0710.3795/ms.tex
1: %\documentclass{emulateapj}
2: \documentclass[10pt,preprint]{aastex}
3: \usepackage{amsmath}
4: \pdfoutput=1
5: 
6: \shorttitle{Localizing coalescing black hole binaries}
7: \shortauthors{Lang \& Hughes}
8: 
9: \begin{document}
10: 
11: \title{Localizing coalescing massive black hole binaries with
12: gravitational waves}
13: 
14: \author{Ryan N.\ Lang and Scott A.\ Hughes}
15: 
16: \affil{Department of Physics and Kavli Institute for Astrophysics and Space Research, Massachusetts Institute of Technology, 77 Massachusetts Avenue, Cambridge, MA 02139}
17: 
18: \begin{abstract}
19: Massive black hole binary coalescences are prime targets for
20: space-based gravitational wave (GW) observatories such as {\it LISA}.  
21: GW measurements can localize the position of a
22: coalescing binary on the sky to an ellipse with a major axis of
23: a few tens of arcminutes to a few degrees, depending on 
24: source redshift, and a minor axis which is $2 - 4$
25: times smaller.  Neglecting weak gravitational lensing,
26: the GWs would also determine the source's luminosity distance to
27: better than percent accuracy for close sources, degrading to several
28: percent for more distant sources.  Weak lensing cannot, in fact, be
29: neglected and is expected to limit the accuracy with which distances
30: can be fixed to errors no less than a few percent.  Assuming a
31: well-measured cosmology, the source's redshift could be inferred with
32: similar accuracy.  GWs alone can thus pinpoint a binary to a
33: three-dimensional ``pixel'' which can help guide searches for the
34: hosts of these events.  We examine the time evolution of this pixel, studying it at merger and at several intervals before
35: merger.  One day before merger, the major axis of the error ellipse 
36: is typically larger
37: than its final value by a factor of $\sim 1.5-6$.  The
38: minor axis is larger by a factor of $\sim 2-9$, and, neglecting lensing, the error
39: in the luminosity distance is larger by a factor of $\sim 1.5-7$.  This
40: large change over a short period of time is due to spin-induced precession, 
41: which is strongest in the final days before
42: merger.  The evolution is slower as we go back further in time.  
43: For $z = 1$, we find that GWs will localize a
44: coalescing binary to within $\sim 10\ \mathrm{deg}^2$ as
45: early as a month prior to merger and determine distance (and hence
46: redshift) to several percent.
47: \end{abstract}
48: 
49: \keywords{black hole physics --- galaxies: nuclei --- gravitation --- gravitational waves}
50: 
51: \section{Introduction}
52: \label{sec:intro}
53: 
54: Among the most important sources of gravitational waves (GWs) in the
55: low-frequency band of space-based detectors are the coalescences of
56: massive black hole binaries (MBHBs).  Binaries containing black holes
57: with masses in the range $10^4 - 10^7\,M_\odot$ are predicted to form
58: through the hierarchical growth of structure as dark matter halos (and
59: the galaxies they host) repeatedly merge; see, for example,
60: {\citet{svh07}} and {\citet{mhsa07}} for recent discussion.  The {\it Laser
61: Interferometer Space Antenna} ({\it LISA}) is being designed to have a
62: sensitivity that would allow detailed measurement of the waves from
63: these binaries.  ``Intrinsic'' parameters --- the masses and spins of
64: the black holes which compose the binary --- should be determined with
65: very high accuracy, with relative errors typically $\sim 10^{-3}$ to 
66: $10^{-1}$, depending on system mass and redshift; see Lang \& Hughes
67: (2006, hereafter Paper I) for recent discussion.  By measuring an
68: ensemble of coalescences over a range of redshifts, MBHB GWs may serve
69: as a kind of structure tracer, tracking the growth and spin
70: evolution of black holes over cosmic time.
71: 
72: ``Extrinsic'' system parameters, describing a binary's location and
73: orientation relative to the detector, are also determined by measuring
74: its GWs.  In Paper I, we showed that a binary's position on the sky
75: can be localized at $z = 1$ to an ellipse with a major axis of a few
76: tens of arcminutes and a minor axis a factor of $2-4$ smaller.  At
77: higher redshift ($z = 3-5$), these values degrade by a factor of a
78: few, reaching a few degrees in the long direction and tens of
79: arcminutes to a degree or two in the short one.  We also found that,
80: neglecting errors due to weak gravitational lensing, a source's
81: luminosity distance typically can be determined to better $1\%$
82: at low redshift ($z = 1$), degrading to several percent at
83: higher redshift ($z = 3-5$).
84: 
85: The intrinsic ability of GWs to determine the distance to a coalescing
86: binary is phenomenal.  Coalescing MBHB systems constitute exquisitely
87: well calibrated distance measures, with the calibration provided by
88: general relativity.  Unfortunately, this percent-level or better
89: accuracy could only be achieved if we measured MBHB coalescences in an
90: empty universe.  In our universe, weak lensing will magnify or
91: demagnify the GWs, and we will infer a luminosity distance smaller
92: (for magnification) or larger (for demagnification) than the true
93: value.  This phenomenon affects all high-redshift standard candles.  Its
94: impact on Type Ia supernovae in particular has been discussed in
95: detail {\citep{frieman97,hw98,h98}}.  It will not be possible to
96: correct for this effect {\citep{dhcf03}} since much of the lensing
97: ``noise'' arises from structure on subarcminute scales that is not
98: probed by shear maps (which map the distribution of matter on scales
99: greater than $1^\prime$ or so).  Since we will not know the extent
100: of the magnification when we measure MBHB waves, we must simply accept
101: the fact that lensing introduces a dispersion of several percent in determining
102: the distance to these GW events (see, e.g., Wang et al. [2002] to
103: compute this dispersion as a function of redshift).  When we quote
104: distance measurement errors, we will typically quote only the
105: intrinsic GW measurement error, neglecting lensing's impact.  When
106: the intrinsic GW distance error is $\lesssim 5\%$, lensing will blur it
107:  to the several percent level.
108: 
109: Note that a source's redshift $z$ {\it cannot} be directly determined
110: using only GWs.  Gravitational wave measurements infer system parameters 
111: through their
112: impact on certain dynamical timescales, such as orbital frequencies
113: and the rate at which these frequencies evolve.  Since these time
114: scales all suffer cosmological redshift, $z$ is degenerate with other
115: parameters.  For example, any mass parameter $m$ is actually measured
116: as $(1 + z)m$.  However, if the binary's luminosity distance is
117: determined, its redshift can then be inferred by assuming a
118: cosmography.  For most binaries, the redshift can be determined to
119: several percent (with an error budget dominated by gravitational
120: lensing\footnote{At redshifts $z \lesssim 0.3$, the error is actually
121: dominated by peculiar velocity effects {\citep{kfhm06}}; however, the
122: event rate is probably negligible at such low redshifts.  As such, we
123: will focus on gravitational lensing as the main source of systematic
124: redshift error.}).  We thus expect that GW measurements will locate a
125: binary to within a three-dimensional ``GW pixel'' which at $z = 1$ has a
126: cross-sectional area of $\sim 10^{-2}$ to $10^{-1}\ \mathrm{deg}^2$ and a
127: depth $\Delta z/z \sim $ several percent.
128: 
129: It is anticipated that there will be great interest in searching the
130: GW pixel for electromagnetic (e.g., optical, X-ray, radio)
131: counterparts to MBHB GW events.  Finding such a counterpart would be
132: much easier if galactic activity were catalyzed in association with the
133: coalescence {\citep{kfhm06}}.  The nature of that activity is likely
134: to depend rather strongly on the mass of the coalescing system
135: \citep{dssch06}.  For example, {\citet{an02}} predict strong outflows
136: and galactic activity prior to the final black hole merger as the
137: smaller member of the binary drives gas onto the larger member,
138: consistent with the high-mass ($M_{\rm tot} \gtrsim 10^7\,M_\odot$)
139: predictions of \citet{dssch06}.  {\citet{mp05}} describe an X-ray
140: afterglow that would ignite after gas refills the volume that is swept
141: clean by the coalescing binary; Dotti et al.\ predict this outcome for
142: smaller systems ($M_{\rm tot} \lesssim \mbox{several}\times
143: 10^6\,M_\odot$).  Recent work by {\citet{bp07}} suggests that the
144: final burst of radiation from a coalescing binary (which can convert
145: $\sim 10\%$ of the system's mass to GWs very suddenly) may excite
146: radial waves, and consequently electromagnetic variability, in an accretion 
147: disk due to the quick change in the disk's Keplerian potential.
148: Such a signature may be essentially mass independent.  On the other
149: hand, the coalescence may be electromagnetically quiet, in which case
150: we face the potentially daunting task of searching the three-dimensional
151: pixel for galaxies with morphology consistent with a (relatively)
152: recent merger, or that have a central velocity dispersion $\sigma$
153: consistent with the inferred final black hole mass (assuming that the
154: $M_{\rm BH}-\sigma$ relation [Ferrarese \& Merritt 2000; Gebhardt et al. 2000] holds at the redshift
155: of these sources, and so soon after merger).
156: 
157: If the host galaxy or some other counterpart can be identified, we
158: could then contemplate combining GW information with electromagnetic
159: data.  For instance, combining {\it LISA} mass measurements with the
160: luminosity of the counterpart may allow us to directly measure the
161: Eddington ratio $L/L_{\mathrm{Edd}}$ {\citep{kfhm06}}.  Identifying a
162: counterpart would also allow us to more accurately characterize the
163: system.  For example, much of the intrinsic luminosity distance error
164: is due to correlations between distance and sky position.  Finding an
165: electromagnetic counterpart essentially determines a binary's position
166: precisely, breaking those correlations.  Previous studies have found
167: that intrinsic distance error can be reduced by almost an order of
168: magnitude if the position is known {\citep{h02,hh05}}.  (Lensing
169: errors still dominate in such a case, so that the distance remains
170: determined only at the few percent level.)  A counterpart may also
171: make it possible to simultaneously determine a source's luminosity
172: distance and redshift.  Such a ``standard siren'' (the GW analog of a
173: standard candle) would very usefully complement other high-redshift
174: standard candles {\citep{hh05}}, such as Type Ia supernovae
175: {\citep{p93, rpk95, wgap03}}.  A direct measurement of redshift will
176: also break the mass-redshift degeneracy more accurately than can be
177: done with just the luminosity distance and some assumed cosmological
178: parameters.  Breaking this degeneracy is critical when studying the
179: growth of black holes with cosmic time {\citep{h02}}.
180: 
181: Many analyses {\citep{c98,h02,v04,bbw05,hh05}} have quantified how
182: well {\it LISA} can determine MBHB parameters, including sky position
183: and distance, using maximum likelihood Fisher matrix estimation. Our
184: results from Paper I, given earlier, include the effects of
185: ``spin-induced precession'' --- precession of both the orbital plane
186: and the individual spins of the black holes due to post-Newtonian
187: spin-interaction effects.  A significant result from that analysis is
188: that spin-induced precession improves sky position accuracy by about
189: half an order of magnitude in each direction versus previous analyses.
190: This result can be partially understood as due to the breaking of a
191: degeneracy between position and orientation angles: thanks to
192: precession, the binary's orientation evolves with time and can be
193: untangled from sky position.  This effect was already known due to
194: pioneering work by {\citet{v04}}; by taking the analysis to higher
195: order and considering a broader range of sources, we were able to show
196: that this improvement held for essentially all astrophysically
197: interesting MBHB sources.
198: 
199: The purpose of this paper is to examine the localization of MBHB
200: systems more thoroughly, in particular how the GW pixel evolves as the
201: final merger is approached.  Paper I only presented results for
202: measurements that proceed all the way to merger.  It will clearly be
203: of some interest to monitor potential hosts for the binary event some
204: time before the merger happens; if nothing else, telescopes will need
205: prior warning to schedule observing campaigns.  Understanding the rate
206: at which localization evolves can also have an important impact on the
207: design of the {\it LISA} mission, clarifying how often it will be
208: necessary to downlink data about MBHB systems in order to effectively
209: guide surveys.
210: 
211: Our main goal is to understand for what range of masses and redshifts prior
212: localization of a binary using GWs will be possible.  A previous
213: analysis by Kocsis et al.\ (2007b; hereafter K07) also examined this
214: problem in great detail, but without including the impact of
215: spin-induced precession.  One of our goals is to see to what extent
216: precession physics changes the conclusions of K07.  We find that
217: precession has a fairly small impact on the time evolution of the GW
218: pixel except in the last few days before the final merger, at which
219: point its impact can be tremendous.  Precession typically changes the
220: area of the sky position error ellipse by a factor of $\sim 3-10$ (up to
221: $\sim 60$ in extreme cases) in just the final day.  This is in accord
222: with the predictions of K07 (and even earlier predictions by N. Cornish
223: 2005, unpublished).
224: 
225: The structure of this paper is as follows.  First, in \S\
226: {\ref{sec:background}}, we briefly review the basics of the MBHB
227: gravitational waveform and the parameter estimation formalism that we
228: use; this section is essentially a synopsis of relevant material from
229: Paper I.  Section {\ref{sec:intrinsicGW}} reviews the form of the GWs
230: that we use in our analysis, while \S\ {\ref{sec:extrinsicGW}}
231: describes how those waves are measured by the {\it LISA}
232: constellation.  We describe the measurement formalism we use in
233: \S\ {\ref{sec:formalism}}.  In \S\ {\ref{sec:review}}, we
234: summarize our results from Paper I regarding the final localization
235: accuracy that {\it LISA} can expect to achieve.
236: 
237: We turn to a detailed discussion of the time evolution of the GW pixel
238: in \S\ {\ref{sec:results}}.  We begin by summarizing the key ideas
239: behind the ``harmonic mode decomposition'' of K07 in \S\
240: {\ref{sec:khmf}}.  This technique cleverly allows calculation of the
241: GW pixel and its time evolution with much less computational effort
242: than our method (albeit without including the impact of spin
243: precession).  Unfortunately, we have discovered that some of the
244: approximations used by K07 introduce a systematic underestimate of
245: the final sky position error by a factor of $2 - 4$ or more in angle;
246: the approximations are much more reliable a week or more prior to the
247: black holes' final merger.  Modulo this underestimate, the K07
248: results agree well with a version of our code which does not include
249: spin precession (particularly a week or more in advance of merger,
250: when their underestimate is not severe).  K07 thus serves as a useful
251: point of comparison to establish the impact of precession on source
252: localization.
253: 
254: Section {\ref{sec:timeevolve}} is dedicated to our results, including
255: comparison to K07 when appropriate.  We find that all relevant
256: parameter errors decrease slowly with time until the last day before
257: merger, when they drop more dramatically.  This sudden drop is not
258: found in K07, nor is it present in a variant of our analysis that
259: ignores spin precession.  It clearly can be attributed to the impact
260: of precession on the waveform.  Before this last day, the major axis
261: is $\sim 1.5-6$ times, the minor axis $\sim 2-9$ times, 
262: and the intrinsic error in
263: the luminosity distance $D_L$ $\sim 1.5-7$ times bigger than at merger
264: for most binaries (i.e., all except the highest masses).  Going back
265: to one week (one month) before merger, these numbers change to $2-9$
266: ($4-11$) for the major axis, $3-14$ ($5-24$) for the minor axis, and
267: $3-14$ ($5-18$) for the error in the luminosity distance.  As a
268: result, for $z = 1$, most binaries can be located within a few square
269: degrees a week before merger and $10\ \mathrm{deg}^2$ a month before
270: merger.  The intrinsic distance errors are also small enough this
271: early that $\Delta z/z$ remains dominated by gravitational lensing
272: errors of several percent.  Advanced localization of MBHB coalescences
273: thus seems plausible for these binaries; the situation is less
274: promising for sources at higher redshift.
275: 
276: As a corollary to our study of the time evolution, we also examine the
277: sky position dependence of errors (\S\ {\ref{sec:angdependence}}).
278: The errors depend strongly on the polar angle with respect to the
279: ecliptic, increasing in the ecliptic plane to as much as $35\%$ over
280: the median for the major axis, $85\%$ over the median for the minor
281: axis, and $15\%$ over the median for errors in the luminosity
282: distance.  The errors have a much weaker dependence on the azimuthal
283: angle.  When we convert to Galactic coordinates, we find that the best
284: localization regions appear to lie fairly far out of the Galactic
285: plane, offering hope that searches for counterparts will not be too badly
286: impacted by foreground contamination.
287: 
288: We conclude this paper in \S\ {\ref{sec:disc}}.  Besides summarizing
289: our results, we discuss shortcomings of this analysis and future work
290: which could help to better understand how well GWs can localize MBHB
291: sources.
292: 
293: Throughout the paper we set $G = c = 1$; a convenient conversion
294: factor in this system is $10^6 M_\odot = 4.92$ s.  When
295: discussing results, we always quote masses as they would be measured
296: in the rest frame of the source.  These masses must be multiplied by a
297: factor of $1 + z$ when used in any of the equations describing the
298: system's dynamics or its GWs (particularly the equations of \S\
299: {\ref{sec:background}}).  We convert between distance and redshift
300: using a spatially flat cosmology with $\Omega_\Lambda = 0.75$ (and
301: hence $\Omega_m = 0.25$) and Hubble constant $H_0 = 75\ {\rm
302: km}\ {\rm s}^{-1}\ {\rm Mpc}^{-1}$.
303: 
304: \section{Determining the sky position and distance accuracy of {\it LISA}}
305: \label{sec:background}
306: 
307: In this section, we present necessary background.  We begin by summarizing the
308: inspiral GW signal and its measurement by {\it LISA}.  We then
309: describe the parameter estimation formalism that we use in our code.
310: More details on these topics can be found in Paper I and references
311: therein.  Finally, we present a synopsis of the results from Paper I,
312: summarizing the {\it final} sky position and distance accuracy that
313: {\it LISA} can expect to achieve.  Detailed discussion of how this
314: accuracy evolves as a function of time is deferred to \S\
315: {\ref{sec:results}}.
316: 
317: \subsection{MBHB gravitational waves}
318: \label{sec:intrinsicGW}
319: 
320: As is by now traditional, we break the GWs generated by a coalescing
321: binary black hole system into three more or less distinct epochs: (1)
322: a slowly evolving {\it inspiral}, in which the black holes gradually
323: spiral toward each other as the orbit decays due to GW emission; (2)
324: a loud {\it merger}, in which the black holes come together and
325: form a single body; and (3) a {\it ringdown}, in which the merged
326: remnant of the binary settles down to its final state.  Our analysis
327: focuses on the inspiral, the most long-lived epoch of coalescence and
328: the epoch in which spin precession plays a major dynamical role.  The
329: inspiral's duration means that {\it LISA}'s position and orientation
330: changes considerably during the measurement; this, plus the impact of
331: precession, is largely responsible for pinning down the sky position
332: of coalescing sources.
333: 
334: The inspiral GW signal is modeled using the post-Newtonian
335: (PN) formalism {\citep{b06,bdiww95,ww96}}.  More
336: specifically, we use the ``restricted'' PN approximation.
337: The waveform can be written schematically as \citep{cf94}
338: \begin{equation}
339: h(t) = \mathrm{Re}\left[\sum_{x,m}
340: h^x_m(t)e^{im\Phi_{\mathrm{orb}}(t)}\right] \, ,
341: \label{eq:restrictedPN}
342: \end{equation}
343: where $x$ labels PN order, $m$ is a harmonic index, and
344: $\Phi_{\mathrm{orb}}(t)$ is the orbital phase.  In the restricted
345: PN case, we discard all amplitudes except the strongest
346: Newtonian quadrupole, $h_2^0$.  The phase, however, is computed to a
347: desired PN order (here the second one, or 2PN).  It has been recognized for
348: some time that important additional information is carried by
349: subleading harmonics (e.g., Hellings \& Moore 2003).  Recent work
350: {\citep{arunetal,ts08}} has confirmed that sky position and distance
351: are more sharply constrained when these terms are included in the wave
352: model (although there still appears to be some disagreement among
353: different groups as to the extent of the improvement).  We plan to
354: include these terms in future work, examining how higher harmonics and
355: precession combine to localize MBHB systems.
356: 
357: Spin-induced precession arises from geodetic and gravitomagnetic
358: influences on the angular momentum vectors present in the binary.
359: Averaged over an orbit, the leading-order spin precession equations
360: are {\citep{acst94,k95}}
361: \begin{eqnarray}
362: \mathbf{\dot{S}}_1 &=&
363: \frac{1}{r^3}\left[\left(2+\frac{3}{2}\frac{m_2}{m_1}\right)\mu
364: \sqrt{Mr} \mathbf{\hat{L}}\right] \times \mathbf{S}_1 +
365: \frac{1}{r^3}\left[\frac{1}{2}\mathbf{S}_2-\frac{3}{2}(\mathbf{S}_2\cdot
366: \mathbf{\hat{L}})\mathbf{\hat{L}}\right] \times \mathbf{S}_1 \, ,
367: \label{eq:S1dot}\\
368: \mathbf{\dot{S}}_2 &=&
369: \frac{1}{r^3}\left[\left(2+\frac{3}{2}\frac{m_1}{m_2}\right)\mu
370: \sqrt{Mr} \mathbf{\hat{L}}\right] \times \mathbf{S}_2 +
371: \frac{1}{r^3}\left[\frac{1}{2}\mathbf{S}_1-\frac{3}{2}(\mathbf{S}_1\cdot
372: \mathbf{\hat{L}})\mathbf{\hat{L}}\right] \times \mathbf{S}_2 \, ,
373: \label{eq:S2dot}
374: \end{eqnarray}
375: where $m_1$ and $m_2$ are the masses of holes 1 and 2, $M = m_1 + m_2$ 
376: is the total mass, $\mu = m_1m_2/M$ is the reduced mass,  
377: $\mathbf{\hat{L}}$ is the normal to the orbital plane, 
378: $\mathbf{S}_1$ and $\mathbf{S}_2$ are the spins of the holes,
379: and $r$ is their orbital separation in harmonic coordinates. 
380: Since total angular momentum is conserved
381: on short timescales (ignoring the inspiral), the orbital angular
382: momentum precesses to compensate.  These precessions drive several
383: modifications to the orbital phase.  One correction is due to
384: spin-orbit and spin-spin terms in the post-Newtonian expression for
385: orbital frequency: Precession makes these terms time dependent, adding
386: a new modulation to the GW phase.  Another correction arises from the
387: changing orientation of the orbital plane \citep{acst94}.
388: 
389: The waveform is described as two polarizations propagating in the
390: $-\mathbf{\hat{n}}$ direction, where $\mathbf{\hat{n}}$ gives the
391: direction to the binary on the sky:
392: \begin{eqnarray}
393: h_+(t) &=& \frac{2{\mathcal M}^{5/3}(\pi
394: f)^{2/3}}{D_L}[1+(\mathbf{\hat{L}}\cdot \mathbf{\hat{n}})^2]
395: \cos[\Phi(t) + \delta_p \Phi(t)] \, ,
396: \label{eq:hplus}\\
397: h_{\times}(t) &=& -\frac{4{\mathcal M}^{5/3}(\pi
398: f)^{2/3}}{D_L}(\mathbf{\hat{L}}\cdot \mathbf{\hat{n}}) \sin[\Phi(t) +
399: \delta_p \Phi(t)] \, ,
400: \label{eq:hcross}
401: \end{eqnarray}
402: where $f$ is the frequency of the waves, $D_L$ is the luminosity
403: distance,
404: $\Phi(t) = \int^t 2\pi f(t') dt'$ is the wave phase (twice the orbital
405: phase in the quadrupole approximation), and $\delta_p \Phi(t)$ is the
406: precessional correction to the orbital phase due to the changing
407: orientation of the orbital plane.  The ``chirp mass'' $\mathcal{M}$,
408: which controls the frequency evolution, or chirp, of the signal, is
409: given by $\mathcal{M} = \mu^{3/5}M^{2/5}$.  Note that the
410: relative weight of the polarizations is determined by the binary's
411: inclination to the line of sight $\iota$, where $\cos \iota =
412: \mathbf{\hat{L}}\cdot \mathbf{\hat{n}}$.
413: 
414: \subsection{MBHB gravitational waves as measured by {\it LISA}}
415: \label{sec:extrinsicGW}
416: 
417: The baseline {\it LISA} mission consists of three spacecraft arranged
418: in an equilateral triangle with arms of length $L = 5 \times 10^6$ km.
419: The center of this constellation orbits the Sun $20^\circ$ behind the
420: earth, with the triangle oriented at $60^\circ$ to the ecliptic.  The
421: individual spacecraft thus orbit in different planes, causing the
422: triangle to ``roll'' about itself in its orbit.  To model {\it LISA}'s
423: response to a signal, we use the long wavelength ($\lambda \gg L$)
424: approximation introduced by \citet{c98}.  The three arms of the
425: triangle act as a pair of two-arm detectors; the signals from the
426: three arms are combined in such a way that the noise spectrum in each
427: ``synthesized'' detector is independent of the other.  Each detector
428: measures both GW polarizations, weighted by antenna response functions
429: $F_i^+(t)$ and $F_i^\times (t)$, where $i \in \{{\rm I, II}\}$ labels the
430: synthesized detector.  These functions depend on the source's sky
431: position as measured in the {\it detector} frame.  Because of the
432: constellation's rolling orbital motion, these functions are time
433: dependent.  The motion around the Sun also causes a Doppler shift in
434: the signal's frequency.  This time dependence allows the sky position
435: of the source to be measured.
436: 
437: An additional time dependence is produced by the precession of the
438: binary's orbital plane.  We saw above that the polarizations strongly
439: depend on the binary's inclination to the line of sight $\iota$ (via
440: $\cos\iota = \mathbf{\hat{L}}\cdot\mathbf{\hat{n}}$).  The antenna
441: pattern functions $F_i^{+,\times}(t)$ also depend on the component of
442: $\mathbf{\hat{L}}$ that is perpendicular to the line of sight.  Since
443: precession changes the direction of the orbital plane, the measured
444: polarizations would vary in time even if the detector were static.
445: The timescale of these precessions is much shorter than the {\it
446: LISA} orbital timescale of 1 yr, so they provide a great deal of
447: additional information about sky position.
448: 
449: We perform the parameter estimation analysis in the frequency domain, so
450: we Fourier transform the signal using the stationary phase approximation.  The final measured waveform is given by
451: \begin{equation}
452: \tilde{h}_i(f) = \sqrt{\frac{5}{96}}\frac{\pi^{-2/3} {\mathcal
453: M}^{5/6}}{D_L}A_{\mathrm{pol}, i}[t(f)] f^{-7/6}
454: e^{i(\Psi(f) - \varphi_{\mathrm{pol},i}[t(f)] -
455: \varphi_D[t(f)] - \delta_p\Phi[t(f)])} \, ,
456: \label{eq:freqdomainsignal}
457: \end{equation}
458: where $A_{\mathrm{pol}, i}(t)$ is the ``polarization amplitude,''
459: $\varphi_{\mathrm{pol},i}(t)$ is the ``polarization phase,'' and
460: $\varphi_D(t)$ is the Doppler phase.  The first two encode modulations
461: due to the relative polarization weightings and antenna pattern functions; the
462: latter is due to the Doppler shift in the GW frequency.  Full
463: expressions for these functions can be found in Paper I.  To second
464: post-Newtonian order, the phase $\Psi(f)$ is given by {\citep{pw95}}
465: \begin{eqnarray}
466: \Psi(f) &=& 2\pi f t_c - \Phi_c - \frac{\pi}{4} + \frac{3}{128}(\pi
467: {\mathcal M} f)^{-5/3} \left[1 +
468: \frac{20}{9}\left(\frac{743}{336} + \frac{11}{4}\eta \right)(\pi
469: Mf)^{2/3}\right.
470: \nonumber\\
471: & &\left. - 4(4\pi - \beta)(\pi Mf) + 10\left(\frac{3058673}{1016064}
472: + \frac{5429}{1008}\eta + \frac{617}{144}\eta^2 - \sigma \right)(\pi
473: Mf)^{4/3}\right] \, ,
474: \label{eq:PNpsi}
475: \end{eqnarray}
476: where $\eta = \mu/M$, $t_c$ is the ``merger time'' (the time at which
477: $f$ formally diverges in the post-Newtonian framework), $\Phi_c$ is
478: the phase at $t_c$, and $\beta$ and $\sigma$ are ``spin-orbit'' and
479: ``spin-spin'' parameters, respectively, encoding the manner in which
480: interactions between the spins and the orbits impact the phase.  See
481: Paper I for precise definitions.  For our purposes, the most important
482: property of $\beta$ and $\sigma$ is that they oscillate due to
483: precession, modulating the phase $\Psi(f)$.  Note that waveform
484: (\ref{eq:freqdomainsignal}) depends on 15 parameters: 2 masses, 6
485: initial spin components, 1 distance, 2 sky angles, 2 initial orbital
486: plane angles, $t_c$, and $\Phi_c$.
487: 
488: The end of inspiral/beginning of merger is a somewhat ad hoc and
489: fuzzy boundary.  Indeed, recent numerical computations have shown that
490: the GWs produced by a binary that coalesces into a single body do not show
491: any particular special feature as the black holes come together,
492: instead smoothly chirping through this transition
493: \citep{bcp07,panetal08}.  Since we are not including numerical merger
494: waves in our analysis, we require some point to terminate our
495: post-Newtonian expansion.  Most studies show that the inspiral comes
496: to an end when the separation of the bodies in harmonic coordinates is
497: roughly $r \sim 6M$; at this point, the system's GW frequency is
498: approximately given by
499: \begin{equation}
500: f_{\mathrm{merge}} \simeq \frac{2}{2\pi}\Omega_{\mathrm{Kepler}}(r =
501: 6M) \simeq \frac{0.02}{M} \, ,
502: \label{eq:fmerge}
503: \end{equation}
504: where $\Omega_{\mathrm{Kepler}} = (M/r^3)^{1/2}$ is the Keplerian orbital 
505: angular frequency.
506: (The factor $1/2\pi$ converts from angular frequency to frequency; the
507: additional factor of $2$ accounts for the quadrupolar nature of
508: gravitational waves.)  We use equation (\ref{eq:fmerge}) throughout our
509: analysis to terminate the inspiral.
510: 
511: \subsection{Summary of parameter measurement formalism}
512: \label{sec:formalism}
513: 
514: To estimate how well {\it LISA} will be able to measure MBHB
515: parameters, we use the maximum likelihood formalism developed by \citet{f92}.  The GW signal $h_i(t)$ described above is corrupted by noise
516: to produce a measured signal $s_i(t) = h_i(t) + n_i(t)$.  We assume
517: the noise $n_i(t)$ is zero mean, wide-sense stationary, Gaussian, and
518: uncorrelated between the two detectors ($i \in \{{\rm I, II}\}$).  It can
519: be characterized by its (one sided) power spectral density (PSD) $S_n(f)$, which is described in detail in \S\ IIIB of Paper I.  The PSD
520: includes instrumental noise intrinsic to the detector, based on the
521: model described by \citet{lhh00} and implemented on the
522: World Wide Web\footnote{See http://www.srl.caltech.edu/\~{}shane/sensitivity/.}, with
523: important numerical factors described in \S\ IIC of \citet{bbw05}.
524: It also includes astrophysical noise arising from a confused
525: background of Galactic \citep{nyz01} and extragalactic \citep{fp03}
526: white dwarf binaries.  With the noise PSD, we can define the inner product between two signals $a$ and $b$ as 
527: \begin{equation}
528: (a|b) = 4 \, \mathrm{Re} \left[\int_0^{\infty} df
529: \frac{\tilde{a}^*(f)\tilde{b}(f)}{S_n(f)}\right]
530: = 2 \int_0^{\infty} df \frac{\tilde{a}^*(f)\tilde{b}(f) +
531: \tilde{a}(f)\tilde{b}^*(f)}{S_n(f)}  \, .
532: \label{eq:innerproduct}
533: \end{equation}
534: 
535: We write a GW as $h(\boldsymbol{\theta})$, where the components of the vector $\boldsymbol{\theta}$ represent the 15 parameters describing MBHB waves.  We assume that a detection has occurred --- that a signal with particular parameters $\boldsymbol{\tilde{\theta}}$ is present in the data --- and that maximum likelihood (ML) estimates $\boldsymbol{\hat{\theta}}$ \citep{cf94} of these parameters have been found.  The signal-to-noise ratio (S/N) is then given by
536: \begin{equation}
537: \mathrm{S/N} \approx (h(\boldsymbol{\hat{\theta}}) |
538: h(\boldsymbol{\hat{\theta}}))^{1/2} \, .
539: \label{eq:finalSNR}
540: \end{equation}
541: 
542: Our ultimate goal is to understand the errors in the ML parameters.  To do so, we examine the probability that the GW parameters are given by $\boldsymbol{\tilde{\theta}}$, expanded around the ML estimate $\boldsymbol{\hat{\theta}}$ \citep{cf94, pw95}:
543: \begin{equation}
544: p(\boldsymbol{\tilde{\theta}} | s) \propto \exp \left(-\frac{1}{2}\Gamma_{ab}\delta
545: \theta^a \delta \theta^b \right) \, ,
546: \label{eq:linearizedprob}
547: \end{equation}
548: where $\delta \theta^a = \tilde{\theta}^a - \hat{\theta}^a$ and
549: \begin{equation}
550: \Gamma_{ab} = \left(\frac{\partial h}{\partial \theta^a}\left|
551: \frac{\partial h}{\partial \theta^b}\right.\right) \, ,
552: \label{eq:fishermatrix}
553: \end{equation}
554: evaluated at $\boldsymbol{\theta} = \boldsymbol{\hat{\theta}}$, is
555: known as the Fisher matrix.  This expression holds for
556: large values of the S/N and is known as the Gaussian
557: approximation.  For the two-detector case, we can exploit the
558: independence of the noises to write the total Fisher matrix $\Gamma_{ab}^{\rm tot}$ as the sum of the individual Fisher matrices.  If we define the covariance
559: matrix $\Sigma^{ab} = (\Gamma_{\rm tot}^{-1})^{ab}$, we can then estimate the measurement error in some parameter
560: $\theta^a$,
561: \begin{equation}
562: \Delta \theta^a \equiv \sqrt{\langle(\delta \theta^a)^2\rangle} =
563: \sqrt{\Sigma^{aa}} \, ,
564: \label{eq:error}
565: \end{equation}
566: as well as the correlation between two parameters,
567: \begin{equation}
568: c^{ab} \equiv \frac{\langle\delta \theta^a \delta
569: \theta^b\rangle}{\Delta \theta^a \Delta \theta^b} =
570: \frac{\Sigma^{ab}}{\sqrt{\Sigma^{aa}\Sigma^{bb}}} \, .
571: \label{eq:correlation}
572: \end{equation}
573: 
574: We are particularly interested in the sky position, which we parameterize by the coordinates
575: ($\bar{\mu}_N = \cos \bar{\theta}_N$, $\bar{\phi}_N$).\footnote{The bar
576: over the angles conforms to the notation in Paper I for quantities
577: measured in the solar system barycenter frame, as opposed to those
578: measured in the time-varying detector frame.}  We want to convert
579: from errors in these two parameters to an error ellipse on the sky.
580: To do so, we first perform a change of coordinates from $\bar{\mu}_N$
581: to $\bar{\theta}_N$.  For small deviations from the ML estimate, we
582: have $\delta\bar{\theta}_N = (d\bar{\theta}_N/d\bar{\mu}_N)\delta
583: \bar{\mu}_N = -\delta \bar{\mu}_N/\sin \bar{\theta}_N$.  Next, we
584: recognize that due to the geometric properties of the sphere, the
585: same $\delta \bar{\phi}_N$ corresponds to a different ``proper'' angle
586: depending on the value of $\bar{\theta}_N$: $\delta \bar{\phi}_N^p =
587: \sin \bar{\theta}_N \delta \bar{\phi}_N$.  With these modifications,
588: equation (\ref{eq:linearizedprob}) becomes
589: \begin{equation}
590: p(\boldsymbol{\tilde{\theta}}|s) \propto \exp
591: \left(-\frac{1}{2}\Gamma^p_{a^\prime b^\prime}\delta
592: \theta^{a^\prime}_p \delta \theta^{b^\prime}_p \right) \, .
593: \label{eq:properundiagonalized}
594: \end{equation}
595: Here $\delta \theta^{a^\prime}_p \equiv
596: (\delta\bar\theta_N,\delta\bar\phi_N^p)$ denotes the proper errors
597: accounting for the metric of the sphere, and $\Gamma^p_{a^\prime
598: b^\prime}$ represents the equivalent Fisher matrix with all conversion
599: factors absorbed inside:
600: \begin{align}
601: \Gamma^p_{\bar{\theta}_N \bar{\theta}_N} &= \sin^2 \bar{\theta}_N
602: \Gamma_{\bar{\mu}_N \bar{\mu}_N} \, , \\
603: \Gamma^p_{\bar{\phi}_N \bar{\phi}_N} &= \csc^2 \bar{\theta}_N
604: \Gamma_{\bar{\phi}_N \bar{\phi}_N} \, , \\
605: \Gamma^p_{\bar{\theta}_N \bar{\phi}_N} &= \Gamma^p_{\bar{\phi}_N
606: \bar{\theta}_N} = (-\sin \bar{\theta}_N)(\csc \bar{\theta}_N)\Gamma_{\bar{\mu}_N \bar{\phi}_N} = -\Gamma_{\bar{\mu}_N \bar{\phi}_N} \, , 
607: \label{eq:properfisher}
608: \end{align}
609: and so on for the rest of the elements.
610: The inverse of this matrix is the proper covariance matrix,
611: $\Sigma_p^{a^\prime b^\prime}$.  Consider now just the $2 \times 2$ subspace
612: of the covariance matrix containing the sky position variables.
613: Let the eigenvalues of this subspace be $\lambda_\pm$.
614: If we define the error ellipse such that the probability that the
615: source lies outside of it is $e^{-1}$ (corresponding to a $\approx
616: 63\%$ confidence interval), then the major and minor axes are given by
617: $2a = 2(2\lambda_+)^{1/2}$ and $2b = 2(2\lambda_-)^{1/2}$, respectively.  
618: Expressed in terms of the original covariance matrix, these are
619: \begin{equation}
620: 2\left[ \vphantom{\sqrt{(\Sigma^{\bar{\mu}_N\bar{\mu}_N} -
621: \Sigma^{\bar{\phi}_N\bar{\phi}_N})^2 +
622: 4(\Sigma^{\bar{\mu}_N\bar{\phi}_N})^2}}
623: \csc^2 \bar{\theta}_N \Sigma^{\bar{\mu}_N\bar{\mu}_N} + 
624: \sin^2 \bar{\theta}_N \Sigma^{\bar{\phi}_N\bar{\phi}_N}
625: \pm \sqrt{(\csc^2 \bar{\theta}_N \Sigma^{\bar{\mu}_N\bar{\mu}_N} -
626: \sin^2 \bar{\theta}_N \Sigma^{\bar{\phi}_N\bar{\phi}_N})^2 +
627: 4(\Sigma^{\bar{\mu}_N\bar{\phi}_N})^2} \right]^{1/2} \, .
628: \label{eq:2a2b}
629: \end{equation}
630: Many previous analyses have reported the ellipse's area $\Delta
631: \Omega_N = \pi a b$ or $(\Delta\Omega_N)^{1/2}$, the side of a square of
632: equivalent area, as the sky position error {\citep{c98,v04,bbw05,hh05}}.
633: Information about the ellipse's shape, crucial input to coordinating
634: GW observations with telescopes, is not included in such a measure.
635: By examining both $2a$ and $2b$, this information is restored.
636: 
637: In an empty universe, the third dimension of our pixel would be
638: determined from the luminosity distance error, $\Delta D_L/D_L =
639: (\Sigma^{\ln D_L, \ln D_L})^{1/2}$.  As already discussed, this
640: quantity is often so small that weak gravitational lensing is
641: expected to dominate the distance error budget.  We thus expect
642: \begin{equation}
643: \frac{\Delta D_L}{D_L} \simeq \sqrt{\Sigma^{\ln D_L, \ln D_L} +
644: \sigma^2_{\rm lens}(z)} \simeq {\rm max}\left[\sqrt{\Sigma^{\ln D_L,
645: \ln D_L}}, \sigma_{\rm lens}(z)\right] \, ,
646: \end{equation}
647: where the lensing dispersion $\sigma_{\rm lens}(z)$ is described in,
648: for example, \citet{whm02}.  If cosmological parameter errors can be
649: neglected, then one typically finds that the redshift error is about
650: equal to the distance error {\citep{h02}}: $\Delta z/z \approx \Delta
651: D_L/D_L$, independent of redshift.
652: 
653: \subsection{Final position and distance accuracy}
654: \label{sec:review}
655: 
656: We conclude this section by reviewing the final accuracy with which
657: MBHB sky position and distance can be determined with GWs.  In Paper
658: I, we described the code which implements the measurement formalism
659: described in \S \S\ {\ref{sec:intrinsicGW}} -- {\ref{sec:formalism}}
660: above.  This formalism is coupled to a Monte Carlo engine: We specify
661: the rest frame masses and redshift and then randomly choose the sky
662: position, initial angular momentum and spin directions, and spin
663: magnitudes for each binary.  We also uniformly distribute the time
664: parameter $t_c$ of each binary over the assumed duration of the {\it
665: LISA} mission (which we take to be 3 yr).  Some binaries are
666: therefore already ``on'' when the mission begins, and as such may
667: radiate in band for less time than other binaries which enter the band
668: later.  The random distribution of merger times also means that the
669: relative azimuth between a binary's sky position and {\it LISA}'s
670: orbital position at merger, $\delta\phi = \bar{\phi}_N - \phi_{\it
671: LISA}(t_c)$, is itself randomly distributed, even if we choose to
672: focus on a particular $\bar{\phi}_N$.  We discuss the implications of
673: this in more detail in \S\ {\ref{sec:angdependence}}.  Computing the
674: errors for a large number of binaries ($10^4$ for each mass and
675: redshift set, unless otherwise stated) allows us to make statements
676: about the distribution of parameter errors for different mass and
677: redshift combinations.
678: 
679: Table {\ref{tab:finalvals}} summarizes the final accuracy of MBHB sky
680: position and distance measurements.  Here we show the {\it median}
681: measurement accuracy from many Monte Carlo surveys.  Several trends
682: are particularly noteworthy.  At all redshifts, the accuracy is worst
683: for the largest masses.  This is because the most massive systems are
684: in band for the least amount of time: The frequency at which inspiral
685: ends is inversely proportional to mass (see eq. [\ref{eq:fmerge}]),
686: and more massive systems move more rapidly from low frequency to high
687: frequency.  The short time these systems spend in band means that {\it
688: LISA} measures a relatively small number of modulations (whether
689: induced by the constellation's orbit or by spin precession).  Second,
690: note that the results for $m_1 = m_2$ tend to be less accurate than
691: results with similar total mass but for which $m_1 > m_2$.  The cause
692: of this phenomenon lies in precession equations (\ref{eq:S1dot})
693: and (\ref{eq:S2dot}): When $m_1 \ne m_2$, the two spins precess at
694: different rates, imposing richer modulations on the measured GWs.
695: Since $m_1 = m_2$ is a rather artificial limit, we expect that the
696: more accurate results for nonunity mass ratio will be the rule.
697: 
698: 
699: \begin{table}[p]
700: \begin{center}
701: \caption{Sky position and distance accuracy at merger}
702: \begin{tabular}{ccccccc}
703: \\
704: \tableline \tableline
705: $z$ &
706: $m_1\ (M_\odot)$ &
707: $m_2\ (M_\odot)$ &
708: $2a\ (\mathrm{arcmin})$ &
709: $2b\ (\mathrm{arcmin})$ &
710: $\Delta\Omega_N (\mathrm{deg}^2)$ &
711: $\Delta D_L/D_L$ \\
712: 
713: \tableline
714: 
715: 1 & $10^5$ & $10^5$ & 27.3 & 13.3 & 0.0729 & 0.00398 \\ 
716: & $3\times 10^5$ & $10^5$ & 16.9 & 7.33 & 0.0233 & 0.00240 \\ 
717: & & $3\times 10^5$ & 23.3 & 11.8 & 0.0556 & 0.00357 \\ 
718: & $10^6$ & $10^5$ & 27.2 & 6.62 & 0.0235 & 0.00320 \\ 
719: & & $3\times 10^5$ & 31.3 & 13.2 & 0.0705 & 0.00393 \\ 
720: & & $10^6$ & 40.2 & 21.9 & 0.176 & 0.00560 \\
721: & $3\times 10^6$ & $3\times 10^5$ & 34.1 & 9.20 & 0.0445 & 0.00376 \\ 
722: & & $10^6$ & 32.3 & 14.7 & 0.0839 & 0.00419 \\ 
723: & & $3\times 10^6$ & 43.3 & 22.3 & 0.193 & 0.00689 \\ 
724: & $10^7$ & $10^6$ & 37.6 & 12.2 & 0.0670 & 0.00457 \\ 
725: & & $3\times 10^6$ & 42.1 & 19.0 & 0.142 & 0.00610 \\ 
726: & & $10^7$ & 81.3 & 38.6 & 0.680 & 0.0250 \\ 
727: 
728: \tableline
729: 
730: 3 & $10^5$ & $10^5$ & 81.0 & 40.8 & 0.665 & 0.0123 \\ 
731: & $3\times 10^5$ & $10^5$ & 92.5 & 39.5 & 0.656 & 0.0126 \\ 
732: & & $3\times 10^5$ & 142 & 75.7 & 2.15 & 0.0201 \\ 
733: & $10^6$ & $10^5$ & 141 & 36.6 & 0.739 & 0.0155 \\ 
734: & & $3\times 10^5$ & 129 & 56.7 & 1.25 & 0.0161 \\ 
735: & & $10^6$ & 158 & 84.3 & 2.64 & 0.0237 \\ 
736: & $3\times 10^6$ & $3\times 10^5$ & 132 & 40.3 & 0.751 & 0.0153 \\ 
737: & & $10^6$ & 142 & 64.6 & 1.65 & 0.0193 \\ 
738: & & $3\times 10^6$ & 224 & 111 & 5.08 & 0.0422 \\ 
739: & $10^7$ & $10^6$ & 206 & 78.5 & 2.74 & 0.0293 \\ 
740: & & $3\times 10^6$ & 297 & 152 & 9.40 & 0.0805 \\ 
741: & & $10^7$ & 2000 & 583 & 256 & 2.41 \\ 
742: 
743: \tableline
744: 
745: 5 & $10^5$ & $10^5$ & 169 & 85.7 & 2.93 & 0.0260 \\ 
746: & $3\times 10^5$ & $10^5$ & 217 & 95.8 & 3.73 & 0.0284 \\ 
747: & & $3\times 10^5$ & 295 & 161 & 9.29 & 0.0409 \\ 
748: & $10^6$ & $10^5$ & 248 & 66.8 & 2.35 & 0.0273 \\ 
749: & & $3\times 10^5$ & 233 & 101 & 3.96 & 0.0294 \\ 
750: & & $10^6$ & 315 & 162 & 10.2 & 0.0501 \\ 
751: & $3\times 10^6$ & $3\times 10^5$ & 265 & 86.4 & 3.27 & 0.0318 \\ 
752: & & $10^6$ & 304 & 139 & 7.52 & 0.0436 \\ 
753: & & $3\times 10^6$ & 538 & 260 & 29.5 & 0.140 \\ 
754: & $10^7$ & $10^6$ & 577 & 290 & 31.9 & 0.124 \\ 
755: & & $3\times 10^6$ & 1720 & 621 & 234 & 1.24 \\ 
756: & & $10^7$ & 180000 & 29600 & $1.15\times10^6$ & 377 \\
757: 
758: \tableline
759: \end{tabular}
760: 
761: \tablecomments{Median errors in sky position and distance for binaries
762: at various masses and redshifts $z = 1$, $3$, and $5$; $2a$ is the
763: major axis of the sky position error ellipse, $2b$ is the minor axis, and $\Delta \Omega_N$ is the ellipse's area.  (Note that since all of these quantities are medians, it is {\it not} true that $\Delta \Omega_N = \pi a b$.)
764: At each mass and redshift point, $10^4$ binaries are drawn with
765: randomly distributed sky positions, initial orbit and spin orientations, spin magnitudes, and $t_c$.}
766: \label{tab:finalvals}
767: \end{center}
768: \end{table}
769: 
770: Independent of these trends, an important result is that 
771: MBHB systems are pinned down on
772: the sky fairly accurately at $z = 1$.  Modulo the higher mass
773: binaries, the median major axis of the sky position error ellipse is
774: typically about $15^\prime - 45^\prime$, and the median minor axis is
775: about $5^\prime - 20^\prime$, with a total ellipse area considerably
776: smaller than $1\ \mathrm{deg}^2$ (ranging from about 0.02 to 0.2 
777: $\mathrm{deg}^2$).  Sources at $z = 1$ are located accurately enough that one
778: can comfortably contemplate searching the GW error ellipse for MBHB
779: counterparts with future survey instruments, such as 
780: the Large Synoptic Survey Telescope (LSST; Tyson et al. 2002).
781: 
782: At higher redshift, positional accuracy degrades.  This is due to the
783: weakening of the signal with distance and to the redshifting of the
784: waves' spectrum, so that the signal tends to spend less time in band.
785: At $z = 3$, the major axis of the error ellipse is $\sim 1^\circ-4^\circ$
786: across, and the minor axis is $\sim 40^\prime-110^\prime$.  The total
787: area of this ellipse is $\sim 0.5 - 5\ \mathrm{deg}^2$.  At $z = 5$,
788: this degrades further to $\sim 3^\circ - 5^\circ$ for the major axis,
789: $\sim 1^\circ - 3^\circ$ for the minor axis, and $\sim 2 - 10\ \mathrm{deg}^2$ 
790: for the total area.  These degraded accuracies are still
791: sufficiently well determined that telescopic searches for MBHB
792: counterparts have a good chance to be fruitful (although not nearly as
793: simple as they would be at $z \sim 1$).
794: 
795: In all cases, the GW distance determination is extremely accurate: For
796: all but the highest masses, $\delta D_L/D_L \lesssim 0.7\%$ at $z =
797: 1$, $\lesssim 4\%$ at $z = 3$, and $\lesssim 5\%$ at $z = 5$.
798: Distance is determined so precisely that these errors are in fact
799: irrelevant --- weak gravitational lensing will dominate the distance
800: error budget for all but the most massive MBHB events.
801: 
802: Although the median values reported in Table {\ref{tab:finalvals}}
803: indicate the typical sky position accuracies we expect, it
804: should be emphasized that they are taken from broad distributions.
805: Figure {\ref{fig:final_axes}} presents the distributions we computed
806: for binaries at $z = 1$ with masses $m_1 = 10^6\,M_\odot$ and $m_2 =
807: (10^5, 3 \times 10^5, 10^6) M_\odot$.  Note that the major axis
808: distribution ({\it left}) is rather flat when compared to the minor
809: axis distribution ({\it right}).  It lacks a single well-defined peak;
810: in fact, it is actually bimodal for $m_1/m_2 > 1$, with one peak near
811: $10^\prime - 20^\prime$ and another closer to $1^\circ - 2^\circ$.  We find
812: that this behavior holds over all mass and redshift cases of interest,
813: with only slight variations.  At smaller masses, the distribution is
814: broader than the cases pictured, without strong bimodality; for larger
815: masses, the distribution is somewhat narrower and tends to develop two
816: very distinct peaks.  Higher mass ratios tend to accentuate the peaks.
817: These results hold for higher redshift, except that transitions
818: between various behaviors occur at smaller total mass [in keeping with
819: the fact that it is not mass but $(1 + z)$ times the mass that
820: determines dynamical behavior].
821: 
822: The minor axis distribution exhibits a rather long tail to very small
823: values, especially when the mass ratio is large.  For $m_1/m_2 = 10$,
824: the distribution peaks near a minor axis $\sim 10^\prime$, but
825: extends from roughly $1^{\prime \prime}$ to about $100^\prime$.  As
826: the mass ratio approaches 1, the peak moves to slightly larger values
827: (slightly more than $10^\prime$ for $m_1/m_2 = 3$; roughly $30^\prime$ 
828: for $m_1/m_2 = 1$), and the tail becomes less populated
829: (although the distributions span roughly the same extent as when
830: $m_1/m_2 = 10$).  We find that this tail exists for all interesting
831: mass and redshift combinations, with the same strong dependence on
832: mass ratio as shown in Figure {\ref{fig:final_axes}}.
833: 
834: \begin{figure}[t]
835: \begin{center}
836: \includegraphics[scale=0.52]{final_2a.pdf}
837: \includegraphics[scale=0.52]{final_2b.pdf}
838: \caption{Distribution of the major axis $2a$ ({\it left}) and minor axis $2b$
839: ({\it right}) of the sky position error ellipse for $10^4$ binaries with
840: $m_1 = 10^6 M_\odot$ and $m_2 = 10^5 M_\odot$ ({\it solid line}), $3\times
841: 10^5 M_\odot$ ({\it dashed line}), and $10^6 M_\odot$ 
842: ({\it dash-dotted line}) at
843: $z = 1$.  Note the bimodal character of the major axis as well as
844: the long tail down to small minor axis, both of which are particularly prominent for
845: larger mass ratios.}
846: \label{fig:final_axes}
847: \end{center}
848: \end{figure}
849: 
850: Finally, it is worth emphasizing that spin-induced precession has a
851: significant effect on the accuracy which we report here.  In Paper I,
852: we compared these results to those obtained with a code which does not
853: include precession physics.  Position and distance accuracy are
854: determined only by the detector's motion in this case.  We found that
855: precession effects reduce the major axis of the sky position error
856: ellipse by a factor of $\sim 2 - 7$ and the minor axis by a factor of $\sim
857: 2 - 10$.  The distance error is likewise improved by a factor of $\sim 2
858: - 7$.  Factors of a few or more improvement are also seen when
859: additional wave harmonics are included {\citep{arunetal,ts08}}.  We
860: are eager to see whether the improvements from precession and from
861: higher harmonics can be combined.
862: 
863: \section{Time and location dependence of the GW pixel}
864: \label{sec:results}
865: 
866: We turn now to a detailed discussion of how well GWs localize an MBHB
867: system as a function of time before final merger and as a function of
868: sky location.  We begin by discussing the analysis of K07, 
869: which presents a clever algorithm for estimating
870: extrinsic parameter errors as a function of time until merger (although
871: at present it does not include spin precession).  We demonstrate that
872: their analysis unfortunately underestimates final position errors by
873: roughly a factor of $\sim 2 - 4$ or more (in angle) due to neglect of
874: certain parameter correlations; the underestimate is much less severe
875: a week or more prior to merger.  We then present our own results for
876: the time dependence of the GW pixel, using K07 for comparison where
877: appropriate.  Finally, we conclude with a brief study of the pixel's
878: dependence on sky location.
879: 
880: \subsection{Summary of K07}
881: \label{sec:khmf}
882: 
883: K07 have devised a new method, the harmonic mode
884: decomposition (HMD), to solve for the extrinsic parameter errors as a
885: function of time to merger.  In the HMD, modulations caused by {\em
886: LISA}'s motion are decoupled from the much faster inspiral timescales
887: and are then expanded in a Fourier series.  The resulting expression
888: for the measured signal features a time-dependent piece with no
889: dependence on the extrinsic parameters and a time-independent piece
890: with all the parameter dependence.  As a result, when Monte Carlo
891: simulations are done across parameter space, the time-dependent
892: integrals do not need to be recomputed for each sample of the
893: distribution.  This makes it possible to quickly survey the estimated
894: parameter errors across a wide range of parameter space.
895: 
896: As already emphasized, the waveform model used by K07 does not (yet)
897: include the impact of spin precession.  As such, we intend to use
898: their results as a baseline against which the
899: impact of spin precession can be compared.  Before doing so, 
900: we first checked to make sure that their
901: results were in agreement with a variant of our code which does not
902: include spin precession {\citep{h02}}.  To our surprise, we found that
903: the final position accuracy predicted by K07 was typically a factor
904: $\sim 2$ (in angle) more accurate than our code predicted.
905: 
906: After detailed study of the HMD algorithm and comparison with our
907: (precession free) code, we believe we understand the primary source of
908: this disagreement.  K07 define a set of ``slow'' parameters
909: $\boldsymbol{\theta}_{\mathrm{slow}}$, which correspond (with some
910: remappings) to our extrinsic parameters: $\ln D_L$, $\cos
911: \bar{\theta}_{L}$, $\cos \bar{\theta}_N$, $\bar{\phi}_{L}$, and
912: $\bar{\phi}_N$.  Here $(\bar{\theta}_L, \bar{\phi}_L)$ defines the
913: orientation of the orbital angular momentum $\mathbf{\hat{L}}$, which
914: is constant when spin precession is ignored.  K07 also define a set
915: of ``fast'' parameters $\boldsymbol{\theta}_{\mathrm{fast}}$, which,
916: modulo the exclusion of spin, map to our intrinsic parameters: $t_c$,
917: $\Phi_c$, $\ln \mathcal{M}$, and $\ln \eta$.  In their formulation of
918: the HMD, K07 approximate the cross-correlation between intrinsic and
919: extrinsic parameters to be zero.  Although the correlations between
920: intrinsic and extrinsic parameters tend to be small, they are not
921: zero.  We find that they typically range in magnitude from about 0.1
922: to 0.4, sometimes reaching $\sim 0.8$.  Neglecting these correlations
923: altogether leads to a systematic underestimate in extrinsic parameter
924: errors.
925: 
926: An example of this is shown in Tables {\ref{tab:full_errs_and_corrs}}
927: and {\ref{tab:khmf_errs_and_corrs}}.  To produce the data shown in
928: Table {\ref{tab:full_errs_and_corrs}}, we compute the Fisher matrix
929: $\Gamma_{ab}^{\rm tot}$ and then invert for the covariance matrix,
930: $\Sigma^{ab} = (\Gamma_{\rm tot}^{-1})^{ab}$.  Table
931: {\ref{tab:full_errs_and_corrs}} then presents a slightly massaged
932: representation of this matrix: Diagonal elements are the 1 $\sigma$
933: errors $(\Sigma^{aa})^{1/2}$, and off-diagonal elements are the
934: correlation coefficients $c^{ab} =
935: \Sigma^{ab}(\Sigma^{aa}\Sigma^{bb})^{-1/2}$.  Take particular note of
936: the magnitude of the correlations between intrinsic and extrinsic
937: parameters (the upper right-hand portion of Table 
938: \ref{tab:full_errs_and_corrs}).  Many entries
939: have values $\sim 0.2 - 0.3$, and two have values $\sim 0.7 - 0.8$.
940: 
941: \begin{table}[p]
942: \begin{center}
943: \caption{Full errors and correlations}
944: \begin{tabular}{cccccc|cccccc}
945: \tableline \tableline
946: &
947: $\ln D_L$ &
948: $\cos\bar{\theta}_L$ &
949: $\cos\bar{\theta}_N$ &
950: $\bar{\phi}_L$ &
951: $\bar{\phi}_N$ &
952: $t_c$ &
953: $\Phi_c$ &
954: $\ln \mathcal{M}$ &
955: $\ln\eta$ &
956: $\beta$ &
957: $\sigma$ \\
958: 
959: \tableline
960: 
961: $\ln D_L$ & $0.233$ & $-0.984$ & $0.878$ & $0.509$ & $0.213$ & $0.0801$ & $0.246$ & $0.227$ & $-0.186$ & $0.205$ & $-0.106$ \\
962: $\cos\bar{\theta}_L$ & $\cdots$ & $0.467$ & $-0.861$ & $-0.350$ & $-0.071$ & $-0.040$ & $-0.098$ & $-0.178$ & $0.138$ & $-0.156$ & $0.0622$ \\
963: $\cos\bar{\theta}_N$ & $\cdots$ & $\cdots$ & $0.0006$ & $0.465$ & $0.203$ & $0.0709$ & $0.231$ & $0.201$ & $-0.166$ & $0.181$ & $-0.095$ \\
964: $\bar{\phi}_L$ & $\cdots$ & $\cdots$ & $\cdots$ & $0.687$ & $0.782$ & $0.232$ & $0.827$ & $0.337$ & $-0.317$ & $0.328$ & $-0.259$ \\
965: $\bar{\phi}_N$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $0.0017$ & $0.193$ & $0.691$ & $0.252$ & $-0.244$ & $0.250$ & $-0.210$ \\
966: \tableline
967: $t_c$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $63.1$ & $0.705$ & $0.923$ & $-0.955$ & $0.942$ & $-0.993$ \\
968: $\Phi_c$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $4.00$ & $0.742$ & $-0.747$ & $0.748$ & $-0.726$ \\
969: $\ln \mathcal{M}$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $0.0010$ & $-0.995$ & $0.998$ & $-0.956$ \\
970: $\ln\eta$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $0.303$ & $-0.999$ & $0.981$ \\
971: $\beta$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $1.11$ & $-0.971$ \\
972: $\sigma$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $0.722$ \\
973: \tableline
974: \end{tabular}
975: 
976: \tablecomments{Example of errors (diagonal elements) and correlations
977: (off-diagonal elements) for a binary with $m_1 = 3 \times 10^6\,M_\odot$ and $m_2
978: = 10^6\,M_\odot$ at $z = 1$.  The errors in $\bar{\phi}_L$, $\bar{\phi}_N$, and $\Phi_c$ are measured in
979: radians; the error in $t_c$ is measured in seconds.  This example was taken from the same
980: Monte Carlo distribution used to make Table
981: {\ref{tab:compare_full_vs_khmf}}; in this particular case, the
982: randomly distributed parameters have the values $\cos\bar{\theta}_L = -0.628$, $\cos\bar{\theta}_N = 0.850$, $\bar{\phi}_L =
983: 3.50$ rad, $\bar{\phi}_N = 0.514$ rad, $t_c = 6.90 \times 10^7$ s, $\beta = 1.48$, and $\sigma = 0.107$.  Entries containing
984: ellipses can be found by symmetry.}
985: \label{tab:full_errs_and_corrs}
986: \end{center}
987: \end{table}
988: 
989: \begin{table}[p]
990: \begin{center}
991: \caption{Errors and correlations neglecting correlations between
992: intrinsic and extrinsic parameters}
993: \begin{tabular}{cccccc|cccccc}
994: \tableline \tableline
995: &
996: $\ln D_L$ &
997: $\cos\bar{\theta}_L$ &
998: $\cos\bar{\theta}_N$ &
999: $\bar{\phi}_L$ &
1000: $\bar{\phi}_N$ &
1001: $t_c$ &
1002: $\Phi_c$ &
1003: $\ln \mathcal{M}$ &
1004: $\ln \eta$ &
1005: $\beta$ &
1006: $\sigma$ \\
1007: \tableline
1008: $\ln D_L$ & $0.142$ & $-0.999$ & $0.721$ & $0.999$ & $0.0602$ & 0 & 0 & 0 & 0 & 0 & 0 \\
1009: $\cos \bar{\theta}_L$ & $\cdots$ & $0.294$ & $-0.721$ & $-0.999$ & $-0.0611$ & 0 & 0 & 0 & 0 & 0 & 0 \\
1010: $\cos \bar{\theta}_N$ & $\cdots$ & $\cdots$ & $0.0004$ & $0.719$ & $0.0576$ & 0 & 0 & 0 & 0 & 0 & 0 \\
1011: $\bar{\phi}_L$ & $\cdots$ & $\cdots$ & $\cdots$ & $0.120$ & $0.0664$ & 0 & 0 & 0 & 0 & 0 & 0 \\
1012: $\bar{\phi}_N $ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $0.0010$ & 0 & 0 & 0 & 0 & 0 & 0 \\
1013: \tableline
1014: $t_c$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $61.3$ & $0.990$ & $0.929$ & $-0.959$ & $0.946$ & $-0.993$ \\
1015: $\Phi_c$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $2.10$ & $0.969$ & $-0.988$ & $0.981$ & $-0.999$ \\
1016: $\ln \mathcal{M}$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $0.0010$ & $-0.995$ & $0.998$ & $-0.960$ \\
1017: $\ln\eta$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $0.288$ & $-0.999$ & $0.983$ \\
1018: $\beta$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $1.05$ & $-0.974$ \\
1019: $\sigma$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ & $0.697$ \\
1020: \tableline
1021: \end{tabular}
1022: 
1023: \tablecomments{Example of errors (diagonal elements) and correlations
1024: (off-diagonal elements) for a binary with $m_1 = 3 \times 10^6\,M_\odot$ and $m_2
1025: = 10^6\,M_\odot$ at $z = 1$ if correlations between intrinsic and extrinsic parameters are neglected.  The errors in $\bar{\phi}_L$, $\bar{\phi}_N$, and $\Phi_c$ are measured in
1026: radians; the error in $t_c$ is measured in seconds.  This example was taken from the same
1027: Monte Carlo distribution used to make Table
1028: {\ref{tab:compare_full_vs_khmf}}; in this particular case, the
1029: randomly distributed parameters have the values $\cos\bar{\theta}_L = -0.628$, $\cos\bar{\theta}_N = 0.850$, $\bar{\phi}_L =
1030: 3.50$ rad, $\bar{\phi}_N = 0.514$ rad, $t_c = 6.90 \times 10^7$ s, $\beta = 1.48$, and $\sigma = 0.107$.  Entries containing
1031: ellipses can be found by symmetry.}
1032: \label{tab:khmf_errs_and_corrs}
1033: \end{center}
1034: \end{table}
1035: 
1036: To see what effect neglecting the intrinsic-extrinsic correlations
1037: has, we repeat this exercise, with a slight modification: We compute
1038: $\Gamma^{\rm tot}_{ab}$ as before, but we now set to zero entries
1039: corresponding to mixed intrinsic/extrinsic parameters.  For example,
1040: we set by hand $\Gamma_{\ln D_L,\ \ln \mathcal {M}} = 0$.  We then
1041: invert this matrix to obtain $\Sigma^{ab}$.  The result is shown in
1042: Table {\ref{tab:khmf_errs_and_corrs}}.  Note that mean parameter
1043: error (diagonal entries) is often significantly smaller than errors
1044: when these correlations are not ignored.  The impact of correlations
1045: between intrinsic and extrinsic parameters is clearly not negligible.
1046: 
1047: Table {\ref{tab:compare_full_vs_khmf}} gives further examples
1048: illustrating the impact of neglecting these correlations on our
1049: estimates of {\it LISA}'s localization accuracy.  We show 10 points
1050: drawn from a $10^4$ binary Monte Carlo run; all use the same masses
1051: and redshifts ($m_1 = 3 \times 10^6\,M_\odot$, $m_2 = 10^6\,M_\odot$, and
1052: $z = 1$) but have different (randomly distributed) sky positions,
1053: orientations, spins, and $t_c$.  For these parameters, we find that
1054: neglecting intrinsic-extrinsic correlations causes one to
1055: underestimate the major axis of the position ellipse by a (median)
1056: factor of $\sim 2$ and the minor axis by a factor of $\sim 3-4$; the area is
1057: underestimated by a factor of $\sim 6-7$.
1058: 
1059: \begin{table}[t]
1060: \begin{center}
1061: \caption{Example sky position error measures from Monte Carlo sample, comparing full Fisher matrix technique with the K07 approximation}
1062: \begin{tabular}{cccccccc}
1063: \\
1064: \tableline \tableline
1065: \multicolumn{2}{c}{$2a$ (arcmin)} & 
1066: & 
1067: \multicolumn{2}{c}{$2b$ (arcmin)} &
1068: & 
1069: \multicolumn{2}{c}{$\Delta \Omega_N\ (\mathrm{deg}^2)$} \\
1070: \cline{1-2} \cline {4-5} \cline{7-8} 
1071: Full &
1072: K07 &
1073: & 
1074: Full &
1075: K07 &
1076: & 
1077: Full &
1078: K07 \\
1079: 
1080: \tableline
1081: % seed used for all this is 4589.
1082: % Data actually run for 1e6, 3e6, but I switched m1, m2 to match our
1083: % conventions.  Makes a difference if we try to replicate the numbers!
1084: 
1085: $201$  & $63.5$ & & $59.0$ & $15.1$ & & $2.59$  & $0.210$ \\
1086: $165$  & $120$  & & $108$  & $88.6$ & & $3.90$  & $2.32$  \\
1087: $117$  & $61.6$ & & $81.0$ & $14.3$ & & $2.07$  & $0.193$ \\
1088: $197$  & $69.7$ & & $46.7$ & $13.4$ & & $2.01$  & $0.204$ \\
1089: $10.9$ & $7.23$ & & $8.21$ & $5.03$ & & $0.0196$ & $0.00793$ \\
1090: $197$  & $51.2$ & & $55.2$ & $13.3$ & & $2.37$  & $0.149$ \\
1091: $46.7$ & $26.8$ & & $36.5$ & $9.19$ & & $0.372$ & $0.0538$ \\
1092: $18.1$ & $12.7$ & & $15.0$ & $6.37$ & & $0.0595$ & $0.0177$ \\
1093: $155$  & $92.4$ & & $88.5$ & $16.2$ & & $2.98$  & $0.326$ \\
1094: $146$  & $143$  & & $139$  & $10.9$ & & $4.43$  & $0.342$ \\
1095: \tableline
1096: \end{tabular}
1097: 
1098: \tablecomments{Ten Monte Carlo points comparing sky position accuracy
1099: for binaries with $m_1 = 3 \times 10^6\,M_\odot$ and $m_2 = 10^6\,M_\odot$
1100: at $z = 1$.}
1101: \label{tab:compare_full_vs_khmf}
1102: \end{center}
1103: \end{table}
1104: 
1105: Our conclusion is that the HMD technique developed by K07 is overly
1106: optimistic by a factor of $\sim 2-4$ or more (in angle) regarding the
1107: final accuracy with which GWs can locate an MBHB event on the sky.  As
1108: a prelude to the time evolution study we present in \S\
1109: {\ref{sec:timeevolve}}, we also examined how this underestimate
1110: evolves as merger is approached.  To our relief, it appears that this
1111: underestimate is {\it much} smaller prior to merger: For the handful
1112: of cases we examined, the factor of $2 - 4$ underestimate in angle
1113: falls to a mere $10\% - 25\%$ offset one week prior to merger.  We find
1114: that the offset plateaus at this level, remaining at a few tens of percent
1115: up to 28 days before merger.
1116: 
1117: Accounting for this systematic final underestimate, we thus find
1118: K07's results to be a good baseline against which to compare our
1119: results.  This comparison makes it possible to assess the extent to
1120: which spin precession improves our ability to locate massive black
1121: hole binaries prior to the final merger.
1122: 
1123: \subsection{Results I: Time evolution of localization accuracy}
1124: \label{sec:timeevolve}
1125: 
1126: We finally come to the main results of this paper, the time evolution
1127: of our ability to localize MBHB systems using GWs when spin precession
1128: is included.  The results summarized in \S\ {\ref{sec:review}}
1129: describe the size of the GW pixel (sky position error ellipse and
1130: luminosity distance) at the end of inspiral.  We do not at this time
1131: incorporate any information regarding the merger and ringdown phases.  
1132: The end-of-inspiral\footnote{Throughout this
1133: paper, ``final'' accuracy refers to the end of inspiral.} localization
1134: accuracies are good enough that searching the GW pixel for
1135: counterparts to MBHB coalescences is likely to be fruitful.  However,
1136: given how little is understood about electromagnetic counterparts to
1137: these events, it is unclear if waiting until these final moments is
1138: the best strategy for such a search.  It will surely be desirable to
1139: also monitor the best-guess location some days or weeks in advance
1140: for electromagnetic precursors to the final merger.  The rate at which
1141: the GW pixel evolves as we approach the merger will have strong
1142: implications for determining the rate at which {\it LISA} data is sent
1143: to the ground.
1144: 
1145: To this end, we now examine the time dependence of the {\em LISA}
1146: pixel.  We have modified the code from Paper I to stop the calculation
1147: at a specified time before the fiducial merge frequency, equation
1148: \eqref{eq:fmerge}.  We begin the evolution of each binary at the
1149: moment it enters the {\it LISA} band (which we take to occur at
1150: $f_{\rm low} = 3 \times 10^{-5}\,{\rm Hz}$).  Because we randomly
1151: distribute $t_c$ over our (assumed) 3 yr {\it LISA} mission, some
1152: sources are already in band at the mission's start; consequently,
1153: these sources begin at $f > f_{\rm low}$.  The binary's evolution is
1154: then followed until it reaches a GW frequency $f_{\mathrm{stop}} =
1155: f(t(f_{\mathrm{merge}})-N)$, where $N$ is the number of days before
1156: merger that we want to stop the signal.  (Choosing $N = 0$ duplicates
1157: the analysis of Paper I.)
1158: 
1159: \begin{figure}[!ht]
1160: \begin{center}
1161: \includegraphics[scale=0.9]{pixels.pdf}
1162: \caption{Evolution of the sky position error ellipse for nine
1163: individual binaries selected from a set of $10^4$.  All have $m_1 =
1164: 10^6 M_\odot$, $m_2 = 3\times 10^5 M_\odot$, and $z = 1$.  The
1165: ellipses are oriented so their major axes are parallel to the $x$-axis
1166: and their minor axes are parallel to the $y$-axis; the axes are
1167: labeled in arcminutes.  From outside in, the ellipses are evaluated at
1168: 28, 21, 14, 7, 4, 2, 1, and 0 days before merger.}
1169: \label{fig:pixels}
1170: \end{center}
1171: \end{figure}
1172: 
1173: Figure \ref{fig:pixels} shows the error ellipse evolution for nine
1174: examples taken from a sample of $10^4$ computed with $m_1 = 10^6
1175: M_\sun$, $m_2 = 3\times 10^5 M_\sun$ and $z = 1$.  We show results for
1176: $N = $ 0, 1, 2, 4, 7, 14, 21, and 28.  For each binary, the major axis
1177: is plotted on the $x$-axis, while the minor axis is plotted on the
1178: $y$-axis; we do not show how each ellipse would be oriented on the
1179: sky.  The results shown in Figures \ref{fig:pixels}{\it a} and 
1180: \ref{fig:pixels}{\it b} were selected by hand
1181: from the distribution as examples of contrasting behavior.  The binary 
1182: in Figure \ref{fig:pixels}{\it a}
1183: shows a dramatic change in the error ellipse with time, especially in
1184: the last day before merger.  In that day, the binary is localized to
1185: an ellipse with $2a = 6.67^\prime$ and $2b = 6.25^\prime$, an area
1186: $\sim 60$ times smaller than at $N = 1$.  By contrast, 
1187: the binary in Figure \ref{fig:pixels}{\it b}
1188: shows almost no change in the error ellipse over the entire four weeks
1189: prior to merger.
1190: 
1191: These are clearly extreme cases.  Other extremes exist, including
1192: binaries with a minor axis orders of magnitude smaller than the major
1193: axis (see the tail in Fig.\
1194: \ref{fig:final_axes}, {\it right}), binaries where the evolution of one or both
1195: axes is not strictly monotonic, and binaries which have very large
1196: ellipses (essentially filling the sky) for large $N$.  (Such cases
1197: correspond to binaries which are already well into the {\it LISA} band
1198: when the mission starts; for large $N$, there is little baseline for
1199: the various modulations to encode their position.)  To get a sense of
1200: more typical behavior, we selected the binaries in Figures 
1201: \ref{fig:pixels}{\it c} -- \ref{fig:pixels}{\it i} randomly
1202: from the distribution.  There does not appear to be any ``typical''
1203: evolution; each binary exhibits some unique features.  Most binaries,
1204: however, seem to share with the binary in Figure 
1205: \ref{fig:pixels}{\it a} the property that the final day
1206: before merger gives much more information on the position than any day
1207: before it (albeit to a lesser degree).  We will see below that this
1208: feature holds for the medians of almost all mass and redshift cases.
1209: It is worth noting that although K07 agree with us on most of the
1210: other qualitative features of the time dependence, they do not see the
1211: dramatic change in the final day of inspiral.  As we will discuss in
1212: more detail below, this dramatic improvement toward the end of
1213: inspiral is due to spin precession physics.  This is in good agreement
1214: with the expectations of N. Cornish (2005, unpublished) and K07 that 
1215: spin precession
1216: would most dramatically impact the last week or so of inspiral.
1217: 
1218: \begin{figure}[htb]
1219: \begin{center}
1220: \includegraphics[scale=0.49]{2aevol.pdf}
1221: \includegraphics[scale=0.49]{2bevol.pdf}
1222: \caption{Same as Fig.\ \ref{fig:final_axes}, but with 
1223: $m_1 = 10^6 M_\odot$, $m_2 = 3 \times 10^5 M_\odot$, and $z = 1$ at
1224: different values of $N$ (the number of days before merger).  Reading
1225: from left to right, $N = 0$ ({\it solid line}), 1 ({\it dashed line}), 
1226: 2 ({\it dash-dotted line}), 4 ({\it dotted line}), 
1227: 7 ({\it solid line again}), 14 ({\it dashed line}), 
1228: 21 ({\it dash-dotted line}), and 28 ({\it dotted line}).  
1229: Clearly, the largest change --- in
1230: shape as well as median --- happens between merger and one day before.}
1231: \label{fig:axes_evol}
1232: \end{center}
1233: \end{figure}
1234: 
1235: While the evolution of parameter errors for individual binaries is
1236: interesting, of more relevance is the evolution of the errors'
1237: distribution.  The left panel of Figure \ref{fig:axes_evol} shows
1238: the time dependence of the distribution of the major axis $2a$ for our
1239: model system of $m_1 = 10^6 M_\odot$, $m_2 = 3\times 10^5 M_\odot$,
1240: and $z = 1$.  We can clearly see the evolution to smaller major axis
1241: as the binary nears merger.  Four weeks before merger, the median
1242: major axis is $4.8$ times larger than at merger; this number shrinks
1243: to $3.9$ two weeks before merger, $3.2$ one week before, and $2.5$ two
1244: days before.  As expected from the individual binaries, the most
1245: dramatic change in the distribution occurs during the last day before
1246: merger.  Not only is the median substantially reduced (by a factor of
1247: $2.2$), but the shape sharply changes.  For $N > 0$, the distribution
1248: is distinctly peaked, becoming gradually flatter as $N$ gets smaller.
1249: Over the last day of inspiral, the distribution evolves into the
1250: almost entirely flat, slightly bimodal shape first seen in
1251: Figure \ref{fig:final_axes}.  We find that this same behavior holds for
1252: all masses and redshift cases of interest: A sharply peaked
1253: distribution at $N > 0$ evolves into the flatter, sometimes bimodal
1254: final distributions described in \S\ \ref{sec:review}.  As the total
1255: mass increases, however, the final distributions become so narrow that
1256: the shape change is no longer very clear.  As we might expect, this
1257: transition occurs at smaller total mass for higher $z$.
1258: 
1259: The right panel of Figure \ref{fig:axes_evol} shows the evolution
1260: of the minor axis $2b$.  Again the distribution slowly changes shape
1261: over time, with the most drastic change occurring in the last day.
1262: Here the final distribution still retains a slight peak, along with
1263: the previously discussed long tail of small errors.  Interestingly,
1264: this tail is present to some degree throughout the evolution.  As the
1265: total mass increases, the final distribution moves to the right until,
1266: as with $2a$, the sharp change of shape disappears.  The same
1267: evolution occurs at higher $z$, again with a shift in the mass scale.
1268: 
1269: \begin{figure}[!p]
1270: \includegraphics[scale=0.49]{2amedians1.pdf}
1271: \includegraphics[scale=0.49]{2bmedians1.pdf}
1272: \includegraphics[scale=0.49]{2amedians2.pdf}
1273: \includegraphics[scale=0.49]{2bmedians2.pdf}
1274: \includegraphics[scale=0.49]{2amedians3.pdf}
1275: \includegraphics[scale=0.49]{2bmedians3.pdf}
1276: \caption{Medians of the sky position ellipse axes for Monte Carlo runs
1277: of $10^4$ binaries as a function of time before merger.  Major axes
1278: $2a$ are on the left; minor axes $2b$ are on the right.  Data were only
1279: outputted at the marked points; the lines are there just to guide the
1280: eye.  The masses have been subdivided into ``low,'' ``intermediate,'' and
1281: ``high'' groups; the exact values (in units of solar masses) are given in
1282: the legends.  Note also the different scales (arcminutes and degrees)
1283: on the $y$-axis.}
1284: \label{fig:ellipsemedians} 
1285: \end{figure}
1286: 
1287: To further understand how the error ellipse evolves, consider Figure
1288: {\ref{fig:ellipsemedians}}.  Here we show the evolution of the median
1289: axes for a wide range of masses (``low,'' ``intermediate,'' and
1290: ``high'') at $z = 1$.  As in Figures {\ref{fig:pixels}} and
1291: {\ref{fig:axes_evol}}, our calculations produce output for $N = $ 0,
1292: 1, 2, 4, 7, 14, 21, and 28; they are connected by lines only to guide
1293: the eye.  Almost all the cases we present show similar behavior: The
1294: ellipses gradually shrink with time before sharply decreasing in size
1295: during the final day.  Significant deviations from this behavior come
1296: from the high-mass binaries, which evolve more drastically at large
1297: $N$ before settling in to resemble the lower mass curves.  This
1298: high-mass deviation is an artifact of our choice of maximum $N$.
1299: Smaller binaries may spend many months or even years in the {\it LISA}
1300: band before merger.  In those cases, enough signal has already been
1301: measured at $N = 28$ to locate the binary reasonably well.  By
1302: contrast, the high-mass binaries spend much less time in
1303: band\footnote{The two highest mass binaries in Fig.\ \ref{fig:ellipsemedians}
1304: are not even
1305: in band a full 28 days.} and have not been measured so well by $N =
1306: 28$.  They have to ``catch up'' to the smaller mass binaries over the
1307: first few weeks of our measurement window.  Nearly identical results
1308: were found by K07: Figure 2 of K07 plots the evolution of
1309: sky position for an intermediate-mass binary from $N \simeq 300$.
1310: They find that the measurement accuracy rapidly evolves early in the
1311: measurement, with slopes of angular error versus time very similar to
1312: what we show in Figure {\ref{fig:ellipsemedians}} for high-mass
1313: binaries.
1314: 
1315: Although these curves are qualitatively quite similar, there are
1316: significant quantitative differences.  For example, the evolution of
1317: the intermediate-mass binaries is less than that of the low-mass
1318: binaries, especially in the last day.  The intermediate-mass major
1319: axes shrink by a factor of $\sim 4-6$ over the entire four-week period,
1320: whereas the low-mass axes shrink by a factor of $\sim 5-11$.  Another important
1321: quantitative difference can be seen by comparing the major and minor
1322: axes.  We find that the ratio $2a/2b$ grows with time in most cases,
1323: indicating that the minor axis tends to shrink more rapidly than the
1324: major axis.  The only exceptions are the previously described
1325: high-mass cases, in which $2a/2b$ shrinks during most or all of the
1326: inspiral.  Presumably the same behavior would also be seen for
1327: lower mass binaries at higher values of $N$; this conclusion is
1328: supported by Figure 4 of {\cite{khm07}}.  For all other cases, $2a/2b$
1329: $\sim 1.3-1.7$ at $N = 28$ and increases to $\sim 1.8-4$ at $N =
1330: 0$.  The largest increase is typically in the final day.
1331: 
1332: What causes this dramatic improvement in the last day of inspiral?
1333: Three factors primarily contribute to our ability to localize a source
1334: on the sky: modulations due to {\it LISA}'s orbital motion,
1335: modulations due to spin precession, and S/N accumulated over time.
1336: Since {\it LISA} moves the same amount in the final day as in any
1337: other day, orbital-induced modulations cannot be the cause.  A great
1338: deal of S/N is accumulated in the last day (typically increasing by a
1339: factor of $\sim 2$ or more), and many parameter errors scale as
1340: (S/N)$^{-1}$.  However, K07 demonstrate that sky position and distance
1341: errors do not scale as (S/N)$^{-1}$ in the last few weeks before merger.
1342: Our ``no precession'' code supports their conclusion: The final jump
1343: in S/N cannot make up for the lack of orbital modulation over such a
1344: short timescale.
1345: 
1346: The remaining possibility is spin-induced precession.  Indeed, we
1347: found in Paper I that the number of modulations due to spin precession
1348: increases dramatically as the binary approaches merger.  This suggests
1349: that the improvement we see is due to the impact of precession.  To
1350: examine this hypothesis, we plot the ``precession'' and ``no
1351: precession'' results together on the same axes.  Figure
1352: \ref{fig:NPcomparison} shows such a plot for a low-mass system and an
1353: intermediate-mass system at $z = 1$.  We see that at $N = 28$ days,
1354: the two codes give similar results for localization accuracy.  Their
1355: predictions gradually diverge as merger is approached.  The greatest
1356: jump between the two codes occurs on the last day before merger,
1357: agreeing with our expectation that precession effects are maximal
1358: then.  The effect is greater in the low-mass case than in the
1359: intermediate-mass case.  Similarly, the ratio $2a/2b$ starts about the
1360: same in both codes, growing very slowly until $N = 1$.  At this point,
1361: it jumps dramatically in the precession code, while staying
1362: roughly the same in the no precession code.  Interestingly, one
1363: effect of precession is that the localization errors track (S/N)$^{-1}$
1364: rather closely.  By breaking the various correlations which made them
1365: deviate from (S/N)$^{-1}$, precession-induced modulations allow the
1366: errors to evolve in a manner that is more consistent with our naive
1367: expectations.
1368: 
1369: \begin{figure}[htb]
1370: \includegraphics[scale=0.49]{NPcomparisonlow.pdf}
1371: \includegraphics[scale=0.49]{NPcomparisonint.pdf}
1372: \caption{Medians of $2a$ and $2b$ as a function of time, comparing an
1373: analysis that accounts for spin-induced precession to one that
1374: neglects it.  Solid lines trace the evolution of $2a$, dashed lines
1375: trace $2b$.  Precession results are marked with crosses, no
1376: precession with circles.  The left plot shows a low-mass case,
1377: $m_1 = 3\times 10^5 M_\odot$ and $m_2 = 10^5 M_\odot$; the right plot
1378: shows an intermediate-mass case, $m_1 = m_2 = 10^6 M_\odot$.  Both
1379: plots are for $z = 1$.}
1380: \label{fig:NPcomparison}
1381: \end{figure}
1382: 
1383: \begin{figure}[!ht]
1384: \begin{center}
1385: \includegraphics[scale=0.52]{muNphiNcovar.pdf}
1386: \caption{Off-diagonal covariance matrix entries illustrating
1387: correlation between sky position and binary orientation as a function
1388: of time, for the binary in Fig.\ \ref{fig:pixels}{\it a}.  The correlations
1389: decrease rapidly in the final day before merger, when precession
1390: effects are maximal.}
1391: \label{fig:muNphiNcovar}
1392: \end{center}
1393: \end{figure}
1394: 
1395: The time evolution of the correlations of interest, those between the
1396: sky position and binary orientation, are illustrated in Figure
1397: \ref{fig:muNphiNcovar}.  Here we show the off-diagonal components of
1398: the covariance matrix $\Sigma^{ab}$ (where $a \in
1399: \{\bar{\mu}_N,\bar{\phi}_N\}$ and $b \in
1400: \{\bar{\mu}_L(0),\bar{\phi}_L(0)\}$) for the binary in Figure
1401: \ref{fig:pixels}{\it a}.  We found that examining the normalized correlation
1402: coefficients $c^{ab} = \Sigma^{ab}(\Sigma^{aa}\Sigma^{bb})^{-1/2}$ can
1403: mislead since $\Sigma^{aa}$ and $\Sigma^{bb}$ are rapidly evolving at
1404: the same time as $\Sigma^{ab}$.  At large $N$, the two sets of angles
1405: are relatively strongly correlated due to degeneracies in the measured
1406: waveform (\ref{eq:freqdomainsignal}). However, in the last day before
1407: merger, the correlations sharply decrease as precession effects
1408: accumulate.  The reduction of these correlations coincides with the
1409: sudden drop in parameter errors seen in Figure \ref{fig:pixels} (and,
1410: by extension to the entire Monte Carlo run, Figs.\ \ref{fig:axes_evol}
1411: -- \ref{fig:NPcomparison}).
1412: 
1413: Finally, we investigate the time evolution of errors in the luminosity
1414: distance $D_L$.  Figure \ref{fig:Devol} shows the distribution of
1415: $\Delta D_L/D_L$ (determined solely by taking into account GW measurement
1416: effects) evolving in time for a binary with $m_1 = 10^6 M_\sun$, $m_2
1417: = 3 \times 10^5 M_\sun$, and $z = 1$.  We see that in contrast to sky
1418: position, the shape of the distribution does not change very much with
1419: time.  It typically spreads enough to reduce its height, but it maintains
1420: a well-defined peak.  However, the progression of the median is very
1421: similar to the sky position case: It decreases slowly with time until the
1422: last day, when it jumps drastically.  The evolution of median values
1423: of $\Delta D_L/D_L$ follows tracks very similar in shape to those
1424: shown in Figure {\ref{fig:ellipsemedians}}, so we do not show them
1425: explicitly. 
1426: 
1427: \begin{figure}[!ht]
1428: \begin{center}
1429: \includegraphics[scale=0.52]{Devol.pdf}
1430: \caption{Distribution of $\Delta D_L/D_L$ for $10^4$ binaries with $m_1 =
1431: 10^6 M_\odot$, $m_2 = 3 \times 10^5 M_\odot$, and $z = 1$ at different
1432: values of $N$ (the number of days before merger).  Reading from left to
1433: right, $N = 0$ ({\it solid line}), 1 ({\it dashed line}), 
1434: 2 ({\it dash-dotted line}), 4 ({\it dotted line}), 7 ({\it solid line again}), 
1435: 14 ({\it dashed line}), 21 ({\it dash-dotted line}), and
1436: 28 ({\it dotted line}).}
1437: \label{fig:Devol}
1438: \end{center}
1439: \end{figure}
1440: 
1441: Table {\ref{tab:positionsummary}} summarizes all of our results on the
1442: time evolution of localization.  At low redshift, the ability to
1443: locate an event on the sky is quite good over much of the mass range
1444: even as much as a month in advance of the final merger.  In most
1445: cases, the localization ellipse at $z = 1$ is never larger than about
1446: $10\ \mathrm{deg}^2$ in size, which is comparable to the field of view
1447: of proposed future surveys, such as LSST \citep{t02}.  This
1448: ability degrades fairly rapidly as redshift increases, especially for
1449: larger masses.  At $z = 3$, the ellipse can be $\sim 10\ \mathrm{deg}^2$
1450: a few days in advance of merger for small and intermediate
1451: masses.  At the highest masses, GWs provide very little localization
1452: information.  Going to $z = 5$ makes this even worse; an ellipse of
1453: $\sim 10\ \mathrm{deg}^2$ can be found at most a day prior to merger,
1454: and only for relatively small mass ranges.
1455: 
1456: In many cases, $\Delta D_L/D_L$ is determined so well by GWs that
1457: gravitational lensing errors are expected to dominate.  As such, the
1458: GW-determined values of $\Delta D_L/D_L$ are essentially irrelevant
1459: for locating these binaries in redshift space; lensing will instead
1460: determine how well redshifts can be measured.  However, in some cases,
1461: the intrinsic distance error exceeds the lensing error, so it is worth
1462: knowing when the ``lensing limit'' can be achieved.  At $z = 1$, the
1463: limit is achieved for most binaries as long as a month before merger.
1464: For $z = 3$, the limit is only achieved a few days to a week
1465: (depending on mass) before merger; for $z = 5$, intrinsic errors
1466: generally exceed the lensing errors even at a day before merger.
1467: 
1468: 
1469: \begin{table}[p]
1470: \caption{Summary of time evolution of localization accuracy}
1471: \begin{tabular}{llccccc}
1472: \\
1473: \tableline \tableline
1474: $z$ &
1475: Binary Range &
1476: $N$ (days) &
1477: $2a$ (deg) &
1478: $2b$ (deg) &
1479: $\Delta \Omega_N\ (\mathrm{deg}^2)$&
1480: $\Delta D_L/D_L$ \\
1481: 
1482: \tableline
1483: 
1484: $1$ & Low ($M \lesssim 10^6\,M_\odot$) &
1485: 
1486: $0$ & $(17 - 27)$\tablenotemark{a} & $(6.6 - 13)$\tablenotemark{a} & $0.02 - 0.07$ & $(2.4 - 4)\times 10^{-3}$\\
1487:  &  & $1$ & $(71 - 140)$\tablenotemark{a} & $(26 - 91)$\tablenotemark{a} & $0.3 - 2.8$ & $(8 - 21) \times 10^{-3}$ \\
1488:  &  & $7$ & $(100 - 180)$\tablenotemark{a} & $(54 - 120)$\tablenotemark{a} & $0.9 - 5.1$ & $(1.3 - 3) \times 10^{-2}$ \\
1489:  &  & $28$ & $(140 - 240)$\tablenotemark{a} & $(94 - 180)$\tablenotemark{a} & $2.6 - 9.6$ & $(2 - 4.3) \times 10^{-2}$ \\
1490: 
1491: \tableline
1492: 
1493: & Int. ($10^6\,M_\odot \lesssim M \lesssim 4\times 10^6\,M_\odot$) & $0$ & $(31 - 40)$\tablenotemark{a} & $(9.2 - 22)$\tablenotemark{a} & $0.04 - 0.18$ & $(3.8 - 5.6) \times 10^{-3}$\\
1494: 
1495: & & $1$ & $(56 - 75)$\tablenotemark{a} & $(20 - 48)$\tablenotemark{a} & $0.17 - 0.77$ & $(6.4 - 11) \times 10^{-3}$ \\
1496:  &  & $7$ & $(85 - 110)$\tablenotemark{a} & $(42 - 74)$\tablenotemark{a} & $0.6 - 1.9$ & $(1.1 - 1.7) \times 10^{-2}$ \\
1497:  &  & $28$ & $(130 - 190)$\tablenotemark{a} & $(80 - 130)$\tablenotemark{a} & $2 - 5.4$ & $(1.8 - 3) \times 10^{-2}$ \\
1498: 
1499: \tableline
1500: 
1501: & High ($M \gtrsim 6\times 10^6\,M_\odot$) &
1502: 
1503: $0$ & $0.6 - 1.4$ & $0.2 - 0.64$ & $0.07 - 0.68$ & $(4.6 - 25)\times 10^{-3}$ \\
1504:  &  & $1$ & $1 - 4.7$ & $0.4 - 2$ & $0.2 - 7.1$ & $(0.7 - 16)\times 10^{-2}$ \\
1505:  &  & $7$\tablenotemark{b} & $2.1 - 6.2$ & $0.9 - 3$ & $1 - 14$ & $(1.6 - 8.7)\times 10^{-2}$ \\
1506:  &  & $28$\tablenotemark{b} & $7.1 - 23$ & $3.4 - 8.7$ & $16 - 150$ & $(8.4 - 30)\times 10^{-2}$ \\
1507: 
1508: \tableline
1509: \tableline
1510: 
1511: $3$ & Low ($M \lesssim 10^6\,M_\odot$) &
1512: 
1513: $0$ & $1.3 - 2.4$ & $0.6 - 1.3$ & $0.7 - 2.1$ & $(1.2 - 2)\times 10^{-2}$\\
1514:  &  &  $1$ & $3.9 - 7.2$ & $1.4 - 4.7$ & $3 - 27$ & $(2.7 - 6.3) \times 10^{-2}$ \\
1515:  &  &  $7$ & $5.8 - 9.6$ & $2.9 - 6.5$ & $10 - 51$ & $(4.4 - 9.3) \times 10^{-2}$ \\
1516:  &  & $28$ & $8.5 - 13$ & $5.4 - 9.8$ & $32 - 100$ & $(7.2 - 14) \times 10^{-2}$ \\
1517: 
1518: \tableline
1519: 
1520: & Int. ($10^6\,M_\odot \lesssim M \lesssim 4\times 10^6\,M_\odot$) &
1521: 
1522: $0$ & $2.1 - 2.6$ & $0.7 - 1.4$ & $0.8 - 2.6$ & $(1.5 - 2.4) \times 10^{-2}$\\
1523: 
1524: & & $1$ & $3.6 - 5$ & $1.3 - 3.1$ & $2.4 - 12$ & $(2.5 - 4.5) \times 10^{-2}$ \\
1525:  &  & $7$ & $6 - 11$ & $2.8 - 6.3$ & $9.8 - 48$ & $(4.5 - 10) \times 10^{-2}$ \\
1526:  &  & $28$ & $9.8 - 91$ & $6.3 - 42$ & $43 - 2900$ & $(0.9 - 14) \times 10^{-1}$ \\
1527: 
1528: \tableline
1529: 
1530: & High ($M \gtrsim 6\times 10^6\,M_\odot$) &
1531: 
1532: $0$ & $3.4 - 33$ & $1.3 - 9.7$ & $2.7 - 260$ & $(2.9 - 240)\times 10^{-2}$ \\
1533:  &  & $1$\tablenotemark{b} & $5.7 - 17$ & $2.4 - 7.3$ & $8.5 - 93$ & $(5.2 - 53)\times 10^{-2}$ \\
1534:  &  & $7$\tablenotemark{b} & $25 - 75$ & $13 - 27$ & $220 - 1500$ & $(5.8 - 19)\times 10^{-1}$ \\
1535:  &  & $28$\tablenotemark{c} & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ \\
1536: 
1537: \tableline
1538: \tableline
1539: 
1540: $5$ & Low ($M \lesssim 10^6\,M_\odot$) &
1541: 
1542: $0$ & $2.8 - 4.9$ & $1.1 - 2.7$ & $2.3 - 9.3$ & $(2.6 - 4.1)\times 10^{-2}$\\
1543:  &  & $1$ & $6.8 - 12$ & $2.4 - 7.6$ & $9 - 71$ & $(4.6 - 10) \times 10^{-2}$ \\
1544:  &  & $7$ & $10 - 16$ & $5 - 11$ & $32 - 140$ & $(7.8 - 15) \times 10^{-2}$ \\
1545:  &  & $28$ & $16 - 22$ & $9.6 - 16$ & $100 - 290$ & $(1.3 - 2.3) \times 10^{-1}$ \\
1546: 
1547: \tableline
1548: 
1549: & Int. ($10^6\,M_\odot \lesssim M \lesssim 4\times 10^6\,M_\odot$) &
1550: 
1551: $0$ & $3.9 - 5.2$ & $1.4 - 2.7$ & $3.3 - 10$ & $(3.2 - 5)\times 10^{-2}$ \\
1552: 
1553: & & $1$ & $7.2 - 11$ & $2.5 - 6.2$ & $9.6 - 48$ & $(5.2 - 10)\times10^{-2}$ \\
1554:  &  & $7$ & $12 - 41$ & $5.9 - 21$ & $45 - 610$ & $(1.1 - 5.2)\times10^{-1}$ \\
1555:  &  & $28$\tablenotemark{b} & $21 - 170$ & $15 - 65$ & $250 - 8000$ & $(2 - 21)\times10^{-1}$ \\
1556: 
1557: \tableline
1558: 
1559: & High ($M \gtrsim 6\times 10^6\,M_\odot$) &
1560: 
1561: $0$\tablenotemark{b} & $9 - 29$ & $4.3 - 10$ & $30 - 230$ & $(1.2 - 12)\times10^{-1}$ \\
1562:  &  & $1$\tablenotemark{b} & $19 - 27$ & $9.9 - 12$ & $130 - 250$ & $(4.4 - 7.3)\times10^{-1}$ \\
1563:  &  & $7$\tablenotemark{c} & $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ \\
1564: % &  & $28$\tablenotemark{c}& $\cdots$ & $\cdots$ & $\cdots$ & $\cdots$ \\
1565: \tableline
1566: \end{tabular}
1567: 
1568: \tablecomments{Typical ranges of sky position and distance measurement accuracy as a function
1569: of time until merger for low-, intermediate-, and high-mass binaries.  Angles are in degrees unless otherwise noted;
1570: solid angles are in square degrees.}
1571: 
1572: \tablenotetext{a}{These angles are in arcminutes; all others
1573: are in degrees.}
1574: 
1575: \tablenotetext{b}{Some very massive systems are excluded from these
1576: data.  In those cases, the position and distance are very poorly constrained this far in advance of merger.  In some cases, the binary is even out of band.}
1577: 
1578: \tablenotetext{c}{All of the binaries of this mass and redshift are either very poorly measured or completely out of band this far in advance of merger.}
1579: 
1580: \label{tab:positionsummary}
1581: \end{table}
1582: 
1583: \subsection{Results II: Angular dependence of localization accuracy}
1584: \label{sec:angdependence}
1585: 
1586: We now examine one final interesting property of the errors: their
1587: dependence on the sky position of the source.  As we design future
1588: surveys to find counterparts to MBHB coalescences, it will be
1589: important to understand if there is a bias for good (or bad)
1590: localization in certain regions of the sky.  It is also useful to know
1591: in advance whether the ``best'' regions are likely to be blocked by
1592: foreground features such as the Galactic center.
1593: 
1594: Before discussing this dependence in detail, it is worth reviewing
1595: some details of how our Monte Carlo distributions are constructed.  As
1596: described in \S\ {\ref{sec:review}}, in most of our analysis we
1597: randomly distribute the sky position of our binaries, drawing from a
1598: uniform distribution in $\bar{\mu}_N = \cos\bar{\theta}_N$ and
1599: $\bar{\phi}_N$, where $\bar{\theta}_N$ and $\bar{\phi}_N$ are the
1600: polar and azimuthal angles of a binary in solar system barycenter
1601: coordinates.  We also randomly choose our binaries' final merger time.
1602: In this section, rather than distributing the sky position, we examine
1603: parameter accuracies for particular given positions; all other Monte
1604: Carlo parameters are distributed as usual.  Because we continue to
1605: randomly distribute the final merger time, the {\it relative} azimuth
1606: between a binary's sky position and {\it LISA}'s orbital position at
1607: merger, $\delta\phi = \bar{\phi}_N - \phi_{\it LISA}(t_c)$, remains
1608: randomly distributed.  As such, we expect our analysis to effectively
1609: average over $\bar{\phi}_N$, washing out any strong dependence on this
1610: angle in our analysis.
1611: 
1612: We begin by examining the dependence of errors on $\bar{\mu}_N$.  We
1613: evenly divide the range $-1 \le \bar{\mu}_N \le 1$ into 40 bins and
1614: run a Monte Carlo simulation with $10^4$ points in each.  That is, we
1615: pick $\bar{\mu}_N$ only from the bin range, but we pick all other
1616: random parameters in the usual manner.  The results for a
1617: representative binary ($m_1 = 10^6 M_\odot, m_2 = 3 \times 10^5
1618: M_\odot$, and $z = 1$) are shown in Figure {\ref{fig:mudependence}}.
1619: 
1620: 
1621: \begin{figure}[t]
1622: \includegraphics[scale=0.32]{mudependence2a.pdf}
1623: \includegraphics[scale=0.32]{mudependence2b.pdf}
1624: \includegraphics[scale=0.32]{mudependenceD.pdf}
1625: \caption{Dependence of the localization errors on $\bar{\mu}_N$.
1626: The major axis $2a$ of the sky position error ellipse is on the left, the 
1627: minor axis $2b$ in the center, and the 
1628: luminosity distance errors $\Delta D_L/D_L$ on the right.  Each datum
1629: represents the median of $10^4$ binaries with $m_1 = 10^6 M_\sun$,
1630: $m_2 = 3 \times 10^5 M_\sun$, and $z = 1$; all other parameters are
1631: selected randomly (except for $\bar{\mu}_N$, whose range is limited to
1632: the bin width).}
1633: \label{fig:mudependence}
1634: \end{figure}
1635: 
1636: Note that all of the error distributions are symmetrically peaked
1637: around the plane of {\em LISA}'s orbit ($\bar\mu_N = 0$).  Any slight
1638: asymmetry is due only to statistical effects.  This is reassuring;
1639: {\em LISA} should not favor one hemisphere over the other.  There is
1640: also additional structure that is parameter dependent.  At its peak,
1641: the major axis $2a$ is almost $35 \%$ greater than the
1642: position-averaged median value\footnote{Note that the
1643: position-averaged medians quoted here are slightly different from
1644: those quoted in Table 1, since this sample has 40 times more points.}
1645: of $31.1^\prime$.  It then decreases with $|\bar{\mu}_N|$ and
1646: reaches a minimum of about $25^\prime$ for $0.75 < |\bar{\mu}_N| <
1647: 0.8$.  Finally, there are subpeaks near the ecliptic poles, although
1648: they still lie below the position-averaged median.  The dependence of
1649: the minor axis $2b$ on angle is even more dramatic.  At its peak, $2b$
1650: differs from the position-averaged median of $13.0^\prime$ by over
1651: $85\%$.  Just like the major axis, it drops to a minimum, but it does
1652: so more rapidly.  The minimum also occurs at a slightly larger value of $|\bar{\mu}_N|$
1653: than for $2a$.  The minor axis also shows fairly strong subpeaks near
1654: the ecliptic poles, with values higher than the position-averaged
1655: median.
1656: 
1657: The luminosity distance errors behave slightly differently.  First of
1658: all, the variation with $\bar{\mu}_N$ is weaker than for the sky
1659: position: At the central peak, $\Delta D_L/D_L$ is only $\sim 15 \%$
1660: greater than its position-averaged median value ($0.00392$).  In
1661: addition, while the distribution again peaks near the poles, it does
1662: so at a smaller value of $|\bar{\mu}_N|$.  (In fact, the peaks occur
1663: very close to where sky position errors are {\it minimized}.)
1664: 
1665: By binning the data sets developed for Paper I, we were able to
1666: confirm this behavior over a wide range of masses and redshifts,
1667: albeit with poorer statistics: $10^4$ binaries in total, rather
1668: than per bin.  Thanks to the poorer statistics, we cannot
1669: resolve the small polar subpeaks in the distribution of $2a$.  In addition, in
1670: some cases it appears that the side peaks can be larger than the
1671: central peak in the distribution of $\Delta D_L/D_L$.  Aside from
1672: these minor variations, the shapes and relative amplitudes seen in
1673: these distributions hold robustly over all masses and redshifts we
1674: considered.
1675: 
1676: %To examine the role that precession plays in determining the
1677: %dependence of localization accuracy on $\bar\mu_N$, we performed the
1678: %same analysis using the ``no precession'' code.  We find similar
1679: %distributions for $2a$ and $2b$ (though somewhat broader); we also
1680: %find that $2a$ more closely resembles $2b$ at the poles.  $\Delta
1681: %D_L/D_L$ is quite a bit different: It is {\it minimized} in the
1682: %ecliptic plane and increases monotonically to the poles.  Careful
1683: %checking shows that the distributions including precession physics at
1684: %28 days before merger have the same shape as the distributions
1685: %neglecting precession do at merger; they then evolve to those in Fig.\
1686: %\ref{fig:mudependence} over time.
1687: 
1688: \begin{figure}[!t]
1689: \includegraphics[scale=0.32]{phidependence2a.pdf}
1690: \includegraphics[scale=0.32]{phidependence2b.pdf}
1691: \includegraphics[scale=0.32]{phidependenceD.pdf}
1692: \caption{Same as Fig. \ref{fig:mudependence}, but for dependence on $\bar{\phi}_N$ and with $10^5$ binaries in each bin rather than $10^4$.}
1693: \label{fig:phidependence}
1694: \end{figure}
1695: 
1696: We next investigate the dependence of the errors on the azimuthal
1697: angle $\bar{\phi}_N$.  To improve the statistics, we now calculate
1698: $10^5$ binaries in each bin; the results are shown in Figure
1699: \ref{fig:phidependence}.  The errors have a very weak (although
1700: nonzero) dependence on $\bar{\phi}_N$: The maximum deviation from the
1701: overall median is only $\sim 1\% - 2\%$.  This is to be expected; as
1702: discussed at the beginning of this section, the randomness of our
1703: binaries' merger times effectively averages over azimuth.  If we did
1704: not average over azimuth in this way, we would expect a moderately
1705: strong $\bar{\phi}_N$-dependence due to the functional form of {\it
1706: LISA}'s response to GWs.  Even after averaging this dependence away,
1707: we might expect some residual $\bar{\phi}_N$ structure due to the
1708: ``rolling'' motion of the {\it LISA} constellation.  The phase
1709: associated with {\it LISA}'s roll angle puts an additional oscillation
1710: on the measured waves (see the $\alpha$-dependence in eqs.\ [47] and
1711: [48] of K07, where $\alpha$ encodes the roll angle), and we do not
1712: average over this angle.  Indeed, on close inspection, we can make
1713: out roughly two peaks in each plot in Figure \ref{fig:phidependence}
1714: (although the statistics are still too poor to resolve them clearly),
1715: consistent with the $\cos2\alpha$ and $\sin2\alpha$ behavior shown in
1716: K07.  For comparison, we also examined the $\bar{\phi}_N$-dependence for
1717: this mass and redshift with precession ``turned off.''  The
1718: oscillatory behavior appears very clearly in this case.  Because of
1719: the weakness of the $\bar{\phi}_N$ behavior, we were unable to easily
1720: check it for other masses and redshifts.
1721: 
1722: \begin{figure}[!p]
1723: \begin{center}
1724: \includegraphics[scale=0.48]{2aecliptic.pdf}
1725: \includegraphics[scale=0.48]{2agalactic.pdf}
1726: \includegraphics[scale=0.48]{2becliptic.pdf}
1727: \includegraphics[scale=0.48]{2bgalactic.pdf}
1728: \includegraphics[scale=0.48]{OmSecliptic.pdf}
1729: \includegraphics[scale=0.48]{OmSgalactic.pdf}
1730: \includegraphics[scale=0.48]{Decliptic.pdf}
1731: \includegraphics[scale=0.48]{Dgalactic.pdf}
1732: \caption{Sky maps of major axis $2a$ ({\it top row}, in arcminutes), minor axis
1733: $2b$ ({\it second row}, in arcminutes), localization ellipse area $\Delta
1734: \Omega_N$ ({\it third row}, in square degrees) and $\Delta D_L/D_L$ 
1735: ({\it bottom row}) for {\em LISA} observations of MBHBs
1736: in different parts of the sky.  Data in the left column are 
1737: presented in ecliptic coordinates;
1738: data on the right are in Galactic coordinates, with the Galactic
1739: center at the middle.  Note that the level of $\bar\phi_N$ variation
1740: is so small that it would not show up in these figures; accordingly,
1741: they are essentially just remappings of Fig.\
1742: {\ref{fig:mudependence}}.}
1743: \label{fig:skymaps}
1744: \end{center}
1745: \end{figure}
1746: 
1747: Finally, to give an overall sense as to how localization varies on the
1748: sky, we present in Figure {\ref{fig:skymaps}} sky maps of the median
1749: major axis $2a$, minor axis $2b$, localization ellipse area $\Delta
1750: \Omega_N$, and distance accuracy $\Delta D_L/D_L$.  We show these data
1751: both in ecliptic and Galactic coordinates.  Note that most of the
1752: region of small error lies outside of the Galactic plane.  This
1753: potentially bodes well for searches for MBHB electromagnetic
1754: counterparts --- the regions where instruments like {\it LISA} ``see''
1755: most sharply are less likely to be hidden by foreground features.
1756: Certain portions of the sky that {\it LISA} sees well will be easier
1757: to search telescopically than others.  It will be an important task
1758: for future surveys over all electromagnetic bands to identify regions
1759: that are particularly amenable to finding counterparts to MBHB events.
1760: 
1761: \section{Summary and conclusions}
1762: \label{sec:disc}
1763: 
1764: As discussed at length in Paper I, accounting for the general
1765: relativistic precession of the angular momentum vectors in an MBHB system has a
1766: dramatic impact on what we can learn by observing the system's
1767: gravitational waves.  Spin-induced precession breaks degeneracies
1768: among different parameters, making it possible to measure them more
1769: accurately than they could be determined if precession were not
1770: present.  This has a particularly important impact on our ability to
1771: locate such a binary on the sky and to determine its luminosity
1772: distance --- the degeneracy between sky angles, distance, and
1773: orientation angles is severe in the absence of precession.
1774: 
1775: Our analysis shows that the improvement that precession imparts to
1776: measurement accumulates fairly slowly.  In using one code which includes
1777: the impact of spin precession and a second which neglects this effect,
1778: we find little difference in the accuracy with which GWs determine
1779: sky position and distance for times more than a few days in advance of the
1780: final merger.  The difference between the two codes grows quite
1781: rapidly in these final days.  In the last day alone, the localization 
1782: ellipse area decreases by a factor of $\sim 3-10$ (up to $\sim 60$ in a 
1783: few low-mass systems) when precession effects are included.  
1784: Distance determination is likewise improved by
1785: factors of $\sim 1.5-7$ in that final day.  
1786: 
1787: Not all of the precession effects occur in the final days.  We saw in
1788: Figure \ref{fig:axes_evol} that the long tail of small minor axes can be
1789: seen, to some degree, throughout the inspiral.  We could get lucky and
1790: find a binary with a very small value of $2b$ weeks before merger.
1791: But the improvement in the median that we found in Paper I appears to
1792: take effect only in the final days of inspiral.  Therefore, while
1793: precession may in fact help improve the {\it final} localization of a
1794: coalescing binary by a factor of $\sim 2-10$ in each direction, it
1795: will not be much help in {\it advanced} localization of a typical
1796: binary.
1797: 
1798: Nevertheless, the pixel sizes that we find are small enough that
1799: future surveys should not have too much trouble searching the region
1800: identified by GWs, at least over certain ranges of mass and redshift.
1801: At $z = 1$, the GW localization ellipse is $\sim 10\ \mathrm{deg}^2$ or
1802: smaller for most binaries as early as a month in advance of merger.
1803: (At high masses, the ellipse can be substantially larger than this a
1804: month before merger, but it shrinks rapidly, reaching a comparable
1805: size $1-2$ weeks before merger.)  This bodes well for future surveys
1806: with large fields of view that are likely to search the GW pixel for
1807: counterparts.  In addition, GWs determine the source luminosity
1808: distance so well that the distance errors we find are essentially
1809: irrelevant --- gravitational lensing will dominate the distance error
1810: budget for all but the highest masses.
1811: 
1812: As redshift increases, the GW pixel rapidly degrades, particularly for the
1813: largest masses.  Let us adopt $10\ \mathrm{deg}^2$ (the approximate LSST
1814: field of view) as a benchmark localization for which counterpart
1815: searches may be contemplated.  At $z = 3$, this benchmark is reached
1816: at merger for almost the entire range of masses we considered.  As
1817: little as a day in advance of merger, however, some of the least
1818: massive and most massive systems are out of this regime.  One week
1819: prior to merger, the most massive systems are barely located at all
1820: (ellipses hundreds of square degrees or larger).  The intermediate
1821: masses do best, but even in their cases the positions are determined
1822: with $\sim 10\ \mathrm{deg}^2$ accuracy no earlier than a few days in
1823: advance of merger.  The resolution degrades further at higher
1824: redshift.  At $z = 5$, systems with $M \gtrsim 6 \times 10^6\,M_\odot$
1825: are not located more accurately than $\sim 30\ \mathrm{deg}^2$ even at
1826: merger.  Smaller systems are located within $\sim 10\ \mathrm{deg}^2$
1827: at merger, but very few are at this accuracy even one day in advance
1828: of merger.  The luminosity distance errors also increase, so much that
1829: they exceed lensing errors a few days to a week before merger at $z =
1830: 3$, and only a day before merger at $z = 5$.  This degradation hurts
1831: the ability to search for counterparts by redshift and subsequently
1832: use them as standard candles.
1833: 
1834: Our main conclusion is that future surveys are likely to have good
1835: advanced knowledge (a few days to one month) of the location of MBHB
1836: coalescences at low redshift ($z \sim 1 - 3$), but only a day's notice
1837: at most at higher redshift ($z \sim 5$).  This conclusion may be
1838: excessively pessimistic.  As mentioned earlier, recent work examining
1839: the importance of subleading harmonics of MBHB GWs is finding that
1840: including harmonics beyond the leading quadrupole has an important
1841: effect on the final accuracy of position determination
1842: {\citep{arunetal,ts08}}.  For most masses, these analyses show a
1843: factor of a few improvement in position, comparable to the improvement
1844: that we find when spin precession is added to the waveform model.  For
1845: high-mass systems, the higher harmonics increase the (previously
1846: small) overlap with the {\it LISA} band; consequently, the improvement
1847: can be much larger, up to 2 or 3 orders of magnitude in area.
1848: Since these two improvements arise from very different physical
1849: effects, it is likely that their separate improvements can be combined
1850: for an overall improvement significantly better than each effect on
1851: its own.  We plan to test this in future work (which is just now
1852: getting underway).
1853: 
1854: Finally, we have also studied the sky position dependence of {\it
1855: LISA}'s ability to localize sources.  We have found that the regions
1856: of best localization lie fairly far out of the Galactic plane.
1857: However, as emphasized by N. Cornish (2007, private communication), a proper
1858: anisotropic confusion background might impact this dependence.  In our
1859: calculations, we have assumed an isotropic background, neglecting the
1860: likely spatial distribution of Galactic binaries.  Properly accounting
1861: for this background is likely to strengthen our conclusion that LISA's
1862: ability to ``see'' is best for MBHB sources out of the Galactic plane.
1863: 
1864: \acknowledgments
1865: 
1866: We are very grateful to Bence Kocsis for detailed discussions on this
1867: work.  We have also benefitted from discussions with Emanuele Berti,
1868: Neil Cornish, Zoltan Haiman, Bala Iyer, Kristen Menou, B.\
1869: Sathyaprakash, Rob Simcoe, Alicia Sintes, Michele Vallisneri, and
1870: Alberto Vecchio.  This work was supported by NASA grants NAGW-12906
1871: and NNG05G105G, as well as NSF grant PHY04-49884.  S.A.H. gratefully
1872: acknowledges the support of the MIT Class of 1956 Career Development
1873: fund, as well as the hospitality of the Harvard-Smithsonian Center for
1874: Astrophysics' Institute for Theory and Computation, where a portion of
1875: an early draft of this manuscript was written.
1876: 
1877: \begin{thebibliography}{99}
1878: 
1879: \bibitem[Apostolatos et al.(1994)]{acst94} Apostolatos, T.\ A.,
1880:   Cutler, C., Sussman, G.\ J., \& Thorne, K.\ S.\ 1994, \prd, 49, 6274
1881: 
1882: \bibitem[Armitage \& Natarajan(2002)]{an02} Armitage, P.\ J., \&
1883:   Natarajan, P.\ 2002, \apj, 567, L9
1884: 
1885: \bibitem[Arun et al.(2007)]{arunetal} Arun, K.\ G., Iyer, B.\ R.,
1886:   Sathyaprakash, B.\ S., Sinha, S., \& Van Den Broeck, C.\ 2007, \prd,
1887:   76, 104016
1888: 
1889: \bibitem[Berti et al.(2005)]{bbw05} Berti, E., Buonanno,
1890:   A., \& Will, C.\ M.\ 2005, \prd, 71, 084025
1891: 
1892: \bibitem[Blanchet(2006)]{b06} Blanchet, L.\ 2006, Living Rev.\
1893:   Relativ., 9, 4
1894: 
1895: \bibitem[Blanchet et al.(1995)]{bdiww95} Blanchet, L., Damour, T.,
1896:   Iyer, B.\ R., Will, C.\ M., \& Wiseman, A.\ G.\ 1995, \prl, 74,
1897:   3515
1898: 
1899: \bibitem[Bode \& Phinney(2007)]{bp07} Bode, N., \& Phinney, E.\ S.\ 2007, APS Abstr., http://meetings.aps.org/link/BAPS.2007.APR.S1.10
1900: 
1901: \bibitem[Buonanno et al.(2007)]{bcp07} Buonanno, A.,
1902:   Cook, G.\ B., \& Pretorius, F.\ 2007, \prd, 75, 124018
1903: 
1904: \bibitem[Cutler(1998)]{c98} Cutler, C.\ 1998, \prd, 57, 7089
1905: 
1906: \bibitem[Cutler \& Flanagan(1994)]{cf94} Cutler, C., \& Flanagan, \'E.\
1907:   \'E.\ 1994, \prd, 49, 2658
1908: 
1909: \bibitem[Dalal et al.(2003)]{dhcf03} Dalal, N., Holz, D.\ E., Chen,
1910:   X., \& Frieman, J.\ A.\ 2003, \apjl, 585, L11
1911: 
1912: \bibitem[Dotti et al.(2006)]{dssch06} Dotti, M., Salvaterra, R.,
1913:   Sesana, A., Colpi, M., \& Haardt, F.\ 2006, \mnras, 372, 869
1914: 
1915: \bibitem[Farmer \& Phinney(2003)]{fp03} Farmer, A.\ J., \& Phinney, E.\
1916:   S.\ 2003, \mnras, 346, 1197
1917: 
1918: \bibitem[Ferrarese \& Merritt(2000)]{fm00} Ferrarese, L., \& Merritt,
1919:   D.\ 2000, \apj, 539, L9
1920: 
1921: \bibitem[Finn(1992)]{f92} Finn, L. S.\ 1992, \prd, 46, 5236 
1922: 
1923: \bibitem[Frieman(1997)]{frieman97} Frieman, J.\ A.\ 1997, Comments
1924:   Astrophys., 18, 323
1925: 
1926: \bibitem[Gebhardt et al.(2000)]{g00} Gebhardt, K., et al.\ 2000, \apj,
1927:   539, L13
1928: 
1929: \bibitem[Hellings \& Moore(2003)]{hm03} Hellings, R.\ W., \& Moore,
1930:   T.\ A.\ 2003, Classical Quantum Gravity, 20, S181
1931: 
1932: \bibitem[Holz(1998)]{h98} Holz, D.\ E.\ 1998, \apjl, 506, L1
1933: 
1934: \bibitem[Holz \& Hughes(2005)]{hh05} Holz, D.\ E., \& Hughes, S.\ A.\
1935:   2005, \apj, 629, 15
1936: 
1937: \bibitem[Holz \& Wald(1998)]{hw98} Holz, D.\ E., \& Wald, R.\ M.\ 1998,
1938:   \prd, 58, 063501
1939: 
1940: \bibitem[Hughes(2002)]{h02} Hughes, S.\ A.\ 2002, \mnras, 331, 805
1941: 
1942: 
1943: \bibitem[Kidder(1995)]{k95} Kidder, L.\ E.\ 1995, \prd, 52, 821
1944: 
1945: \bibitem[Kocsis et al.(2006)]{kfhm06} Kocsis, B., Frei, Z., Haiman,
1946:   Z., \& Menou, K.\ 2006, \apj, 637, 27
1947: 
1948: \bibitem[Kocsis et al.(2007a)]{khm07} Kocsis, B., Haiman, Z., \&
1949:   Menou, K.\ 2007a, \apj, submitted (arXiv:0712.1144)
1950: 
1951: \bibitem[Kocsis et al.(2007b)]{khmf07} Kocsis, B., Haiman, Z., Menou,
1952:   K., \& Frei, Z.\ 2007b, \prd, 76, 022003 (K07)
1953: 
1954: \bibitem[Lang \& Hughes(2006)]{lh06} Lang, R.\ N., \& Hughes, S.\ A.\
1955:   2006, \prd, 74, 122001 (Paper I)
1956: 
1957: \bibitem[Larson et al.(2000)]{lhh00} Larson, S.\ L.,
1958:   Hiscock, W.\ A., \& Hellings, R.\ W.\ 2000, \prd, 62, 062001
1959: 
1960: \bibitem[Micic et al.(2007)]{mhsa07}
1961:   Micic, M., Holley-Bockelmann, K., Sigurdsson, S., \& Abel, T.\
1962:   2007, \mnras, 380, 1533
1963: 
1964: \bibitem[Milosavljevi\'c \& Phinney(2005)]{mp05} Milosavljevi\'c, M.,
1965:   \& Phinney, E.\ S.\ 2005, \apj, 622, L93
1966: 
1967: \bibitem[Nelemans et al.(2001)]{nyz01}
1968:   Nelemans, G., Yungelson, L.\ R., \& Portegies Zwart, S.\ F.\ 2001,
1969:   \aap, 375, 890
1970: 
1971: \bibitem[Pan et al.(2008)]{panetal08} Pan, Y., Buonanno, A., Baker,
1972:   J.\ G., Centrella, J., Kelly, B.\ J., McWilliams, S.\ T., Pretorius,
1973:   F., \& van Meter, J.\ R.\ 2008, \prd, 77, 024014
1974: 
1975: \bibitem[Phillips(1993)]{p93} Phillips, M.\ M.\ 1993 \apj, 413, L105
1976: 
1977: \bibitem[Poisson \& Will(1995)]{pw95} Poisson, E., \& Will, C.\ M.\
1978:   1995, \prd, 52, 848
1979: 
1980: \bibitem[Riess et al.(1995)]{rpk95} Riess, A.\ G., Press,
1981:   W.\ H., \& Kirshner, R.\ P.\ 1995, \apj, 438, L17
1982: 
1983: \bibitem[Tyson et al.(2002)]{t02} Tyson, J.\ A., \& the {\it LSST}
1984:   Collaboration 2002, Proc. SPIE Int. Soc. Opt. Eng., 4836, 10
1985: 
1986: \bibitem[Vecchio(2004)]{v04} Vecchio, A.\ 2004, \prd, 70, 042001
1987: 
1988: \bibitem[Sesana et al.(2007)]{svh07} Sesana, A.,
1989:   Volonteri, M., \& Haardt, F.\ 2007, \mnras, 377, 1711
1990: 
1991: \bibitem[Trias \& Sintes(2008)]{ts08} Trias, M., \& Sintes, A.\ M.\
1992:   2008, \prd, 77, 024030
1993: 
1994: \bibitem[Wang et al.(2003)]{wgap03} Wang, L., Goldhaber, G., Aldering,
1995:   G., \& Perlmutter, S.\ 2003, \apj, 590, 944
1996: 
1997: \bibitem[Wang et al.(2002)]{whm02} Wang, Y., Holz, D.\ E.,
1998:   \& Munshi, D.\ 2002, \apjl, 572, L15
1999: 
2000: \bibitem[Will \& Wiseman(1996)]{ww96} Will, C.\ M., \& Wiseman, A.\
2001:   G.\ 1996, \prd, 54, 4813
2002: 
2003: \end{thebibliography}
2004: 
2005: \end{document}
2006: