  BEGIN DOCUMENT  %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
51: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
52: \begin{document}
53: 
54: 
55: 
56: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
57: %%%%%%%%%%%%%%%%%%%%  TITLE  %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
58: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
59: \title{Relativistic Effects in Extreme Mass Ratio Gravitational Wave Bursts}
60: 
61: 
62: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
63: %%%%%%%%%%%%%%%%%%%%  AUTHORS  %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
64: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
65: 
66: \author{Nicol\'as Yunes, Carlos F.  Sopuerta~\altaffilmark{1,2,3},
67:   Louis J. Rubbo~\altaffilmark{4}, Kelly
68:   Holley-Bockelmann~\altaffilmark{5}}
69: 
70: \affil{Institute for Gravitational Physics and Geometry and
71:   Center for Gravitational Wave Physics, The Pennsylvania State
72:   University, University Park, PA 16802, USA.}
73: 
74: \altaffiltext{1}{Department of Physics, University of Guelph,
75:   Guelph, Ontario, Canada N1G 2W1.}
76: 
77: \altaffiltext{2}{Institut de Ci\`encies de l'Espai (CSIC),
78:   Facultat de Ci\`encies, Campus UAB, Torre C5 parells, E-08193
79:   Bellaterra (Barcelona), Spain.}
80: 
81: \altaffiltext{3}{Institut d'Estudis Espacials de Catalunya
82:   (IEEC), Ed. Nexus-201, c/ Gran Capit\`a 2-4, E-08034 Barcelona,
83:   Spain.}
84: 
85: \altaffiltext{4}{Department of Physics, Coastal Carolina University
86:   Conway, SC 29528, USA.}
87: 
88: \altaffiltext{5}{Department of Physics and Astronomy, Vanderbilt
89:   University, Nashville, TN 37235, USA.}
90: 
91: 
92: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
93: %%%%%%%%%%%%%%%%%%%  ABSTRACT  %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
94: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
95: 
96: \begin{abstract}
97:   Extreme mass ratio bursts (EMRBs) have been proposed as a possible
98:   source for future space-borne gravitational wave detectors, such as
99:   the Laser Interferometer Space Antenna (LISA).  These events are
100:   characterized by long-period, nearly-radial orbits of compact
101:   objects around a central massive black hole. The gravitational
102:   radiation emitted during such events consists of a short burst,
103:   corresponding to periapse passage, followed by a longer, silent
104:   interval. In this paper we investigate the impact of including
105:   relativistic corrections to the description of the compact object's
106:   trajectory via a geodesic treatment, as well as including
107:   higher-order multipole corrections in the waveform calculation.  The
108:   degree to which the relativistic corrections are important depends
109:   on the EMRB's orbital parameters.  We find that relativistic EMRBs
110:   ($v_{\mathrm{max}}/c>0.25$) are not rare and actually account for
111:   approximately half of the events in our astrophysical model.  The
112:   relativistic corrections tend to significantly change the waveform
113:   amplitude and phase relative to a Newtonian description, although
114:   some of this dephasing could be mimicked by parameter errors.  The
115:   dephasing over several bursts could be of particular importance not
116:   only to gravitational wave detection, but also to parameter
117:   estimation, since it is highly correlated to the spin of the massive
118:   black hole.  Consequently, we postulate that if a relativistic EMRB
119:   is detected, such dephasing might be used to probe the relativistic
120:   character of the massive black hole and obtain information about its
121:   spin.
122: \end{abstract}
123: 
124: 
125: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
126: %%%%%%%%%%%%%%%%%%%  KEYWORDS  %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
127: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
128: 
129: \keywords{black hole physics --- Galaxy: nucleus --- gravitational
130:   waves --- stellar dynamics}
131: 
132: 
133: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
134: %%%%%%%%%%%%%%%%%%%  MAKETITLE  %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
135: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
136: 
137: \maketitle
138: 
139: 
140: 
141: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
142: %%%%%%%%%%%%%%%%%%%  INTRODUCTION  %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
143: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
144: 
145: \section{Introduction} \label{intro}
146: 
147: Low-frequency ($10^{-5} \lesssim f \lesssim 0.1$~Hz) gravitational
148: wave interferometers, such as the proposed Laser Interferometer Space
149: Antenna (LISA) \citep{Bender:1998,Danzmann:2003tv,Sumner:2004}, will
150: open a completely new window to the Universe. Through observations of
151: low-frequency gravitational radiation we will be able to witness the
152: inspiral and merger of massive black hole binaries; the inspiral of
153: compact objects into massive black holes; and millions of
154: quasi-stationary compact galactic binaries.  Recently, a new source of
155: low-frequency gravitational radiation has been suggested: extreme mass
156: ratio bursts (EMRBs) \citep{Rubbo:2006dv,Rubbo:2007}.
157: 
158: EMRBs consist of a stellar-mass compact object (SCO) orbiting a
159: massive black hole (MBH) of $10^{4-8}~\Msun$ with orbital periods
160: greater than $T_{\cut} = 3 \times 10^4$~s.  The defining orbital
161: period cutoff is derived from LISA's lower frequency limit of
162: $f_{\cut} = 3 \times 10^{-5}$~Hz.  Systems with orbital periods less
163: than $T_{\cut}$ will radiate continuously inside the LISA band.  Such
164: continuous systems are more appropriately categorized as extreme mass
165: ratio inspirals (EMRIs) and have been studied extensively elsewhere:
166: recent estimations of the event rate are given by \citet{Gair:2004iv}
167: and \citet{Hopman:2006ha}; a discussion on a possible EMRI background
168: is given by \citet{Barack:2004b}; accounts on the theoretical
169: description of EMRIs can be found in the reviews by
170: \citet{Poisson:2004lr} and \citet{Glampedakis:2005hs}, while the
171: recent review by \citet{Amaro-Seoane:2007aw} describes the
172: astrophysical and detection applications.
173: 
174: Although EMRB events are distinct from EMRI events, their evolutionary
175: track could be connected. In the burst scenario, the SCO orbits the
176: MBH emitting a beamed burst of gravitational radiation during
177: pericenter passage.  The emitted radiation carries away energy and
178: angular momentum from the system so that after multiple pericenter
179: passages the orbital period decreases, and possibly the system becomes
180: an EMRI.  However, this evolutionary track is most likely disrupted by
181: scattering interactions with other stars and/or if the SCO plunges
182: directly into the central MBH on one of its passages.
183: 
184: The EMRB event rate has recently been investigated using simplified
185: galactic models and data analysis techniques
186: \citep{Rubbo:2006dv,Rubbo:2007,Hopman:2006fc}.  Using a density
187: profile described by an $\eta$-model \citep{Tremaine:1994},
188: \citet{Rubbo:2006dv,Rubbo:2007} suggested an event rate of
189: $\sim\!15~{\textrm{yr}}^{-1}$ for events with signal-to-noise ratios
190: (SNRs) greater than five out to the Virgo cluster.  When mass
191: segregation and different inner cusp models are considered, the
192: predicted rate decreases by an order of magnitude
193: \citep{Hopman:2006fc}.  These preliminary studies were aimed at
194: understanding if EMRB event rates are interesting for low-frequency
195: gravitational wave detectors such as LISA.  More work is still needed
196: to improve the predicted event rate in the context of realistic
197: galaxies, where the role of non-equilibrium dynamics, anisotropy,
198: complex star formation histories, substructure, and non-sphericity may
199: act to change the rate from these fiducial estimates by orders of
200: magnitude \citep{HB:2006af, Rubbo:2006dv, Rubbo:2007}.
201: 
202: In addition to the astrophysical uncertainties, there are no
203: investigations of the impact of relativistic corrections to EMRB
204: dynamics.  All EMRB studies have been carried out in a
205: \textit{quasi-Newtonian} framework. In this framework, one uses the
206: Newtonian equations of motion and extracts the gravitational waveforms
207: by means of the {\em quadrupole formula}.  This approximation ignores
208: the black hole nature of the central potential, including the black
209: hole's rotation (spin), and is technically valid only for orbits with
210: non-relativistic velocities.  However, a considerable number of EMRBs
211: are characterized by large pericenter velocities ($v_{p} \gtrsim
212: 0.25\,c$) and these \textit{relativistic} EMRBs should produce
213: gravitational wave signals with larger SNRs, as we will show later.
214: 
215: In this paper, we shall not consider the issue of EMRB event rates,
216: but instead we shall study the effects of relativistic corrections to
217: such events.  For extreme-mass-ratio systems a simple way of
218: introducing relativistic corrections is by using the so-called
219: \textit{semi-relativistic} approximation introduced by
220: \citet{Ruffini:1981rs}, and used recently in the context of EMRIs by
221: \citet{Gair:2005is,Gair:2005hu}.  In this approximation, the MBH and
222: surrounding area are modeled using the Kerr solution to Einstein's
223: field equations, which describes a stationary spinning black hole (the
224: Schwarzschild solution corresponds to the non-spinning case).  The SCO
225: is considered to be a point-like object (neglecting its own
226: self-gravity) whose trajectory is described by a geodesic of the Kerr
227: spacetime.  In other words, relative to previous work, we have
228: replaced the Newtonian equations of motion by relativistic geodesic
229: equations of motion.
230: 
231: The relativistic description introduces effects such as orbital
232: precession and frame dragging, but it does not account for effects due
233: to the gravitational field induced by the SCO. These effects, for
234: example, lead to changes in the (geodesic) constants of motion due to
235: radiation reaction. Even though these effects introduce errors that
236: scale with the system's mass ratio \citep[e.g.
237: see][]{Glampedakis:2005hs}, they cannot be neglected for EMRIs. This
238: is because in the late stages of the EMRI the SCO spends a substantial
239: fraction of cycles in the strong-field region of the MBH.  On the
240: other hand, in the case of EMRBs, the SCO sling-shots around the MBH
241: and its interaction time during pericenter passage is relatively small
242: ($< 10^{5}$~s).  Radiation reaction effects can then be neglected
243: in EMRBs since the radiation reaction timescale is always much larger
244: than the period of pericenter passage.
245: 
246: In this paper, we also improve on the semi-relativistic approximation
247: by using a more precise gravitational wave extraction procedure. The
248: procedure employed is the multipole-moment wave generation formalism
249: for slow-motion objects with arbitrarily strong internal gravity
250: \citep{Thorne:1980rm}.  We consider terms up to the mass-octopole and
251: current-quadrupole multipoles, thus improving on the mass-quadrupole
252: analysis of \citet{Rubbo:2006dv,Rubbo:2007} and \citet{Hopman:2006fc}.
253: Higher multipoles will become important if the system becomes even
254: more relativistic, but pericenter velocities for EMRBs are typically
255: small to moderate relative to the speed of light (typically $0.1
256: \lesssim v_{p}/c \lesssim 0.5$).  Such higher multipolar corrections
257: were taken into account for EMRIs by~\citet{Babak:2006uv}, but for
258: those sources the phase evolution must be tracked very accurately,
259: requiring techniques from black hole perturbation 
260: theory~\citep{Poisson:2004lr, Glampedakis:2005hs} that we shall not 
261: consider here.
262: 
263: The study of the relativistic corrections considered in this work
264: leads to the following conclusions. First, we find that relativistic
265: effects are significant for approximately $50\%$ of the orbits
266: contained in the EMRB phase space considered by
267: \citet{Rubbo:2006dv,Rubbo:2007}.  These relativistic EMRB orbits
268: differ from their Newtonian counterparts in such a way that the
269: associated waveforms present a noticeably different structure.  In
270: particular, we find that there is a dephasing relative to Newtonian
271: waveforms that is due to precessional effects and depend strongly on
272: the MBH spin.  These findings show that EMRB events are relativistic
273: enough that they should be treated accordingly, as was previously
274: found for EMRIs \citep{Glampedakis:2005hs}.
275: 
276: Second, we find that the corrections to the trajectories affect the
277: waveforms much more than the corrections in the waveform generation
278: over several bursts.  For example, for a given relativistic
279: trajectory, we find that the difference between the SNR of a waveform
280: obtain from the quadrupole formula to that obtained from the
281: quadrupole-octopole formula is of the order of $10 \%$ (depending on
282: the location of the observer.)  On the other hand, using the same
283: waveform generation formula (quadrupole or quadrupole-octopole), the
284: difference between the SNR of a Newtonian waveform to that of a Kerr
285: waveform is of the order of $100 \%$. These findings show that
286: modeling EMRB waveforms with a quasi-Newtonian treatment might not be
287: sufficient for certain data analysis applications.
288: 
289: Third, we find that the relativistic corrections accumulate with
290: multiple bursts and, thus, they may have an important impact in
291: improving the SNR. It is also conceivable that such corrections might
292: be important for parameter estimation studies and, perhaps, may be
293: used to determine or bound the spin of the MBH if a high SNR event is
294: detected. Along this same lines, if parameters can be determined
295: accurately enough, it might also be possible to use EMRB measurements
296: to test deviations from General Relativity. We must note, however,
297: that changing the orbital parameters in Newtonian waveforms could
298: somewhat mimic some of the relativistic corrections, but a detailed
299: Fisher analysis of such effects is beyond the scope of this paper. 
300: 
301: The remainder of this paper is divided as follows:
302: Section~\ref{dynamics} deals with the dynamics of EMRBs in the
303: semi-relativistic approximation and justifies the use of this
304: approximation for these systems; Section~\ref{waveforms} reviews the
305: inclusion of higher-order multipolar corrections to the waveform
306: generation formalism; Section~\ref{numsim} describes the numerical
307: implementation of the equations of motion and the initial data used;
308: Section~\ref{comps} compares the orbital trajectories and waveforms;
309: Section~\ref{conclusions} concludes and points to future research.
310: 
311: In this paper, we denote the MBH mass by $\Mbh$ and its {\em
312: gravitational radius} by $\Rbh = 2G\Mbh/c^2$, where $c$ is the speed
313: of light and $G$ the Newtonian gravitational constant.  To simplify
314: some expressions we normalize masses with respect to $\Mmw = 4\times
315: 10^6\,\Msun$, which is of the same magnitude as the mass of the MBH at
316: the center of the Milky Way \citep{Ghez:2005}.  The gravitational
317: radius can then be written as:
318: %
319: \begin{eqnarray}
320:   \Rbh &=& (3.82\times 10^{-7}~\textrm{pc}) \frac{\Mbh}{\Mmw} \,.
321: \end{eqnarray}
322: %
323: 
324: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
325: %%%%%%%%%%%%%%%%%%%  EMRB DYNAMICS  %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
326: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
327: 
328: \section{EMRB Dynamics} \label{dynamics}
329: 
330: In this section, we discuss the description of the orbital motion.
331: Newtonian dynamics usually provides an adequate description of many
332: astrophysical sources of gravitational waves, at least from a
333: qualitative point of view.  However, for certain gravitational wave
334: sources, such a description is insufficient and relativistic effects
335: have to be considered.  For EMRB sources with pericenter distances
336: $r_p > 4\Rbh$ and velocities $v_{p}/c < 0.5$, the semi-relativistic
337: approximation to the equations of motion, in combination with a
338: multipolar description of the gravitational radiation, can adequately
339: model the dynamics and gravitational radiation, as we argue below.
340: 
341: The semi-relativistic approximation treats the motion of the SCO in
342: the point-particle limit as a geodesic of the MBH geometry, which is
343: justified based on the small mass-ratios associated with these
344: systems.  In this work, we adopt Cartesian Kerr-Schild coordinates,
345: $\{t,x^i\}$ ($i=1,2,3$), in which the MBH geometry is
346: time-independent, reflecting its stationary character, and tends to a
347: flat-space geometry in Cartesian coordinates far from the MBH.  We
348: denote the geodesic trajectory by $z^i(t)$, its spatial velocity by
349: $v^i(t) = dz^i/dt$, and its spatial acceleration by $a^i(t) =
350: dv^i/dt$.  The latter, in such a coordinate system, and by virtue of
351: the geodesic equations of motion, has the following form
352: \citep[e.g. see][]{Marck:1995kd}:
353: %
354: \begin{equation} \label{geodesics}
355:   a^i = F^i[v^j;\met^{}_{\mu\nu},\partial^{}_j\met^{}_{\mu\nu}] \,,
356: \end{equation}
357: %
358: where $\met^{}_{\mu\nu}$ ($\mu,\nu=0,1,2,3$) are the spacetime
359: components of the MBH metric. These equations describe the influence
360: of the spacetime curvature produced by the MBH and approach the
361: Newtonian equations of motion in the regime where $v/c = |v^i|/c \ll
362: 1$ and $G\Mbh/(c^2\,r) \ll 1$ ($r=|x^i|$).
363: 
364: The effects from the self-gravity of the SCO can be neglected.  To see
365: this, consider the (Keplerian) orbital timescale, $T^{}_{\orb}$, in
366: comparison to a characteristic radiation-reaction timescale,
367: $T^{}_{\rr}$.  For the latter, we can use the timescale associated
368: with the rate of change of the semi-latus rectum, $p$, related to the
369: pericenter distance by $r_p = p/(1 + e)$, namely $T^{}_{\rr} \sim
370: p/|dp/dt|$.  The radiation-reaction timescales of the other orbital
371: elements are comparable or larger \citep[see,
372: e.g.,][]{Glampedakis:2005hs}. The ratio of these timescales is
373: %
374: \begin{equation} \label{eq:tscales}
375:   \frac{T^{}_{\orb}}{T^{}_{\rr}} \sim 2 \pi \mu
376:   \left(\frac{\Rbh}{2p}\right)^{5/2}\,,
377: \end{equation}
378: %
379: where $\mu = m/\Mbh$ is the mass ratio of the system and $m$ the SCO
380: mass.  It is evident that the radiation-reaction timescale is much
381: greater than the orbital timescale due to the extreme mass ratio, $\mu
382: \ll 1$, and because EMRBs have $p \gg \mu^{5/2} \Rbh$. In the unlikely
383: case that more accuracy is required, one could improve the analysis
384: through the use of ``Kludge'' waveforms \citep{Babak:2006uv}, which
385: have been shown to reproduce numerical results in the adiabatic
386: approximation accurately for EMRIs.
387: 
388: Formally, the orbital timescale used is not really the exact timescale
389: of orbital motion. This is because the mass distribution of a
390: MBH-embedded galaxy possesses a non-Keplerian potential that leads to
391: non-Keplerian orbits.  However, most EMRBs (by rate) have apocenters
392: that do not extend far into the stellar population, implying that the
393: contribution from the galaxy potential is minimal.  The orbits we
394: study in later sections have a contamination from the galactic
395: potential that is less than $2\%$ of the MBH mass.  Moreover,
396: \citet{Hopman:2006fc} rightly argue that the inner region is
397: statistically empty of stars, which is due to finite effects realized
398: at the small scales observed near the MBH.
399: 
400: Certain constraints may be derived on the size of $p$ and $r_{p}$ for
401: EMRB events. The most important constraint is derived from the
402: definition of EMRBs: orbits with sufficiently large orbital period
403: $T_{\orb} > T_{\cut}$.  Assuming a Keplerian orbit (which is a rough
404: assumption), this constraint translates to pericenter distances as
405: follows
406: %
407: \begin{equation}
408:   r_p > (7.98 \times 10^{-7}~\textrm{pc}) \frac{(1 -e)}{0.1}
409:   \left(\frac{\Mbh}{\Mmw}\right)^{1/3}
410:   \left(\frac{T}{T_{\cut}}\right)^{2/3} \,, \label{orbitalconstraint}
411: \end{equation}
412: %
413: where we have rescaled quantities assuming a typical eccentricity of
414: $e = 0.9$ \citep{Rubbo:2006dv,Rubbo:2007} and a typical MBH mass of
415: $\Mbh = \Mmw$.  In terms of geometrized units, such a constraint
416: translates roughly to $r_{p} > 2\Rbh$.
417: 
418: This constraint can be compared with the requirement that the SCO does
419: not get captured. Any object that enters the black hole event horizon
420: is captured, where the horizon is located (in Boyer-Lindquist
421: coordinates) at
422: %
423: \begin{equation}
424:   r_{\rm cap} = (1.91 \times 10^{-7}~\textrm{pc})\,
425:   \frac{\Mbh}{\Mmw}\, \left[ 1 + \sqrt{1 -
426:   \frac{\abh^2}{\Mbh^2}}\;\right]\,,
427: \end{equation}
428: %
429: where $\abh$ is the (Kerr) MBH spin parameter, related to its
430: intrinsic angular momentum by $\Sbh = G\Mbh\,\abh/c$, and bounded by
431: $\abh/\Mbh \leq 1$.  Thus, for a maximally spinning Kerr MBH ($\abh =
432: \Mbh$), $r_{\rm cap} = 0.5\Rbh$, while for a Schwarzschild
433: (non-spinning) MBH it is just $\Rbh$.  This condition tell us simply
434: that $r_p > r_{\rm cap}$, which is a condition superseded by the
435: constraint on the orbital period given in
436: equation~\eqref{orbitalconstraint}.  One could explore other possible
437: constraints \citep{Rubbo:2006dv, Rubbo:2007} but they are in general
438: superseded by equation~\eqref{orbitalconstraint}.
439: 
440: These constraints clearly exclude the ergosphere of the MBH ($r_{\rm
441: cap} < r \lesssim \Rbh$) where frame dragging effects are most
442: pronounced.  However, for EMRIs it has been argued
443: \citep{Glampedakis:2005hs} that orbits with $r_p < 10 \Rbh$ cannot be
444: considered Keplerian anymore, mainly due to precessional effects. This
445: statement can be made more quantitative by looking at the ratio of
446: first-order post-Newtonian (1~PN) predictions \citep{Blanchet:2002av}
447: to Newtonian ones (0~PN). For example, for the energy of a circular
448: orbit, this ratio scales as $7 \Rbh/(8 r_p)$, while for the perihelion
449: precession angle, the ratio scales as $3 \Rbh/(2 r_p)$, for
450: extreme-mass ratios.  Therefore, for orbits with pericenter passage
451: $r_p \sim 5\Rbh$, the 1~PN correction to the energy and the perihelion
452: precession angle is approximately $20 \%$ and $30 \%$ respectively,
453: relative to the Newtonian value. This indicates that, even for orbits
454: outside the ergosphere, relativistic effects are not necessarily
455: negligible.
456: 
457: The relativistic geodesic equations of motion introduce corrections to
458: the Newtonian motion that can be interpreted in terms of a black hole
459: \textit{effective} potential.  By comparing the Newtonian and
460: relativistic potentials one can see that the relativistic corrections
461: dominate over the centrifugal barrier at small distances from the
462: black hole center.  In this work we show that these relativistic
463: corrections can be sampled by EMRBs and hence, one should model these
464: systems accordingly.  Nevertheless, as we argued above, the
465: relativistic treatment of EMRBs does not need to be as sophisticated
466: as in the case of EMRIs, since radiation-reaction can be neglected.
467: 
468: 
469: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
470: %%%%%%%%%%%%%%%%%%% BURST WAVEFORMS %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
471: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
472: 
473: \section{EMRB Waveforms} \label{waveforms}
474: 
475: In this section we describe how we extract gravitational waveforms
476: once we have integrated the geodesic equations of motion. We use a
477: multipole-moment wave generation formalism for slowly-moving objects
478: with arbitrarily strong internal gravity \citep{Thorne:1980rm,
479: Flanagan:2005yc, Glampedakis:2005hs}. In quasi-Newtonian and
480: semi-relativistic treatments, the radiation is modeled by the lowest
481: non-vanishing multipole moment: the mass-quadrupole.  To that order,
482: and for the case of a point-like object orbiting a MBH at a fixed
483: coordinate location, the plus and cross polarizations are given by
484: \citep{Misner:1973cw, Thorne:1980rm}
485: %
486: \begin{equation} \label{waves-quad}
487:   h_{+,\times} = \frac{2 G m}{r c^4} \; \epsilon^{ij}_{+,\times}
488:   \left( a_i z_j + v_i v_j \right),
489: \end{equation}
490: %
491: where $r$ is the (flat-space) distance to the observer and
492: $\epsilon^{ij}_{+,\times}$ are polarization tensors.  This expression
493: assumes, based on the slow motion approximation, that the change in
494: the acceleration with respect to time, the \textit{jerk}, $j^i =
495: da^i/dt$, is a small quantity.  More precisely, we are neglecting
496: terms of order $(v/c)^3$, or in other words, since the (quadrupole)
497: leading order terms are of order $(v/c)^2$, this implies a relative
498: error of order $v/c$.  
499: 
500: One can improve on this description for the gravitational radiation by
501: accounting for higher-order multipole moments. In this paper, we
502: consider the mass-octopole and current-quadrupole multipoles, which
503: require the knowledge of one more time derivative of the trajectory,
504: the jerk.  Adding these contributions to equation~\eqref{waves-quad},
505: the gravitational waveforms are given by~\citep{Thorne:1980rm}
506: %
507: \begin{eqnarray} \label{waves-oct}
508:   h_{+,\times} &=& \frac{2 G m}{r c^4} \; \epsilon^{ij}_{+,\times} \bigg\{
509:   a_i z_j + v_i v_j \nonumber\\
510:   &+& \frac{1}{c} \Big[ \left({\bf{n}} \cdot {\bf{z}}\right) \left(z_i j_j + 3 a_i v_j
511:   \right) + \left({\bf{n}} \cdot {\bf{v}}\right)
512:   \left( a_i z_j + v_i v_j \right) \nonumber\\ 
513:   &-& \left({\bf{n}} \cdot {\bf{a}}\right) v_i z_j - \frac{1}{2}
514:   \left({\bf{n}} \cdot {\bf{j}}\right) z_i z_j \Big] \bigg\}, 
515: \end{eqnarray}
516: %
517: where $n^i = x^i/r$ is a unit vector that points to the observer and
518: the vector product is the flat-space scalar product.  In this case, we
519: are neglecting terms of order $(v/c)^4$ and hence we are making a
520: relative error of order $(v/c)^2$ with respect to the leading order
521: quadrupole term.
522: 
523: The waveforms of equation~\eqref{waves-oct} are a truncated multipole
524: expansion, where we are neglecting the current-octopole,
525: mass-hexadecapole, and higher multipole moments. This expansion is
526: based on a slow-motion approximation which is valid for orbits whose
527: pericenter velocity is small relative to the speed of light.  For
528: closed circular orbits, we can use the Virial theorem to argue that
529: this is equivalent to requiring $r_{p} > \Mbh$. For a relativistic
530: EMRB event with $v_{p}/c = 0.4$, the maximum relative contribution of
531: the octopole to the quadrupole is of the order of~$40\%$, since the
532: octopolar term is of order $v/c$ smaller than the quadrupolar one. In
533: this paper, we shall study EMRBs from the sample of Milky Way sources
534: studied in~\cite{Rubbo:2006dv}. These sources have initial pericenter
535: distances of $r_{p} > 8 \Mbh$, thus justifying the use of a
536: low-multipolar expansion in the wave-generation formalism and the
537: neglect of radiation reaction effects in the orbital motion.
538: 
539: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
540: %%%%%%%%%%%%%%%%%%%%%   NUMERICAL SIMULATIONS %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
541: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
542: 
543: \section{Numerical Simulations} \label{numsim}
544: 
545: In this section we describe the EMRB simulations that were carried
546: out, including the choice of initial conditions.  The simulations
547: involve integrating the equations for geodesic motion around a Kerr
548: black hole, equation~\eqref{geodesics}, forward in time. \citep[For a
549: detailed exposition of Kerr geodesics see][]{Chandrasekhar:1992bo}.
550: Since Cartesian Kerr-Schild coordinates are used, the initial
551: conditions can be denoted by $(z^i_0,v^i_0)$.  The numerical
552: implementation does not use the Kerr geodesic constants of motion
553: (energy, angular momentum, and Carter constant) in order to reduce the
554: number of variables of the resulting system of ordinary differential
555: equations.  Instead, we have used the constants of motion to monitor
556: the accuracy of the time integration.  The integration accuracy is set
557: so that we obtain fractional errors for the constants of motion
558: smaller than one part in $10^{10}$.  The code uses a Bulirsh-Stoer
559: extrapolation method as the evolution algorithm \citep[see,
560: e.g.][]{Press:1992nr,Stoer:1993sb}.  We have also introduced in the
561: code the possibility of switching between Kerr geodesics and Newtonian
562: equations of motion.  The gravitational waveforms are then obtained
563: directly by applying expressions~\eqref{waves-quad} and
564: \eqref{waves-oct} to the numerically obtained trajectory $z^i(t)$.
565: 
566: Comparisons are carried out by choosing a representative relativistic
567: orbit within the allowed phase space for EMRB's.  We made the
568: following choices for the test case:
569: %
570: \begin{itemize}
571: \item The central MBH mass is $\Mbh = \Mmw$ and the SCO mass is $\Mco
572:   = 1~\Msun$, such that the mass ratio is $\mu = \Mco/\Mbh = 2.5
573:   \times 10^{-7} \ll 1$.
574: 
575: \item The MBH spin parameter is either $\abh=0$ (Schwarzschild) or $\abh
576:   = 0.998~\Mbh$ (Kerr).  The angular momentum is aligned along the
577:   $z$-axis and equal to either $S^z = 0$ or $S^z = 0.998 G\Mbh^2/c$.
578:   
579: \item The observer is located at $r_{obs} = 8$ kpc along the z-axis,
580:   which corresponds to the approximate distance from Earth to the center
581:   of the Milky Way \citep{Eisenhauer:2005}.
582: \end{itemize}
583: %
584: Furthermore, we make the following choices for the orbital initial
585: conditions: 
586: %
587: \begin{eqnarray} \label{orbit:ID}
588:   z^i_0 &=& \left(-1.59, 1.05, -0.185\right) \times 10^{-5}~{\textrm{pc}}, 
589:   \nonumber\\
590:   v^i_0 &=& \left(1.70,-2.89,0.510\right)\times 10^{4}~{\textrm{km
591:   s}}^{-1}\,.
592: \label{testorbit}
593: \end{eqnarray}
594: %                  
595: The initial conditions are such that $r_{0} = |z^i_0| = 50 \Rbh = 1.91
596: \times 10^{-5}$~pc, and $|v^i_0| = 0.11\,c = 3.39 \times
597: 10^{4}~{\textrm{km s}}^{-1}$.  The orbital plane is inclined by
598: $10^{\circ}$ with respect to the $x-y$ plane to demonstrate the
599: effects of orbital plane precession, which only occurs for spinning
600: MBHs. Since this paper is concerned with the effect of relativistic
601: corrections to EMRB events, we choose to give all orbits the same
602: initial conditions. The possibility of relaxing this choice and its
603: effect on the conclusions derived in this paper shall be discussed in
604: a later section.
605: 
606: Although the test orbit is in the phase space of EMRB events studied
607: in~\citet{Rubbo:2006dv}, one might worry that it is too relativistic
608: to actually have a significant probability to occur in nature.  In 
609: particular, one can think that the SCO may be tidally disrupted. To
610: address this question, let us consider a Newtonian description of the 
611: central potential, which leads to the following values of the pericenter 
612: distance and velocity:
613: %
614: \begin{eqnarray}
615:   r_{p} &=& 6.45~\Rbh = 2.48 \times 10^{-6}~{\textrm{pc}},
616:   \nonumber \\
617:   |v_{p}| &=& 0.384\,c = 1.15 \times 10^{5}~{\textrm{km s}}^{-1} \,.
618: \end{eqnarray}
619: %
620: One might worry that stars might be tidally disrupted with such small
621: pericenters. However, as shown by \citet{Hopman:2006fc}, most SCOs in
622: EMRB scenarios consist of stellar-mass black holes, which cannot be
623: tidally disrupted.
624: 
625: Such relativistic orbits are actually naturally occurring in the phase
626: space of possible EMRBs studied in~\citet{Rubbo:2006dv}. Of all
627: orbits in the allowed EMRB phase space considered
628: in~\citet{Rubbo:2006dv, Rubbo:2007}, $6\%$ is contained within a small
629: $6$-dimensional phase space volume centered on the test
630: orbit~\footnote{In other words, $6\%$ of the test orbits considered to
631:   be EMRBs in ~\citet{Rubbo:2006dv} are close in phase space to our
632:   test orbit. This does not imply that the probability of such a test
633:   orbit actually occurring in nature is $6 \%$, since any element of
634:   phase space may have small overall probabilities, even down to
635:   $10^{-7} \%$.}. Furthermore, the test case shown possesses a short
636: orbital timescale, bursting approximately $150$ times per year. Such
637: events with small orbital timescale were shown to dominate the EMRB
638: event rate in~\citet{Rubbo:2006dv}. It is in this sense that the test
639: orbit studied here is {\emph{typical}} or representative of EMRBs.
640: 
641: The relative location of the test orbit in the pericenter
642: distance-eccentricity plane of the phase space of allowed
643: EMRBs~\citep{Rubbo:2006dv} is presented in Figure~\ref{ID} (triangle.)
644: The eccentricity was here calculated assuming a Newtonian orbit and
645: the pericenter separation is given in gravitational radii, $\Rbh$.
646: Although the test orbit has a large eccentricity, its apocenter is
647: small enough ($r_{a} \lesssim 150 \Rbh \approx 6 \times 10^{-5}$~pc)
648: that the contribution from the surrounding stellar population to the
649: potential can be neglected.  In general, the left side of the figure
650: corresponds to highly relativistic orbits with large pericenter
651: velocities and small pericenter distances.  Orbits with pericenter
652: velocities $|v_p| > 0.25\,c = 3 \times 10^4~{\textrm{km s}}^{-1} $
653: account for approximately 50\% of the possible orbits within the phase
654: space studied in~\citet{Rubbo:2006dv}.
655: %
656: \begin{figure}
657: \includegraphics[scale=0.375,clip=true]{f1.eps}
658: \caption{\label{ID} Plot of the set of possible EMRB orbits as
659:   computed in \citet{Rubbo:2006dv} in the pericenter
660:   distance-eccentricity plane. The pericenter distance is given in
661:   units of the gravitational radius $\Rbh$. The initial conditions
662:   for the test [Eq.~(\ref{testorbit})] and the extreme 
663:   [Eq.~(\ref{extremeorbit})] orbits are indicated by a triangle and
664:   a circle respectively.}
665: \end{figure}
666: %
667: 
668: Although the test orbit is a good source to demonstrate the
669: differences between the Newtonian and relativistic treatments, we
670: could have chosen an even more relativistic one. Such an event would
671: still be classified as an EMRB in a Newtonian treatment, but it would
672: border with the definition of a continuous source.  An example of such
673: an extreme orbit is shown with a circle in Figure~\ref{ID}, to the
674: left of the test orbit (triangle).  This extreme orbit possesses the
675: following initial conditions:
676: %
677: \begin{eqnarray}
678:   z^i_0 &=& \left(-1.81,0.6,-1.06\right) \times 10^{-6}~{\textrm{pc}},
679:   \nonumber \\
680:   v^i_0 &=& \left(1.72,-1.78,0.31\right) \times 10^{5}~{\textrm{km
681:   s}}^{-1},
682:   \label{extremeorbit}
683: \end{eqnarray}
684: %
685: where $r_p = 4\,\Rbh = 1.53 \times 10^{-6}$~pc and $|v_p| = 0.49c =
686: 1.46 \times 10^5~{\textrm{km s}}^{-1}$ for a Newtonian potential. We
687: will study such an extreme orbit at the end of the next section as an
688: example of a limiting relativistic case.
689: 
690: 
691: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
692: %%%%%%%%%%%%%%%%%%%%%%   COMPARISONS  %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
693: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
694: 
695: \section{Comparison of Trajectories and Waveforms} \label{comps}
696: 
697: In this section we compare the results obtained for both the orbital
698: motion and the gravitational radiation emitted by an EMRB event using
699: both the Newtonian and relativistic description.  Since the {\em plus}
700: and {\em cross} polarization waveforms present similar features, we
701: only plot the {\em plus} polarization waveforms.  In the remainder of
702: this section we use the following nomenclature: a quadrupolar
703: (octopolar) Newtonian waveform is one that was calculated using the
704: quadrupole (octopole) formula and Newtonian equations of motion; a
705: quadrupole (octopole) Schwarzschild waveform is one that was
706: calculated using the quadrupole (octopole) formula and the geodesic
707: equations of motion with no spin ($\abh=0$); a quadrupole (octopole)
708: Kerr waveform is one that was calculated using the quadrupole
709: (octopole) formula and the geodesic equations of motion with spin
710: $\abh = 0.998 \Mbh$.
711: 
712: %----------------------------------------------
713: \subsection{Orbital Trajectories}
714: 
715: Let us begin by comparing the trajectories obtained in our
716: simulations.  In Figure~\ref{Orbits}, we plot the test orbit
717: corresponding to a Newtonian treatment (solid line) and the one
718: corresponding to a relativistic treatment without spin (dashed line)
719: and with spin (dotted line).  The dot and arrow indicate the initial
720: location and velocity projected onto the $x$-$y$ plane. The MBH is
721: located at the origin of the coordinates, and the vectors denoted by
722: $L$ and $S$ describe the direction of the initial orbital angular
723: momentum and the MBH spin respectively.  In the relativistic
724: description there are precessional effects in the SCO orbit that can
725: be clearly observed in Figure~\ref{Orbits}.  These precessional
726: effects are: pericenter precession about the orbital angular momentum
727: axis, which acts in the initial orbital plane; and frame-dragging
728: precession about the total angular momentum axis, which acts out of
729: the initial orbital plane.  While the former always occurs in a
730: relativistic treatment, the latter is present only in the spinning
731: case.
732: %
733: \begin{center}
734: \begin{figure}
735:  \includegraphics[scale=0.85,clip=true]{f2.eps}
736: \caption{\label{Orbits} Trajectories for the SCO, with initial
737:   conditions given by~\eqref{orbit:ID}, corresponding to a Newtonian
738:   description (solid line) and relativistic descriptions with no spin
739:   (dashed line) and with spin (dotted line).  The MBH is located at
740:   the origin and the vectors $L$ and $S$ denote the initial orbital
741:   angular momentum and the MBH spin respectively. }
742: \end{figure}
743: \end{center}
744: %
745: 
746: Different relativistic precessional effects are generally of different
747: magnitude. These effects are usually inversely proportional to the
748: pericenter distance, or equivalently proportional to the pericenter
749: velocity of the SCO.  Precession out of the initial orbital plane,
750: however, is smaller than precession in the orbital plane by a relative
751: factor of order $v_{p}/c$ and it is directly proportional to the spin
752: of the MBH. In terms of post-Newtonian theory \citep[see,
753: e.g.][]{Blanchet:2002av}, the pericenter advance is described by
754: 1st-order post-Newtonian corrections to the equations of motion (order
755: $(v/c)^2$ relative to the Newtonian acceleration), while precession
756: off the orbital plane is due to spin-orbit and spin-spin couplings
757: that correspond to $1.5$ and $2$-order corrections (order $(v/c)^3$
758: and $(v/c)^4$ relative to the Newtonian acceleration.)  Therefore,
759: since EMRBs are characterized as events with small to moderate
760: pericenter velocities, precession out of the initial orbital plane is
761: small to moderate relative to pericenter advance, even for maximally
762: spinning MBHs.
763: 
764: We can estimate the precession rate by comparing the Newtonian and
765: relativistic trajectories.  For the test orbit considered, we find
766: that the rate in the orbital plane is roughly $\pi/3$ radians per
767: orbit for the non-spinning case and $2\pi/3$ radians per orbit for the
768: spinning one. These precessional effects have been studied extensively
769: in the context of EMRIs \citep[see, e.g.][]{Schmidt:2002} and also
770: specifically for S-stars in the central region of our Galaxy in
771: \citet{Kraniotis:2007}.  Nonetheless, these effects have not been
772: previously analyzed in the context of EMRBs, since previous studies
773: employed a quasi-Newtonian treatment. 
774: 
775: %----------------------------------------------
776: \subsection{Waveforms}
777: Let us now analyze how the differences in the SCO trajectories
778: translate into different signatures in the waveforms.  In
779: Figure~\ref{hp1} we plot the quadrupole Newtonian and Schwarzschild
780: waveforms (solid and dashed lines respectively), while in
781: Figure~\ref{hp2} we plot the quadrupole Schwarzschild and Kerr
782: waveforms (dashed and dot-dot-dashed lines respectively.) There are
783: three main differences between the Newtonian and the relativistic
784: waveforms: a modulated phase change, an amplitude change, and a time
785: of arrival change.  The changes in amplitude and time of arrival are
786: due to the test particle experiencing a larger ``force'' of attraction
787: as it approaches the black hole.  Quantitatively, this increase in
788: force is due to the presence of ($r_p^{-n}$)-contributions to the
789: relativistic corrections to the central potential (with $n$ a real
790: positive number.)
791: 
792: %
793: \begin{figure}[tl]
794:   \includegraphics[scale=0.35,clip=true]{f3.eps}
795: \caption{\label{hp1} EMRB waveforms ({\em plus} polarization):
796: The Newtonian waveform corresponds  to the solid line and
797: the Schwarzschild one to the dashed line.}
798: \end{figure}
799: %
800: 
801: %
802: \begin{figure}[tr]
803: \includegraphics[scale=0.35,clip=true]{f4.eps}
804: \caption{\label{hp2} EMRB waveforms ({\em plus} polarization):
805: The Schwarzschild waveform corresponds to the dashed
806: line and the Kerr one to the dot-dot-dashed line.
807: The dephasing of the waveforms can be best observed during the
808: silent transitions between bursts. For example, in the first silent
809: transition the waveforms are roughly $\pi$ radians out of phase, while
810: in the third one they are in phase.}
811: \end{figure}
812: %
813: 
814: Gravitational wave interferometers are most sensitive to the phase,
815: which is clearly different for Newtonian and relativistic waveforms.
816: The dephasing present in Figures~\ref{hp1} and \ref{hp2} parallels the
817: orbital dephasing discussed earlier (the gravitational-wave and
818: orbital frequencies are intimately related) and leads to an amplitude
819: modulation. In terms of the gravitational wave phase, both
820: Figures~\ref{hp1} and \ref{hp2} show a dephasing of $\pi/6$ radians
821: per cycle.  This can be seen after the third burst where the waveforms
822: are back in phase. In fact, there is a significant dephasing even
823: between the relativistic waveforms due to the effect of the MBH spin.
824: If a cursory examination by eye can detect the difference in the
825: waveforms due to differences in the nature of the central potential,
826: it stands to reason that strong-field EMRB waveforms might allow us to
827: probe of the spacetime near a MBH. We should note, however, that no
828: work has yet been done to find best-fit parameters for Newtonian
829: waveforms that maximize the correlation with relativistic ones. In
830: other words, it might be possible to mimic some of the relativistic
831: corrections by choosing different initial data for the Newtonian
832: waveforms, but such a study is beyond the scope of this paper.
833: 
834: The difference in dephasing can be better studied by calculating the
835: signal overlap,
836: %
837: \begin{equation} \label{corr}
838:   (h_{1} | h_{2}) = \frac{\int_{0}^{T} h_{1}(t)
839:   h_{2}(t) dt}{\sqrt{ \int_{0}^{T} h_{1}^{2}(t) dt} \;
840:   \sqrt{\int_{0}^{T} h_{2}^{2}(t) dt}} \;.
841: \end{equation}
842: The overlap indicates how well a signal $h_{1}$ can be extracted via
843: matched filtering with a template $h_{2}$~\footnote{The overlap is
844:   given here in the time domain, but an analogous representation in
845:   the frequency domain could also be used. Such an expression in the
846:   frequency domain can be derived through Parseval's theorem.}.  In
847: Figures~\ref{power1} and~\ref{power2} we plot the normalized integrand
848: as a function of time, with $h_{1}$ given by the quadrupole Kerr
849: waveform and $h_{2}$ given by either the quadrupole Newtonian or
850: Schwarzschild waveforms.  Observe that neither the Newtonian nor the
851: Schwarzschild waveforms match well with the Kerr waveform.  Moreover,
852: note that the correlation with the Newtonian waveform deteriorates
853: greatly after only the first cycle. The integral of
854: equation~\eqref{corr} gives the correlation between different
855: waveforms over nine days (four bursts): for the Newtonian and Kerr
856: plus-polarized waveforms it is $9.6 \%$; for the Schwarzschild and
857: Kerr plus-polarized waveforms it is $-6.3 \%$.  As a point of
858: comparison, a substantial signal overlap should be~$\gtrsim 95 \%$. As
859: already mentioned, the same initial conditions were chosen for both
860: the Newtonian and relativistic orbits, such that their waveforms would
861: be both initially in-phase and any dephasing due to relativistic
862: effects could be clearly studied. However, such a choice forces the
863: SCO to pass through periapsis at slightly different times, because in
864: the relativistic case this object experiences a ``deeper'' potential.
865: Such a difference in timing degrades the overlap somewhat and could,
866: in principle, be ameliorated by choosing different initial conditions
867: for the Newtonian evolution, {\textit{i.~e.~}} by maximizing the
868: overlap over all orbital parameters, but such a study is beyond the
869: scope of this paper.
870: 
871: Figures~\ref{power1} and~\ref{power2} provide some evidence that the
872: use of a relativistic waveform might be required for the data analysis
873: problem of extracting EMRB signals. Such expectations are somewhat
874: confirmed in Table~\ref{table_orbs}, where we present the correlation
875: between Newtonian and Kerr plus-polarized waveforms integrated over
876: both a single day (one burst) and nine days of data (several bursts)
877: for a sample of different EMRB orbits~\footnote{Here we use the
878:   frequency representation of the correlation calculation, employing
879:   the Fourier transform of the waveforms.}. Both a single and several
880: bursts should be studied because, in principle, parameter adjustments
881: might mitigate the between-burst dephasing, but probably not the
882: in-burst dephasing. The orbits chosen were taken directly from the
883: allowed EMRB phase space of~\citet{Rubbo:2006dv} and possess different
884: initial positions and velocities, leading to different eccentricities,
885: initial inclination and orbital periods. The orbital period can be
886: used to classify the orbits into highly-bursting (burst more than once
887: per week) and slowly-bursting (burst less than once per week.)  For
888: highly-bursting EMRBs, the average correlation is $\sim 27 \%$ when
889: all nine days of data (several bursts) are used, while it is $\sim 93
890: \%$ when only one day of data (single burst) is used. For the
891: slowly-bursting EMRBs we considered here, there is actually only one
892: burst per week and its correlation is $\sim 85 \%$. These results
893: indicate that accumulating precession effects lead to a significant
894: loss of correlation between Newtonian and Kerr waveforms if bursts are
895: to be connected. Moreover, we see that even for a single burst the
896: shape of the relativistic waveforms is sufficiently different from its
897: Newtonian counterpart to lead to a significant mismatch (in-phase
898: dephasing). If one maximizes the correlation over intrinsic orbital
899: parameters it might be possible to increase the correlation somewhat,
900: but again that is to be studied elsewhere.
901: 
902: %
903: \begin{figure}[tl]
904: \includegraphics[scale=0.33,clip=true]{f5.eps}
905: \caption{\label{power1} Plot of the overlap integrand of
906:   equation~\eqref{corr} with $h_{1}$ given by the quadrupole Kerr
907:   waveform and $h_{2}$ by the quadrupole Newtonian one.  The inset
908:   zooms to a region near the small peaks.}
909: \end{figure}
910: %
911: 
912: %
913: \begin{figure}[tr]
914: \includegraphics[scale=0.33,clip=true]{f6.eps}
915: \caption{\label{power2} Plot of the overlap integrand of
916:   equation~\eqref{corr} with $h_{1}$ given by the quadrupole Kerr
917:   waveform and $h_{2}$ by the quadrupole Schwarzschild one.}
918: \end{figure}
919: %
920: 
921: Let us now focus on the differences in the waveforms when they are
922: calculated with the quadrupolar approximation versus the
923: quadrupolar-octopolar one. In Figure~\ref{diff} we plot the absolute
924: value of the difference between the octopole and quadrupole waveforms
925: as a function of time for a Schwarzschild (upper panel) and a Kerr
926: (lower panel) central potential. The difference is normalized to the
927: maximum of the first peak of the quadrupole waveform, since other
928: peaks have approximately the same maximum.
929: %
930: \begin{figure}
931: \includegraphics[scale=0.33,clip=true]{f7.eps}
932: \caption{\label{diff} Plot of the absolute magnitude of the difference
933:   between the quadrupole and octopole Schwarzschild (upper panel) and
934:   Kerr (lower panel) waveforms. The difference is normalized to the
935:   maximum value of the first peak of the quadrupole waveform.}
936: \end{figure}
937: %
938: Observe that the inclusion of higher-order multipoles does not affect
939: the phasing of the waveforms, but only the amplitude, which is in
940: general different by roughly 4\% relative to the quadrupole waveform
941: for the Kerr test case. At first sight, this result is in disagreement
942: with the expectation that the octopolar correction is at most
943: $\lesssim 40 \%$ of the quadrupolar one. Note, however, that the $40
944: \%$ estimate is an order-of-magnitude {\emph{upper limit}}, since the
945: octopole correction is dependent on the location of the observer
946: relative to the trajectory, velocity, acceleration and jerk vectors.
947: For the test case, where the observer is located on the $\hat{z}$-axis
948: and the orbit is initially inclined by $10^{\circ}$, the octopolar
949: change is reduced by approximately an order of magnitude, since
950: initially $(\mb{n} \cdot \mb{z}) \approx (\mb{n} \cdot \mb{v}) \approx
951: (\mb{n} \cdot \mb{a}) \approx (\mb{n} \cdot \mb{j}) \approx 0.1$. In
952: Table~\ref{table_obs}, we present the approximate maximum difference
953: between octopolar and quadrupolar waveforms as a function of observer
954: location, focusing only on the first burst of radiation. The location
955: of the observer is rotated about the $\hat{y}$-axis on the
956: $\hat{x}-{\hat{z}}$ plane, always at a fixed radial distance of 8~kpc.
957: Note that for some observers the difference is larger and, in fact, of
958: the order of 40\%, since the dot products are closer to unity.  These
959: results are thus consistent with the expectation that the $n$-th
960: multipolar contribution cannot in general be larger than order
961: $(v/c)^n$ relative to the quadrupolar leading term.
962: 
963: \subsection{Data Analysis}
964: 
965: In order to quantify some of our statements about the change in phase
966: and amplitude, we calculated the SNR for the relativistic waveforms
967: via the standard formula
968: %
969: \begin{equation}
970:   \rho^2 = 4 \int_{0}^{\infty} \frac{|\tilde{h}(f)|^{2}}{S_{n}(f)} \;
971:   df \;,
972: \end{equation}
973: %
974: where the tilde denotes the Fourier transform and $S_{n}(f)$ is the
975: one-sided power spectral noise density.  Here we employ the Online
976: Sensitivity Curve Generator \citep{Larson:2000} with the standard LISA
977: settings and the inclusion of the white-dwarf background contribution.
978: When calculating SNRs, we set the observation time to roughly nine
979: days, so as to include multiple bursts in our single SNR value.
980: 
981: The inclusion of relativistic corrections in the trajectories has a
982: dramatic impact in the SNR. We find that the Schwarzschild waveform
983: increases the SNR by a factor of approximately $59\%$, while the Kerr
984: waveform increases it by $162\%$, relative to the Newtonian SNR.  The
985: difference in SNR is because the relativistic orbits experience a
986: deeper effective potential and, thus, the interaction timescale is
987: smaller.  Therefore, the inverse of the interaction time, $f_{\star} =
988: v_p/r_p$, is larger for the Schwarzschild and Kerr waveforms relative
989: to the Newtonian one.  As a result, the Fourier power is shifted to
990: higher frequencies, where LISA is more sensitive.
991: 
992: The SNR behaves similarly for other EMRB orbits with different orbital
993: periods, eccentricities and pericenter parameters. This can be
994: observed in Table~\ref{table_orbs} where we present the SNR difference
995: between Newtonian and Kerr waveforms for different EMRB orbits for a
996: single and several bursts. Highly-bursting EMRBs lead to a large
997: change in the SNR over a week of data, since they experience the
998: depths of the effective potential several times. Per burst, the change
999: in SNR can range from one, to ten or even ninety percent, depending on
1000: the inclination angle of the orbit, the pericenter distance and other
1001: orbital parameters. Also note that the orbits presented in the table
1002: are not as relativistic as the test case, which is why the SNR
1003: difference is smaller. This study seems to indicate that the SNR is in
1004: general somewhat larger for relativistic waveforms, specially when
1005: several burst are taken into account.~\footnote{Naive intuition might
1006:   suggest that the change in the SNR should scale like the square root
1007:   of the number of bursts, but this is not necessarily correct. First,
1008:   different orbits possess different beaming patterns on the sky due
1009:   to precession. Second, the starting frequencies of these bursts is
1010:   very close to the limit of LISA's sensitivity band [$10^{-5}$ Hz].
1011:   As the orbits burst, precession somewhat increases the frequency of
1012:   the waveform, forcing different bursts to contribute different
1013:   amounts to the SNR.}  Consequently, the event rate calculated in
1014: \citet{Rubbo:2006dv, Rubbo:2007}, which in particular summed over all
1015: bursts in one year of data, is an underestimate for their galaxy
1016: model, because some of the systems with a Newtonian SNR $\lesssim 5$
1017: should have been added to the detectable event rate.  However, the
1018: uncertainty in the event rate is still probably dominated by
1019: astrophysical uncertainties and not by the dynamics modeling.
1020: 
1021: Conversely, the inclusion of higher multipole moments to the wave
1022: generation formalism has little to no effect in the SNR. In the
1023: previous section we showed that there was $\approx 4\%$ difference
1024: between the octopole and quadrupole waveforms relative to the maximum
1025: of the first peak of the quadrupole waveform. We further showed that
1026: this difference depends on the location of the observer (see
1027: table~\ref{table_obs}), but for the test case it does not exceed a
1028: maximum of 40\%, which agrees with the multipole-ordering argument
1029: previously described.  However, we also pointed out that the amplitude
1030: difference is confined to sharp peaks in the time domain. Such a
1031: confined change in the waveform amplitude leads to a Fourier power
1032: being dispersed over a large frequency region, including outside the
1033: LISA band.  As a result, there is a correspondingly small change in
1034: the SNR: of the order of $\approx 1\%$ relative to the quadrupolar
1035: formalism. Such a result is in agreement with the analysis
1036: of~\cite{Babak:2006uv}, which was carried out for EMRIs. Therefore, we
1037: see that the change in SNR is primarily dominated by the modifications
1038: introduced in the geodesic description of the equations of motion, and
1039: not in the octopolar correction to the waveform generation.
1040: 
1041: The analysis presented in this section, in particular
1042: Figure~\ref{power1}, makes it clear that relativistic corrections to
1043: the waveforms accumulate with multiple bursts. In other words, over a
1044: single burst (pericenter passage), a quadrupolar waveform calculation
1045: using Newtonian dynamics might be sufficient.  However, if one wants
1046: to estimate parameters associated with the central MBH, then multiple
1047: bursts might be necessary to associate the events to a single SCO
1048: trajectory.  In terms of data analysis, for a detection search it is
1049: simpler to look for a single burst using techniques such as excess
1050: power and wavelet decompositions \citep[e.q.  see][]{Anderson:2000yy,
1051:   Klimenko:2004, Stuver:2006, Camarda:2006}. For estimating MBH
1052: parameters, the results of this paper suggest that multiple bursts
1053: might have to be connected. For this to occur, a single template may
1054: be used, but as our results indicate, the template will need to
1055: incorporate the effects of general relativity.~\footnote{A note of
1056:   caution should be added here, since a detailed study of the
1057:   maximization of the overlap with respect to orbital parameters in
1058:   the Newtonian waveform has not yet been carried out. Indeed, it
1059:   might be possible to mimic some of the relativistic effects with
1060:   Newtonian waveforms with varying parameters, but such mimicking is
1061:   most probably not possible for highly relativistic EMRBs.}
1062: 
1063: At this junction, we should comment on some of the caveats in the
1064: conclusions derived from the analysis presented here. First, in this
1065: paper we have primarily concentrated on the question of characterizing
1066: the gravitational waves through the study of the SNR and overlap. In
1067: this study, however, we have not maximized these data analysis
1068: measures with respect to (intrinsic or extrinsic) orbital parameters.
1069: Although it might be possible to improve the SNR and overlap via
1070: parameter maximization, our study suggests that the introduction of
1071: relativistic effects, such as precession, lead to a clear imprint on
1072: the waveform that might be difficult to mimic with a purely Newtonian
1073: waveform irrespective of its parameters. Second, in this paper we have
1074: only touched the iceberg of the signal characterization and parameter
1075: estimation problem. A possible route to study this problem is through
1076: a numerical Fisher analysis, with the complications derived from the
1077: fact that the waveforms are known only numerically. Furthermore,
1078: increasing the complexity of the waveform will also increase the
1079: computational cost of these studies and, thus, it might be interesting
1080: to investigate whether it is possible or advantageous to search for
1081: individual bursts with similar frequency and identify them as
1082: belonging to the same physical system.  These issues are beyond the
1083: scope of this paper, but they should be addressed in future
1084: investigations.
1085: 
1086: Setting these caveats momentarily aside, let us conclude with some
1087: discussion of the extreme relativistic case introduced earlier. As we
1088: have seen, relativistic corrections can introduce strong modifications
1089: to EMRB waveforms, which depend on how relativistic the EMRB event is
1090: and, in particular, on the pericenter velocity.  The corrections are
1091: particularly strong for the class of EMRBs that inhabit the boundary
1092: between EMRBs and EMRIs, defined by the $T_{\cut} = 3 \times 10^{4}$~s
1093: value, corresponding to the period between apocenter passages.  An
1094: example of such an event is the extreme orbit discussed in
1095: Section~\ref{numsim}, whose waveform is shown in Figure~\ref{extreme}.
1096: Observe that a simplistic Newtonian description misses the rich
1097: structure, in which the SCO whirls twice about the black hole before
1098: zooming out to apocenter again. This behavior is missed entirely when
1099: we evolve the orbit with the Newtonian equations of motion, even
1100: though the same initial conditions were used. Even though in the
1101: previous cases a Newtonian waveform might be able to extract
1102: relativistic events by adjusting intrinsic parameters, such is
1103: definitely not the case for the highly-relativistic event of
1104: Fig.~\ref{extreme}, since no choice of parameters in the Newtonian
1105: waveform can reproduce its multiple-peak structure.
1106: 
1107: Due to their whirling behavior, the extreme orbit waveform resembles
1108: the zoom-whirl events often mentioned in the EMRI literature
1109: \citep{Hughes:2001:a, Hughes:2001:b, Glampedakis:2002}.  However, the
1110: event is still an EMRB and not an EMRI because the period between
1111: apocenter passages is too long.  For our galactic model, we find the
1112: probability of a small region of phase space around this orbit to be
1113: rather high, 10\%.  If this EMRB is detected with sufficiently high
1114: SNR it seems plausible that a parameter estimation analysis would
1115: allow for a determination of the background parameters, such as the
1116: black hole spin. \citet{Barack:2004a} have already investigated LISA's
1117: ability to measure MBH properties using approximate EMRI signals.
1118: They found that, depending on the actual orbital parameters, it will
1119: be possible to measure the MBH spin with fractional errors of
1120: $10^{-3}$ to $10^{-5}$.  This high precision measurement is the result
1121: of observing up to $\sim\!10^{6}$ complete orbits. Conversely, it is
1122: very unlikely that EMRB measurements will be able to match the
1123: measurement capabilities of EMRI signals, since only a few bursts will
1124: probably be available. Whether accurate parameter extraction is
1125: possible can only be determined with a more detailed data analysis
1126: investigation of EMRBs.
1127: 
1128: %
1129: \begin{figure}
1130: \includegraphics[scale=0.33,clip=true]{f8.eps}
1131: \caption{\label{extreme} Plot of the quadrupole Newtonian (solid) and
1132:   Schwarzschild (dashed) gravitational waveform as a function of
1133:   time.}
1134: \end{figure}
1135: 
1136: 
1137: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1138: %%%%%%%%%%%%%%%%%%%  CONCLUSIONS %%%%$%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1139: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1140: 
1141: \section{Conclusions} \label{conclusions}
1142: 
1143: We have studied the effects of relativistic corrections on the
1144: gravitational waves produced by EMRBs. These events originate from
1145: long period orbits of a SCO around a MBH, leading to large-amplitude,
1146: quasi-periodic gravitational wave bursts. Using a more accurate
1147: relativistic treatment of the phenomenon, we have improved on the
1148: waveforms and trajectories relative to previous work.  The orbital
1149: trajectories were corrected by accounting for the spacetime curvature
1150: of the system for Schwarzschild and Kerr MBHs. The waveform generation
1151: was corrected by accounting for the next order term in the multipolar
1152: expansion of gravitational radiation.
1153: 
1154: We found that relativistic corrections change the waveform shape
1155: relative to its Newtonian counterpart. One of the most significant
1156: changes was found to be an amplitude-modulated dephasing, produced by
1157: the relativistic corrections to the orbital trajectory and, in
1158: particular, by relativistic precessional effects. Other effects
1159: included a change in the amplitude of the waveform, partially produced
1160: by the inclusion of higher-order terms in the gravitational wave
1161: generation scheme.
1162: 
1163: The magnitude of the relativistic corrections was found to be directly
1164: proportional to the pericenter velocity of the orbit, as expected.
1165: Surprisingly, we estimated that at least $50 \%$ of the orbits
1166: analyzed in \citet{Rubbo:2006dv,Rubbo:2007} acquire relativistic
1167: velocities and, thus, non-negligible relativistic corrections. We
1168: investigated these corrections in detail by choosing a test orbit,
1169: that is representative of the kind that dominated the event rate
1170: calculation of~\citet{Rubbo:2006dv}. We also studied a limiting case
1171: of a highly relativistic EMRB and found that it whirls more than once
1172: around the MBH before zooming back to apocenter and becoming silent
1173: again.
1174: 
1175: We have also discussed the possible consequences that relativistic
1176: effects might have on the detection and parameter estimation of
1177: gravitational waves from EMRBs by LISA, namely a change in the SNR and
1178: loss of overlap. These changes are mainly due to the relativistic
1179: treatment of the equations of motion, while a quadrupolar wave
1180: generation formalism seems to suffice. This finding is relevant
1181: particularly to match filtering searches, where a Newtonian treatment
1182: of the orbit might lead to a deterioration of confidence limits.
1183: Furthermore, our study suggests that, given an EMRB gravitational wave
1184: detection, it might be plausible to extract or bound the spin of the
1185: central potential with a template that takes into account the Kerr
1186: character of the MBH. Other astrophysical consequences include a
1187: possible increase in the event rate, which implies that the rates of
1188: \citet{Rubbo:2006dv,Rubbo:2007} and \citet{Hopman:2006fc} might be
1189: lower limits, although these estimates are still dominated by
1190: uncertainties in the astrophysical modeling for the host galaxy.
1191: 
1192: In addition to the astrophysical modeling, future research should
1193: tackle the details of the EMRB data analysis and signal extraction
1194: issues put forward above. Based on the results of this paper, one may
1195: explore the possibility of testing alternative theories of gravity
1196: with EMRBs by performing matched filtering with templates from
1197: alternative theories \citep{will:1998:bmo, scharre:2002:tsg,
1198:   will:2004:tat, berti:2005:tgr, berti:2005:esb}.  Another possible
1199: avenue for future research is the study of confidence limits with
1200: which the spin of the central MBH can be measured.
1201: 
1202: This research could then be used to examine whether EMRB events can
1203: distinguish between a pure Kerr MBH and some other spacetime. Such
1204: studies have already began with the analysis
1205: of~\citet{Collins:2004ex},~\citet{Glampedakis:2005cf}
1206: and~\citet{Barausse:2006vt}, where comparisons between a Kerr and
1207: other non-Kerr spacetimes were performed. Such studies could be
1208: extended to the perturbed Kerr solution found by~\citet{Yunes:2005ve},
1209: where the perturbation is parameterized by the Weyl tensor of the
1210: external universe and could represent some external accretion disk,
1211: planetary system or some other compact object. Ultimately, these
1212: investigations will decide whether EMRB events are worth studying in
1213: further detail by future gravitational wave observatories.
1214: 
1215: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1216: %%%%%%%%%%%%%%%%%%%  ACKNOWLEDGMENTS %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1217: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1218: 
1219: \acknowledgments
1220: 
1221: The authors would like to thank Ben Owen for reading the manuscript
1222: and providing useful comments, as well as the anonymous referee for
1223: providing insightful suggestions. We would also like to acknowledge
1224: the support of the Center for Gravitational Wave Physics funded by the
1225: National Science Foundation under Cooperative Agreement PHY-01-14375,
1226: and support from NSF grants PHY 05-55628, PHY 05-55436, PHY 02-18750,
1227: PHY 02-44788, PHY 02-45649 and PHY 00-99559.  K.~H.~B.~and L.~R.~also
1228: acknowledge the support of NASA NNG04GU99G, NASA NN G05GF71G.
1229: C.~F.~S. acknowledges the support of the Natural Sciences and
1230: Engineering Research Council of Canada.
1231: 
1232: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1233: %%%%%%%%%%%%%%%%%%%  BIBLIOGRAPHY  %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1234: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1235: % Create the reference section using BibTeX:
1236: \begin{thebibliography}{48}
1237: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
1238: 
1239: \bibitem[{Amaro-Seoane {et~al.}(2007)}]{Amaro-Seoane:2007aw}
1240: Amaro-Seoane, P., {et~al.} 2007, ArXiv e-print astro-ph/0703495
1241: 
1242: \bibitem[{Anderson {et~al.}(2001)Anderson, Brady, Creighton, \&
1243:   Flanagan}]{Anderson:2000yy}
1244: Anderson, W.~G., Brady, P.~R., Creighton, J. D.~E., \& Flanagan, E.~E. 2001,
1245:   Phys. Rev., D63, 042003
1246: 
1247: \bibitem[{Babak {et~al.}(2007)Babak, Fang, Gair, Glampedakis, \&
1248:   Hughes}]{Babak:2006uv}
1249: Babak, S., Fang, H., Gair, J.~R., Glampedakis, K., \& Hughes, S.~A. 2007, Phys.
1250:   Rev., D75, 024005
1251: 
1252: \bibitem[{Barack \& Cutler(2004{\natexlab{a}})}]{Barack:2004b}
1253: Barack, L., \& Cutler, C. 2004{\natexlab{a}}, Phys. Rev. D, 70, 122002
1254: 
1255: \bibitem[{Barack \& Cutler(2004{\natexlab{b}})}]{Barack:2004a}
1256: ---. 2004{\natexlab{b}}, Phys. Rev. D, 69, 082005
1257: 
1258: \bibitem[{Barausse {et~al.}(2007)Barausse, Rezzolla, Petroff, \&
1259:   Ansorg}]{Barausse:2006vt}
1260: Barausse, E., Rezzolla, L., Petroff, D., \& Ansorg, M. 2007, Phys. Rev., D75,
1261:   064026
1262: 
1263: \bibitem[{Bender {et~al.}(1998)}]{Bender:1998}
1264: Bender, P., {et~al.} 1998, {Laser Interferometer Space Antenna for the
1265:   Detection and Observation of Gravitational Waves: An International Project in
1266:   the Field of Fundamental Physics in Space}, {LISA Pre-Phase A Report},
1267:   Max-Planck-Institut f{\"u}r Quantenoptik, Garching, mPQ~233
1268: 
1269: \bibitem[{{Berti} {et~al.}(2005{\natexlab{a}}){Berti}, {Buonanno}, \&
1270:   {Will}}]{berti:2005:esb}
1271: {Berti}, E., {Buonanno}, A., \& {Will}, C.~M. 2005{\natexlab{a}}, Phys. Rev. D,
1272:   71
1273: 
1274: \bibitem[{{Berti} {et~al.}(2005{\natexlab{b}}){Berti}, {Buonanno}, \&
1275:   {Will}}]{berti:2005:tgr}
1276: ---. 2005{\natexlab{b}}, Class. Quant. Grav., 22, S943
1277: 
1278: \bibitem[{Blanchet(2006)}]{Blanchet:2002av}
1279: Blanchet, L. 2006, Living Rev. Relativity, 9, 4
1280: 
1281: \bibitem[{Camarda \& Ortolan(2006)}]{Camarda:2006}
1282: Camarda, M., \& Ortolan, A. 2006, Phys. Rev. D, 74, 062001
1283: 
1284: \bibitem[{{Chandrasekhar}(1992)}]{Chandrasekhar:1992bo}
1285: {Chandrasekhar}, S. 1992, {The mathematical theory of black holes} (New York:
1286:   Oxford University Press)
1287: 
1288: \bibitem[{Collins \& Hughes(2004)}]{Collins:2004ex}
1289: Collins, N.~A., \& Hughes, S.~A. 2004, Phys. Rev., D69, 124022
1290: 
1291: \bibitem[{Danzmann \& R{\"u}diger(2003)}]{Danzmann:2003tv}
1292: Danzmann, K., \& R{\"u}diger, A. 2003, Class. Quant. Grav., 20, S1
1293: 
1294: \bibitem[{Eisenhauer {et~al.}(2005)}]{Eisenhauer:2005}
1295: Eisenhauer, F., {et~al.} 2005, ApJ, 628, 246
1296: 
1297: \bibitem[{Flanagan \& Hughes(2005)}]{Flanagan:2005yc}
1298: Flanagan, E.~E., \& Hughes, S.~A. 2005, New J. Phys., 7, 204
1299: 
1300: \bibitem[{Gair {et~al.}(2005)Gair, Kennefick, \& Larson}]{Gair:2005is}
1301: Gair, J.~R., Kennefick, D.~J., \& Larson, S.~L. 2005, Phys. Rev. D, 72, 084009
1302: 
1303: \bibitem[{Gair {et~al.}(2006)Gair, Kennefick, \& Larson}]{Gair:2005hu}
1304: ---. 2006, ApJ, 639, 999
1305: 
1306: \bibitem[{Gair {et~al.}(2004)}]{Gair:2004iv}
1307: Gair, J.~R., {et~al.} 2004, Class. Quant. Grav., 21, S1595
1308: 
1309: \bibitem[{Ghez {et~al.}(2005)}]{Ghez:2005}
1310: Ghez, A.~M., {et~al.} 2005, ApJ, 620, 744
1311: 
1312: \bibitem[{Glampedakis(2005)}]{Glampedakis:2005hs}
1313: Glampedakis, K. 2005, Class. Quant. Grav., 22, S605
1314: 
1315: \bibitem[{Glampedakis \& Babak(2006)}]{Glampedakis:2005cf}
1316: Glampedakis, K., \& Babak, S. 2006, Class. Quant. Grav., 23, 4167
1317: 
1318: \bibitem[{Glampedakis \& Kennefick(2002)}]{Glampedakis:2002}
1319: Glampedakis, K., \& Kennefick, D. 2002, Phys. Rev. D, 66, 044002
1320: 
1321: \bibitem[{{Holley-Bockelmann} \& {Sigurdsson}(2006)}]{HB:2006af}
1322: {Holley-Bockelmann}, K., \& {Sigurdsson}, S. 2006, ArXiv e-print
1323:   astro-ph/0601520
1324: 
1325: \bibitem[{{Hopman} \& {Alexander}(2006)}]{Hopman:2006ha}
1326: {Hopman}, C., \& {Alexander}, T. 2006, ApJ, 645, L133
1327: 
1328: \bibitem[{Hopman {et~al.}(2006)Hopman, Freitag, \& Larson}]{Hopman:2006fc}
1329: Hopman, C., Freitag, M., \& Larson, S.~L. 2006, ArXiv e-print astro-ph/0612337
1330: 
1331: \bibitem[{Hughes(2001{\natexlab{a}})}]{Hughes:2001:a}
1332: Hughes, S.~A. 2001{\natexlab{a}}, Phys. Rev. D, 64, 064004
1333: 
1334: \bibitem[{Hughes(2001{\natexlab{b}})}]{Hughes:2001:b}
1335: ---. 2001{\natexlab{b}}, Class. Quant. Grav., 18, 4067
1336: 
1337: \bibitem[{Klimenko {et~al.}(2004)Klimenko, Yakushin, \&
1338:   Mitselmakher}]{Klimenko:2004}
1339: Klimenko, S., Yakushin, I., \& Mitselmakher, G. 2004, Class. Quant. Grav., 21,
1340:   S1685
1341: 
1342: \bibitem[{Kraniotis(2007)}]{Kraniotis:2007}
1343: Kraniotis, G.~V. 2007, Class. Quant. Grav., 24, 1775
1344: 
1345: \bibitem[{Larson {et~al.}(2000)Larson, Hiscock, \& Hellings}]{Larson:2000}
1346: Larson, S.~L., Hiscock, W.~A., \& Hellings, R.~W. 2000, Phys. Rev. D, 62,
1347:   062001,
1348:   \texttt{http://www.srl.caltech.edu/$\sim$shane/sensitivity/index.html}
1349: 
1350: \bibitem[{Marck(1996)}]{Marck:1995kd}
1351: Marck, J.-A. 1996, Class. Quant. Grav., 13, 393
1352: 
1353: \bibitem[{Misner {et~al.}(1973)Misner, Thorne, \& Wheeler}]{Misner:1973cw}
1354: Misner, C.~W., Thorne, K., \& Wheeler, J.~A. 1973, Gravitation (San Francisco:
1355:   W. H. Freeman \& Co.)
1356: 
1357: \bibitem[{Poisson(2004)}]{Poisson:2004lr}
1358: Poisson, E. 2004, Living Rev. Relativity, 7, 6
1359: 
1360: \bibitem[{Press {et~al.}(1992)Press, Flannery, Teukolsky, \&
1361:   Vetterling}]{Press:1992nr}
1362: Press, W.~H., Flannery, B.~P., Teukolsky, S.~A., \& Vetterling, W.~T. 1992,
1363:   Numerical Recipes: The Art of Scientific Computing (Cambridge (UK) and New
1364:   York: Cambridge University Press)
1365: 
1366: \bibitem[{Rubbo {et~al.}(2006)Rubbo, Holley-Bockelmann, \& Finn}]{Rubbo:2006dv}
1367: Rubbo, L.~J., Holley-Bockelmann, K., \& Finn, L.~S. 2006, ApJ, 649, L25
1368: 
1369: \bibitem[{Rubbo {et~al.}(2007)Rubbo, Holley-Bockelmann, \& Finn}]{Rubbo:2007}
1370: Rubbo, L.~J., Holley-Bockelmann, K., \& Finn, L.~S. 2007, in Laser
1371:   Interferometer Space Antenna: 6th International LISA Symposium, ed. S.~M.
1372:   Merkowitz \& J.~C. Livas, Vol. 873 (AIP Conference Proceedings), 284--288
1373: 
1374: \bibitem[{Ruffini \& Sasaki(1981)}]{Ruffini:1981rs}
1375: Ruffini, R., \& Sasaki, M. 1981, Prog. Theor. Phys., 66, 1627
1376: 
1377: \bibitem[{{Scharre} \& {Will}(2002)}]{scharre:2002:tsg}
1378: {Scharre}, P.~D., \& {Will}, C.~M. 2002, Phys. Rev. D, 65, 042002
1379: 
1380: \bibitem[{Schmidt(2002)}]{Schmidt:2002}
1381: Schmidt, W. 2002, Class. Quant. Grav., 19, 2743
1382: 
1383: \bibitem[{Stoer \& Bulirsch(1993)}]{Stoer:1993sb}
1384: Stoer, J., \& Bulirsch, R. 1993, Introduction to Numerical Analysis (New York:
1385:   Springer-Verlag)
1386: 
1387: \bibitem[{Stuver \& Finn(2006)}]{Stuver:2006}
1388: Stuver, A.~L., \& Finn, L.~S. 2006, Class. Quant. Grav., 23, S733
1389: 
1390: \bibitem[{Sumner \& Shaul(2004)}]{Sumner:2004}
1391: Sumner, T.~J., \& Shaul, D. N.~A. 2004, MPLA, 19, 785
1392: 
1393: \bibitem[{Thorne(1980)}]{Thorne:1980rm}
1394: Thorne, K.~S. 1980, Rev. Mod. Phys., 52, 299
1395: 
1396: \bibitem[{Tremaine {et~al.}(1994)}]{Tremaine:1994}
1397: Tremaine, S., {et~al.} 1994, AJ, 107, 634
1398: 
1399: \bibitem[{Will(1998)}]{will:1998:bmo}
1400: Will, C.~M. 1998, Phys. Rev. D, 57, 2061
1401: 
1402: \bibitem[{Will \& Yunes(2004)}]{will:2004:tat}
1403: Will, C.~M., \& Yunes, N. 2004, Class. Quant. Grav., 21, 4367
1404: 
1405: \bibitem[{Yunes \& Gonzalez(2006)}]{Yunes:2005ve}
1406: Yunes, N., \& Gonzalez, J.~A. 2006, Phys. Rev., D73, 024010
1407: 
1408: \end{thebibliography}
1409: 
1410: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1411: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%  TABLES  %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1412: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1413: \clearpage 
1414: %
1415: \begin{center}
1416: \begin{deluxetable}{c c c c c c c c c}
1417:   \tablewidth{0pt} \tablecaption{\label{table_orbs} SNR and overlap for
1418:     different EMRB orbits.}  \tablehead{ \colhead{$r_p$ [$\mu$pc]} &
1419:     \colhead{$v_p/c$} & \colhead{$P$ [yrs]} &
1420:     \colhead{$e$} & \colhead{$\#$ Bursts} & \colhead{$\frac{|\Delta
1421:     \rho|}{\rho}$} & \colhead{$\frac{|\Delta
1422:     \rho^{(1)}|}{\rho^{(1)}}$} & \colhead{$(h_1|h_2)$} 
1423:     & \colhead{$(h_1|h_2)^{(1)}$} }  
1424: %
1425: \startdata 
1426: %
1427: $7.8$ & $0.196$ & $0.0042$ & $0.603$ & $6$ & $0.38$ & $0.049$ & $0.27$ & $0.98$\\[1mm]
1428: %
1429: $7.5$ & $0.207$ & $0.0055$ & $0.684$ & $5$ & $0.13$ & $0.0007$ & $0.20$ & $0.88$\\[1mm]
1430: %
1431: $7.2$ & $0.217$ & $0.0074$ & $0.749$ & $4$ & $0.22$ & $0.057$ & $0.31$ & $0.90$\\[1mm]
1432: %
1433: $7.1$ & $0.220$ & $0.0086$ & $0.777$ & $3$ & $0.49$ & $0.061$ & $0.33$ & $0.99$\\[1mm]
1434: %
1435:       &                      &          &         &     &     \\[1mm]
1436: %
1437: $13$  & $0.163$ & $0.0450$ & $0.859$ & $1$ & $\cdot$ & $0.002$ & $\cdot$ & $0.90$\\[1mm]
1438:                                 %
1439: $8.0$ & $0.217$ & $0.1953$ & $0.968$ & $1$ & $\cdot$ & $0.033$ & $\cdot$ & $0.97$\\[1mm]
1440: %
1441: $10$  & $0.193$ & $0.7683$ & $0.984$ & $1$ & $\cdot$ & $0.009$ & $\cdot$ & $0.72$\\[1mm]
1442: %
1443: $12$  & $0.173$ & $3.0407$ & $0.992$ & $1$ & $\cdot$ & $0.015$ & $\cdot$ & $0.80$\\[1mm]
1444: %
1445: \enddata \tablecomments{In this table we present the SNR and
1446:   correlation computed in the frequency domain between Kerr quadrupole
1447:   and octopole waveforms for different EMRB orbits. The orbits are
1448:   separated into two groups: highly-bursting (top) and slowly-bursting
1449:   (bottom). The first five columns present information about the
1450:   different orbits, including how many times they burst, their
1451:   eccentricities and periods, which were chosen directly from the
1452:   allowed EMRB phase space of~\citet{Rubbo:2006dv} and thus represent
1453:   Milky Way sources. The sixth and seventh columns present the
1454:   difference in SNR between a Newtonian and Kerr quadrupole waveform
1455:   relative to the former using the entire data set and only one burst
1456:   respectively. Similarly, the eighth and nineth columns show the
1457:   correlation between the plus polarizations using the entire data set
1458:   four and one burst respectively. Since the slowly-bursting sources
1459:   burst only once, the sixth and eighth columns are redundant for
1460:   these sources. All calculations assume the observer is located on
1461:   the $z$-axis and random initial inclination angles.}
1462: \end{deluxetable}
1463: \end{center}
1464: %
1465: 
1466: 
1467: \clearpage 
1468: %
1469: \begin{center}
1470: \begin{deluxetable}{c c c c c}
1471: \tablewidth{0pt}
1472: \tablecaption{\label{table_obs} Comparison between quadrupole and
1473:   octopole waveforms.}
1474: \tablehead{
1475: \colhead{Angle [degrees]} & \colhead{$x_{\textrm{obs}}$ [kpc]} &
1476: \colhead{$y_{\textrm{obs}}$ [kpc]} & \colhead{$z_{\textrm{obs}}$ [kpc]}
1477: & \colhead{Amp.~diff.}} 
1478: \startdata
1479: $0$  & $0$  & $0$ & $8$ & $3.9 \%$ \\[1mm]
1480: $20$ & $2.73$  & $0$ & $7.51$ & $3.3 \%$ \\[1mm]
1481: $40$ & $5.14$  & $0$ & $6.12$ & $11.5 \%$ \\[1mm]
1482: $60$ & $6.93$  & $0$ & $4$    & $15.2 \%$ \\[1mm]
1483: $80$ & $7.88$  & $0$ & $1.39$ & $19.7 \%$ \\[1mm]
1484: $100$ & $7.88$  & $0$ & $-1.39$ & $21.6 \%$ \\[1mm]
1485: $120$ & $6.93$  & $0$ & $-4$ & $17.8 \%$ \\[1mm]
1486: $140$ & $5.14$  & $0$ & $-6.13$ & $43.3 \%$ \\[1mm]
1487: $160$ & $2.74$  & $0$ & $-7.52$ & $8.6 \%$ \\[1mm]
1488: \enddata
1489: \tablecomments{Here we present an approximate measure of the amplitude
1490:   difference between the quadrupole and octopole waveforms. We
1491:   concentrate only on the first burst of radiation and we normalize
1492:   the difference to the maximum of the first peak of the quadrupole
1493:   waveform. The difference is presented as a function of the observer
1494:   location, which is always at a fixed radial distance of
1495:   $r_{\textrm{obs}} = 8$ kpc, but rotated about the $\hat{y}$-axis on
1496:   $\hat{x}-\hat{z}$ plane ($\theta$ is here the usual Euler polar
1497:   angle.) }
1498: \end{deluxetable}
1499: \end{center}
1500: %
1501: 
1502: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1503: %%%%%%%%%%%%%%%%%%%  THAT'S ALL FOLKS!  %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1504: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1505: \end{document}
1506: