1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: %% %%
3: %% Probing Anisotropies of Gravitational-Wave Backgrounds %%
4: %% With a Space-based Interferometer II: perturbative %%
5: %% reconstruction of a skymap in low-frequency regime %%
6: %% %%
7: %% %%
8: %% 2005/10/18 Version 2.0 re-submitted to Phys.Rev.D %%
9: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
10: \documentclass[showpacs,superscriptaddress,preprintnumbers,amsmath,amssymb,nofootinbib]{revtex4}
11:
12:
13: \usepackage{amsmath,amssymb,latexsym}
14: \usepackage{graphicx,bm}% Include figure files
15: \usepackage{dcolumn} % Align table columns on decimal point
16: \usepackage{epsf,psfig}
17: % \usepackage{epsf}
18: %
19: % Define a page header.
20: %
21: %=================================================================
22: %
23: \begin{document}
24: %
25: % =================================================================
26: %
27: %
28: %
29: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
30: \newcommand{\sinc}{ {\mathrm{sinc}} }
31: \newcommand{\hf}{ {\hat{f}} }
32: \newcommand{\Rdots}{\rotatebox[origin=c]{-45}{\vdots}}
33: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
34: %
35: %
36: %
37: %
38: %
39: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
40: \title{ Probing anisotropies of gravitational-wave backgrounds
41: \\
42: with a space-based interferometer II:
43: \\
44: \it{
45: perturbative reconstruction of a low-frequency skymap
46: } }
47: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
48:
49:
50: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
51: \author{Atsushi Taruya} \email{ataruya_at_utap.phys.s.u-tokyo.ac.jp}
52: \affiliation{ Research Center for the Early Universe~(RESCEU), School
53: of Science, The University of Tokyo, Tokyo 113-0033, Japan }
54:
55: \author{Hideaki Kudoh} \email{kudoh_at_utap.phys.s.u-tokyo.ac.jp}
56: \affiliation{ Department of Physics, The University of Tokyo, Tokyo
57: 113-0033, Japan }
58: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
59:
60:
61: \preprint{UTAP-531, RESCEU-34/05}
62: \pacs{04.30.-w, 04.80.Nn, 95.55.Ym, 95.30.Sf}
63: %%
64: %% PACS (Checked 2004/10/20)
65: %%
66: %% 04.30.-w, Gravitational waves: theory 04.80.Nn, Gravitational wave
67: %% detectors and experiments 95.55.Ym Gravitational radiation
68: %% detectors; 95.75.-z Observation and data reduction techniques;
69: %% 95.30.Sf Relativity and gravitation
70: %%
71: %%
72: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
73: \begin{abstract}
74: We present a perturbative reconstruction method to make a skymap of
75: gravitational-wave backgrounds (GWBs) observed via space-based
76: interferometer. In the presence of anisotropies in GWBs, the
77: cross-correlated signals of observed GWBs are inherently time-dependent
78: due to the non-stationarity of the gravitational-wave detector. Since
79: the cross-correlated signal is obtained through an all-sky integral
80: of primary signals convolving with the antenna pattern function of
81: gravitational-wave detectors, the non-stationarity of cross-correlated
82: signals, together with full knowledge of antenna pattern functions,
83: can be used to reconstruct an intensity map of the GWBs.
84: Here, we give two simple methods to reconstruct a skymap of GWBs
85: based on the perturbative expansion in low-frequency regime.
86: The first one is based on harmonic-Fourier representation of data
87: streams and the second is based on ``direct'' time-series data.
88: The latter method enables us to create a skymap in a direct manner.
89: The reconstruction technique is demonstrated for the case of the
90: Galactic gravitational wave background observed via planned space
91: interferometer, LISA. Although the angular resolution of
92: low-frequency skymap is rather restricted, the methodology presented
93: here would be helpful in discriminating the GWBs of galactic origins
94: by those of the extragalactic and/or cosmological origins.
95: \end{abstract}
96: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
97:
98:
99: \maketitle
100:
101:
102: % ==============================================================%
103: % ==============================================================%
104: \section{Introduction}
105: \label{sec:intro}
106: % ==============================================================%
107: % ==============================================================%
108:
109:
110: The gravitational-wave background, incoherent superposition of
111: gravitational waves coming from many unresolved point-sources and/or
112: diffuse-sources would be a cosmological gold mine to probe the dark side
113: of the Universe. While these signals are generally random and act as
114: {\it confusion noises} with respect to a periodic gravitational-wave
115: signal, the statistical properties of gravitational-wave backgrounds
116: (GWBs) carry valuable information such as physical processes of
117: gravitational-wave emission, source distribution and
118: populations. Moreover, the extremely early universe beyond the
119: last-scattering surface of cosmic microwave background (CMB) can be
120: directly explored using GWB. Therefore, GWBs may be regarded as an
121: ultimate cosmological tool alternative to the CMB.
122:
123:
124: Currently, several grand-based gravitational-wave detectors are now
125: under scientific operation, and the search for gravitational waves
126: enters a new era. There are several kinds of target GW sources in these
127: detectors, such as, coalescence of neutron star binaries and
128: core-collapsed supernova. On the other hand, planned space-based
129: detector, LISA (Laser Interferometer Space Antenna) and the
130: next-generation detectors, e.g., DECIGO \cite{Seto:2001qf} and
131: BBO \cite{BBO:2003} aim at detecting gravitational waves in
132: low-frequency band $0.1$mHz --$0.1$Hz, in which many detectable
133: candidates for GWBs exist. Among them, the GWB originated from the
134: inflationary epoch may be detected directly in this band
135: (e.g., \cite{Ungarelli:2000jp,Smith:2005mm}).
136: Therefore, for future application to cosmology, various implications for
137: these GWBs should deserve consideration both from the theoretical and
138: the observational viewpoint.
139:
140:
141: One fundamental issue to access a new subject of cosmology is to make a
142: skymap of GWB as a first step. Similar to the case of the CMB, the
143: intensity map of the GWBs is of particular interest and it provides
144: valuable information, which plays a key role to clarify the origin of
145: GWBs. Importantly, the low-frequency GWBs observed by LISA are expected
146: to be anisotropic due to the contribution of Galactic binaries to the
147: confusion noise (\cite{Hils:1990hg,Bender:1997bc,Nelemans:2001hp}, see
148: also recent numerical simulations
149: \cite{Benacquista:2003th,Edlund:2005ye,Timpano:2005gm}).
150: Thus, the GWB skymap is potentially useful to discriminate the
151: individual backgrounds from many superposed GWBs, as well as to identify
152: the underlying physical processes. The basic idea to create a skymap of
153: GWB is to use the directional information obtained through the
154: time-modulation of the correlation signals, which is caused by the
155: motion of gravitational-wave detector. As detector's antenna pattern
156: sweeps out the sky, the amplitude of the gravitational-wave signal would
157: gradually change in time if the intensity distribution of GWB is
158: anisotropic. Using this, a method to explore anisotropies of GWB has
159: been proposed \cite{Giampieri:1997,Giampieri:1997ie,Allen:1997gp}. Later,
160: the methodology was applied to study the anisotropic GWB observed by
161: space interferometer, LISA
162: \cite{Ungarelli:2001xu,Cornish:2001hg,Cornish:2002bh,Seto:2004ji,Seto:2004np}.
163:
164:
165:
166: In the previous paper \cite{Kudoh:2004he}, which is referred to as
167: {\it Paper I} in the present paper, we have investigated the directional
168: sensitivity of space interferometer to the anisotropy of GWB.
169: Particularly focusing on the geometric properties of antenna pattern
170: functions and their angular power, we found that the angular sensitivity
171: to the anisotropic GWBs is severely restricted by the data combination
172: and the symmetry of detector configuration. As a result, in the case of
173: the single LISA detector, detectable multipole moments are limited to
174: $\ell\lesssim 8$--$10$ with the effective strain sensitivity
175: $h\sim 10^{-20}$ Hz$^{-1/2}$.
176: This is marked contrast to the angular sensitivity to the chirp signals
177: emitted from point sources, in which the angular resolution can reach at
178: a level of a square degree or even better than that
179: \cite{Cutler:1997ta,Moore:1999zw,Peterseim:1997ic,Takahashi:2003wm}.
180:
181:
182:
183: Despite the poor resolution of LISA detector with respect to GWBs,
184: making a skymap of GWBs is still an important general issue and thus
185: needs to be investigated. In the present paper, we continue to
186: investigate the map-making problem and consider how the intensity map of the
187: GWBs is reconstructed from the time-modulation signals observed via
188: space interferometer. In particular, we are interested in the
189: low-frequency GWB, the wavelength of which is longer than the arm-length
190: of the detector. In such case, the frequency dependence of the detector
191: response becomes simpler and a perturbative scheme based on the
192: low-frequency expansion of the antenna pattern functions can be
193: applied. Owing to the least-squares approximation, we present a robust
194: reconstruction method. The methodology is quite general and is also
195: applicable to the map-making problem in the case of the ground detectors.
196: We demonstrate how the present reconstruction method works well in a
197: specific example of GWB source, i.e., Galactic confusion-noise background.
198: With a sufficient high signal-to-noise ratio for each anisotropic
199: components of GWB, we show that the space interferometer
200: LISA can create the low-frequency skymap of Galactic GWB with angular
201: resolution $\ell\leq5$.
202: Since the resultant skymap is obtained in a non-parametric way without
203: any assumption of source distributions, we hope that despite the poor
204: angular resolution, the present methodology will be helpful to give a
205: tight constraint on the luminosity distribution of GWBs if we combine
206: it with the other observational techniques.
207:
208:
209: The organization of the paper is as follows. In Sec.
210: \ref{sec:detection of anisotropy}, a brief discussion on the detection
211: and the signal processing of anisotropic GWBs is presented, together
212: with the detector response of space interferometer. Sec.\ref{sec:method}
213: describes the details of the reconstruction of a GWB skymap in
214: low-frequency regime. Owing to the least-squares approximation and the
215: low-frequency expansion, a perturbative reconstruction scheme is developed.
216: Related to this (in the case of LISA), and we give an important remark
217: on the degeneracy between some multipole moments (Appendix
218: \ref{appendix:on_the_degeneracy}). In Sec.\ref{sec:demonstration}, the
219: reconstruction method is demonstrated in the case of the Galactic GWB.
220: The signal-to-noise ratios for anisotropy of GWB are evaluated and the
221: feasibility to make a skymap of GWB is discussed. Finally,
222: Section \ref{sec:conclusion} is devoted to a summary and conclusion.
223:
224:
225:
226: % ==============================================================%
227: % ==============================================================%
228: \section{Basic formalism}
229: \label{sec:detection of anisotropy}
230: % ==============================================================%
231: % ==============================================================%
232:
233:
234: % ///////////////////////////////////////////////////////////// %
235: \subsection{Correlation analysis}
236: \label{subsec:correlation}
237: % ///////////////////////////////////////////////////////////// %
238:
239:
240: Let us begin with briefly reviewing the signal processing of
241: gravitational-wave backgrounds based on the correlation analysis
242: \cite{Kudoh:2004he}. Stochastic gravitational-wave backgrounds are
243: described by incoherent superposition of plane gravitational waves
244: ${\mathbf h}=h_{ij}$ given by
245: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
246: \begin{eqnarray}
247: {\mathbf h}(t,\mathbf{x})= \sum_{A=+,\times} \int_{-\infty}^{+\infty} df
248: \int d\mathbf{\Omega}\,\,e^{i\,\,2\pi\,f(t-\mathbf{\Omega}\cdot \mathbf{x})}
249: \,\tilde{h}_A(f,\mathbf{\Omega})\,\mathbf{e}^A(\mathbf{\Omega}).
250: \end{eqnarray}
251: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
252: The Fourier amplitude $\tilde{h}_A (f,\mathbf{\Omega})$ of the
253: gravitational waves for the two polarization modes
254: ${\mathbf e}^A$ ($A=+,\times$) is assumed to be characterized by a
255: stationary
256: random process with zero mean $\big\langle \widetilde{h}_A \big\rangle =0 $.
257: The power spectral density $S_h$ is then defined by
258: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
259: \begin{eqnarray}
260: \left\langle
261: \widetilde{h}_A^* (f, {\bf{\Omega}})
262: {{\widetilde{h}}_{A'} } (f', {\bf{\Omega}}')
263: \right\rangle
264: &=&
265: \frac{1}{2} \delta(f-f')
266: \frac{ \delta^2 ( {\bf{\Omega}} - {\bf{\Omega}}')}{4\pi}
267: \delta_{AA'} S_h(|f|, \,\mathbf{\Omega}),
268: \label{eq:h*h gaussian process}
269: \end{eqnarray}
270: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
271: where ${\mathbf{\Omega}}$ is the direction of a propagating plane wave.
272: Note that the statistical independence between two different directions
273: in the sky is implicitly assumed in equation (\ref{eq:h*h gaussian
274: process}), which might not be generally correct. Actually, CMB skymap
275: exhibits a large-angle correlation between the different skies, which
276: reflects the primordial density fluctuations. For the GWB of our
277: interest, the wavelength of the tensor fluctuations detected via space
278: interferometers is much shorter than the cosmological scales and thereby
279: the assumption put in equation (\ref{eq:h*h gaussian process})
280: is practically valid.
281:
282:
283: The detection of a gravitational-wave background is achieved through the
284: correlation analysis of two data streams. The planned space
285: interferometer, LISA and also the next generation detectors DECIGO/BBO
286: constitute several spacecrafts, each of which exchanges laser beams with
287: the others. Combining these laser pulses, it is possible to synthesize
288: the various output streams which are sensitive (or insensitive) to the
289: gravitational-wave signal. The output stream for a specific combination
290: $I$ denoted by $s_I(t)$ is generally described by a sum of the
291: gravitational-wave signal $h_I(t)$ and the instrumental noise $n_I(t)$ by
292: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
293: \begin{equation*}
294: s_I(t) = h_I(t)+ n_I(t).
295: \end{equation*}
296: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
297: We assume that the noise $n_I(t)$ is treated as a Gaussian random
298: process with spectral density $S_{\rm n}(f)$ and zero mean
299: $\big\langle n_I\big\rangle=0 $. The gravitational-wave signal $h_I$ is
300: obtained by contracting $\mathbf{h}$ with detector's response function
301: and/or detector tensor ${\mathbf D}:=D_{ij}$:
302: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
303: \begin{equation}
304: h_I(t) = \sum_{A=+,\times} \int_{-\infty}^{+\infty} df
305: \int d \mathbf{\Omega} ~
306: e^{i2\pi f (t-\mathbf{\Omega} \cdot \mathbf{x}_I) }
307: {\textbf{D}}_I ( {\mathbf{\Omega}} ,f;t) \,
308: {\textbf : }\, \mathbf{e}^{A}(\mathbf{\Omega})\,\,
309: \widetilde{h}_A(f,\mathbf{\Omega}).
310: \label{eq:h_I(t)}
311: \end{equation}
312: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
313: Note that the response function explicitly depends on time.
314: The time variation of response function is caused by the orbital motion
315: of the detector and it plays a key role in reconstructing a skymap of the GWBs.
316: Typically, the orbital frequency of the detector motion is much lower
317: than the observed frequency and within the case, the expression
318: (\ref{eq:h_I(t)}) is validated.
319:
320:
321:
322:
323: Provided the two output data sets, the correlation analysis is examined
324: depending on the strategy of data analysis, i.e., self-correlation
325: analysis using the single data stream or cross-correlation analysis
326: using the two independent data stream.
327: Defining $S_{IJ}(t) \equiv \left\langle s_I(t)s_J(t) \right\rangle$, the Fourier
328: counterpart of it $\widetilde{S}_{IJ}(t,f)$, which is related with
329: $S_{IJ}(t)$ by $S_{IJ}(t)=\int df \widetilde{S}_{IJ}(t,f)$, becomes
330: \footnote{
331: In Paper I, we had used $C_{IJ}$ to denote the output data
332: $\left\langle s_I(t)s_J(t) \right\rangle$ itself. This coincides with
333: the present definition if we neglect the instrumental noises.
334: }
335: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
336: \begin{eqnarray}
337: \widetilde{S}_{IJ}(t,f) & = &
338: \widetilde{C}_{IJ}(t,f) + \delta_{IJ} \,\,S_n(|f|),
339: \label{eq:detector output S(t)}
340: \end{eqnarray}
341: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
342: where we define
343: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
344: \begin{eqnarray}
345: \widetilde{C}_{IJ}(t,f)=\int \frac{d
346: \mathbf{\Omega}}{4\pi} S_h(|f|, \mathbf{\Omega})~
347: \mathcal{ F}_{IJ}^E(f, \mathbf{\Omega};\,t).
348: \label{eq:def_of_C(f)}
349: \end{eqnarray}
350: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
351: Here, the function $\mathcal{F}^E_{IJ}$ is the so-called antenna pattern
352: function defined in an ecliptic coordinate, which is expressed in terms
353: of detector's response function:
354: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
355: \begin{eqnarray}
356: && \mathcal{F}^E_{IJ}(f, \mathbf{\Omega};\,t)=
357: e^{ i \, 2\pi f \,{\bf \Omega \cdot}(\textbf{x}_I -\textbf{x}_J) }
358: \sum_{A=+,\times}
359: F_I^{A*}( \mathbf{\Omega},f;\,t) F_J^A( \mathbf{\Omega},f;\,t)
360: \cr
361: && F_I^A( \mathbf{\Omega},f;\,t) =
362: {\bf D}_I( \mathbf{\Omega},f;\,t)\, \textbf{:} \,
363: \textbf{e} ^A (\mathbf{\Omega})
364: \label{eq:def_of_antenna}
365: \end{eqnarray}
366: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
367:
368:
369: Apart from the second term, equation (\ref{eq:detector output S(t)})
370: implies that the luminosity distribution of GWBs
371: $S_h(f,\mathbf{\Omega})$ can be obtained by deconvolving the all-sky
372: integral of antenna pattern function from the time-series data $S_{IJ}(t)$.
373: To see this more explicitly, we focus on equation (\ref{eq:def_of_C(f)})
374: and decompose the antenna pattern function and the luminosity
375: distribution into spherical harmonics in an ecliptic coordinate, i.e.,
376: sky-fixed frame:
377: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
378: \begin{eqnarray}
379: S_h(|f|,\, \mathbf{\Omega})
380: = \sum_{\ell, m} \,\,[p_{\ell m} (f)]^* \,\,
381: Y_{\ell m}^* ( \mathbf{\Omega}),
382: \cr
383: \mathcal{F}^E_{IJ}(f,\, \mathbf{\Omega};\,t)
384: =
385: \sum_{\ell, m} \,\, a_{\ell m}^E (f,t) \,\,
386: Y_{\ell m} ( \mathbf{\Omega}).
387: \label{eq:Y_lm expansion of S and F}
388: \end{eqnarray}
389: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
390: Note that the properties of spherical harmonics yield
391: $p_{\ell m}^*=(-1)^m p_{\ell,-m}$ and
392: $a_{\ell m}^*=(-1)^{\ell-m}a_{\ell,-m}$, where the latter property comes from
393: ${\mathcal{F}}_{IJ}^* (f, \Omega;t) = {\mathcal{F}}_{IJ} (f, - \Omega;t)$
394: \cite{Kudoh:2004he}. Substituting (\ref{eq:Y_lm expansion of S and F}) into
395: (\ref{eq:def_of_C(f)}) becomes
396: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
397: \begin{eqnarray}
398: \widetilde{C}_{IJ}(t,f) = \frac{1}{4\pi}\, \sum_{\ell m}\,\,
399: \left[ p_{\ell m}(f) \right]^* a^{E }_{\ell m}(f,\,t),
400: \label{eq:C_IJ in E-frame}
401: \end{eqnarray}
402: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
403: where we have dropped the contribution from the detector noise.
404: The expression (\ref{eq:C_IJ in E-frame}) still involves the
405: time-dependence of antenna pattern functions.
406: To eliminate this, one may further rewrite equation
407: (\ref{eq:C_IJ in E-frame}) by employing the harmonic expansion in
408: detector's rest frame.
409: We denote the multipole coefficients of the antenna pattern in
410: detector's rest frame by $a_{\ell m}$.
411: The transformation between the detector rest frame and the sky-fixed
412: frame is described by a rotation matrix by the Euler angles
413: $(\psi, \vartheta ,\varphi)$, whose explicit relation is expressed in
414: terms of the Wigner $D$ matrices \footnote{
415: \label{footnote:orbital motion}
416: We are specifically concerned with the time dependence of directional
417: sensitivity of antenna pattern functions, not the real orbital motion.
418: Hence, only the time evolution of directional dependence is considered
419: in the expression (\ref{eq:Wigner_formula}).}
420: \cite{Allen:1997gp,Cornish:2002bh, Edmonds:1957}:
421: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
422: \begin{eqnarray}
423: a^E_{\ell m }(f,t) = \sum_{n=-\ell}^\ell
424: e^{-i \,n \,\psi}\, d^\ell_{nm}( \vartheta)\, e^{-i \,m \,\varphi}
425: \,\,a_{\ell n}(f).
426: \label{eq:Wigner_formula}
427: \end{eqnarray}
428: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
429: Here the Euler rotation is defined to perform a sequence of rotation,
430: starting with a rotation by $\psi$ about the original $z$ axis, followed
431: by rotation by $\vartheta$ about the original $y$ axis, and ending with
432: a rotation by $\varphi$ about the original $z$ axis.
433: The Wigner $D$ matrices for $n\ge m$ is
434: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
435: \begin{eqnarray}
436: d_{n m}^\ell ({\mathcal{\theta}}) =
437: (-1)^{\ell-n}
438: \sqrt{\frac{ (\ell+n)! (\ell-n)! }{ (\ell+m)!(\ell-m)! } }
439: \left( \cos \frac{\theta}{2}\right)^{n+m}
440: \left(-\sin \frac{\theta}{2}\right)^{n-m}
441: P_{\ell-n}^{(n+m, n-m)}(-\cos\theta)
442: \end{eqnarray}
443: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
444: with $P_n^{(a,b)}$ being Jacobi polynomial. For
445: $n<m$, we have $d^{\ell}_{nm}=(-1)^{n-m} d^\ell_{mn}$.
446:
447:
448:
449: %%%%%%%%%%%%%%%%%%%%%% Figure %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
450: \begin{figure}[t]
451: \begin{center}
452: \includegraphics[width=8cm,angle=0,clip]{figs/ecliptic_coordinates.eps}
453: \end{center}
454: \caption{
455: LISA configuration in the ecliptic coordinates. The Galactic center is at
456: RA $267.4^\circ$ and dec. $-5.5^{\circ}$ in this coordinate system.
457: }
458: \label{fig:LISA}
459: \end{figure}
460: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
461:
462:
463: Let us now focus on the orbital motion of the LISA constellation (see
464: Fig.\ref{fig:LISA}).
465: The LISA orbital motion can be expressed by
466: $\psi=-\omega t$, $\vartheta=-\pi/3$, $\varphi= \omega t$, where
467: $\omega = 2\pi/T_0$ is the orbital frequency of LISA ($T_0=1$ sidereal year)
468: \footnote{
469: The relation $\psi=-\varphi$ does not necessarily hold for orbital motion of
470: LISA and there may be some possibilities to impose a constant phase
471: difference, i.e., $\psi=-\varphi + c$. However, for the sake of
472: simplicity, we put $c=0$.
473: }
474: Since the antenna pattern function is periodic in time due to the
475: orbital motion, the expected signals also vary in time with the same
476: period $T_0$. It is therefore convenient to express the output signals by
477: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
478: \begin{equation}
479: \widetilde{C}_{IJ}(t,f)=\sum_{k=-\infty}^{+\infty}\,\,
480: \widetilde{C}_{IJ,k}(f)\,\,e^{i\,k\,\omega\,t}.
481: \end{equation}
482: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
483: Using the relation (\ref{eq:Wigner_formula}) with specific parameter
484: set, the Fourier component $\widetilde{C}_{IJ,k}(f)$ is then given by
485: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
486: \begin{eqnarray}
487: \tilde{C}_k(f) &\equiv& \frac{1}{T_0}
488: \int^{T_0}_{0} dt~ e^{-ik \omega t}\,\, \tilde{C}_{IJ}(t,f)
489: \nonumber\\
490: &=& \frac{1}{4\pi}
491: \sum_{\ell=0}^{\infty} \sum_{m=- \ell}^{\ell-k}~ [p_{\ell m}(f)]^*\,\,
492: d^{\ell}_{(m+k),m} \left( -\frac{\pi}{3} \right) ~a_{\ell,(m+k)}(f)
493: \label{eq:deconvolution}
494: \end{eqnarray}
495: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
496: for $k\ge 0$. As for $k<0$, the lower and the upper limit of the sum
497: over $m$ are changed to $m=-\ell-k$ and $m=\ell$, respectively.
498:
499:
500: Equation (\ref{eq:deconvolution}) as well as (\ref{eq:def_of_C(f)}) is
501: the theoretical basis to reconstruct the skymap of GWBs.
502: Given the output data $\tilde{C}_k(f)$ (or $\tilde{C}_{IJ}(t,f)$)
503: experimentally, the task is to solve the linear system
504: (\ref{eq:deconvolution}) with respect to $p_{\ell m}(f)$ for given
505: antenna pattern functions. One important remark deduced from equation
506: (\ref{eq:deconvolution}) is that the accessible multipole coefficients
507: $p_{\ell m}$ are severely restricted by the angular sensitivity of
508: antenna pattern functions. The important properties of the antenna
509: pattern functions are summarized in next subsection.
510: Another important message is that the above linear systems are generally
511: either over-constrained or under-determined. For the expression
512: (\ref{eq:deconvolution}), if one truncates the multipole expansion with
513: $\ell=\ell_\mathrm{max}$, the system consists of
514: $\frac{1}{2}(\ell_{max}+1)(2\ell_{max}+1)$ unknowns and
515: $N(2\ell_{max}+1)$ equations (for $k \ge0$), where $N$ is the number of
516: available modes of the antenna pattern functions. Thus this
517: deconvolution problem is, in principle, over-determined for a relatively
518: small truncation multipole $\ell_{\text{max}}$, while it becomes
519: under-determined for a larger value of $\ell_{\text{max}}$. We will
520: later discuss the over-determined case in
521: Sec.\ref{subsubsec:over-determined}, where
522: $\ell_{\text{max}} = 5$ and $N \simeq 5$.
523: %
524: %
525: %
526: %
527: %
528: %
529: %
530: % ///////////////////////////////////////////////////////////// %
531: \subsection{Detector response and antenna pattern functions}
532: \label{subsec:antenna_pattern}
533: % ///////////////////////////////////////////////////////////// %
534: %
535: %
536: %
537: %
538: %
539: %%%%%%%%%%%%%%%%%%%%%%%%%%% TABLE %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
540: \begin{table}[b]
541: \caption{\label{tab:summay_antenna}
542: Important properties of antenna pattern function.
543: Note that the cross-correlated data are blind to isotropic GWBs.
544: }
545: \begin{ruledtabular}
546: \begin{tabular}{cll}
547: Combination of variables & Visible multipole moments in low-frequency
548: regime $(\hat{f}\lesssim1)$ & General properties \\
549: \hline
550: (A,A), (E,E) & $\mathcal{O}(\hat{f}^2):$ $\ell=0,\,\,2,\,\,4$
551: & visible only to $\ell=$even \\
552: (T,T) & $\mathcal{O}(\hat{f}^4):$ $\ell=0,\,\,4,\,\,6$
553: & visible only to $\ell=$even \\
554: (A,E) & $\mathcal{O}(\hat{f}^2):$ $\ell=4$,\quad \quad
555: $\mathcal{O}(\hat{f}^3):$ $\ell=3,\,\,5$
556: & blind to $\ell=0,\,\,1$ \\
557: (A,T), (E,T) & $\mathcal{O}(\hat{f}^3):$ $\ell=1,\,\,3,\,\,5$
558: & blind to $\ell=0$ \\
559: \end{tabular}
560: \end{ruledtabular}
561: \end{table}
562: %%%%%%%%%%%%%%%%%%%%%%%%%%% TABLE %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
563:
564:
565: The output signals of space interferometer sensitive to the
566: gravitational waves are constructed by time-delayed combination of laser
567: pulses. In the case of LISA, the technique to synthesize data streams
568: canceling the laser frequency noise is known as time-delay
569: interferometry (TDI), which is crucial for our subsequent analysis.
570: In the present paper, we use the optimal set of TDI variables
571: independently found by Prince et al. \cite{Prince:2002hp} and
572: Nayak et al.\cite{Nayak:2002ir}, which are free from the noise correlation
573: \footnote{
574: The optimal TDI variables adopted here may not be the best TDI
575: combinations for the present purpose. There might be a better choice of
576: the signal combinations, although a dramatic improvement of the angular
577: sensitivity would not be expected. }.
578: A simple realization of such data set is obtained from a combination of
579: Sagnac signals, which are the six-link observables using all six LISA
580: oriented arms. For example, the Sagnac signal S$_1$ measures the phase
581: difference accumulated by two laser beams received at space craft $1$,
582: each of which travels around the LISA array in clockwise or
583: counter-clockwise direction.
584: The explicit expression of detector tensor for such signal, which we denote by
585: ${\bf D}_{\scriptscriptstyle\rm S_1}$, is given in equation (21) of
586: Paper I (see also \cite{Cornish:2001bb,Armstrong:1999}).
587: The analytic expression for other detector tensors ${\bf D}_{\rm S_2}$
588: and ${\bf D}_{\rm S_3}$ are also obtained by the cyclic permutation of
589: the unit vectors $\mathbf{a},\,\mathbf{b}$ and $\mathbf{c}$.
590:
591:
592: Combining the three Sagnac signals, a set of optimal data combinations
593: can be constructed
594: \cite{Prince:2002hp,Nayak:2002ir} (see also \cite{Krolak:2004xp}):
595: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
596: \begin{eqnarray}
597: && {\bf D}_{\rm A} =
598: \frac{1}{\sqrt{2} }( {\bf D}_{\rm S_3}- {\bf D}_{\rm S_1} ),
599: \nonumber
600: \\
601: && {\bf D}_{\rm E} =
602: \frac{1}{\sqrt{6}}(
603: {\bf D}_{\rm S_1} - 2{\bf D}_{ \rm S_2 } + {\bf D}_{ \rm S_3} ),
604: \nonumber
605: \\
606: && {\bf D}_{\rm T} =
607: \frac{1}{\sqrt{3}}( {\bf D}_{\rm S_1}
608: + {\bf D}_{\rm S_2} + {\bf D}_{\rm S_3} ).
609: \label{eq:def AET mode}
610: \end{eqnarray}
611: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
612: These three detector tensors are referred to as $A,E,T$-variables,
613: which generate six kinds of antenna pattern functions.
614: Notice that in the equal arm-length limit, frequency dependence of these
615: functions is simply expressed in terms of the normalized frequency:
616: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
617: \begin{equation}
618: \hat{f}\equiv\frac{f}{f_*},
619: \label{eq:normalized_f}
620: \end{equation}
621: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
622: where the characteristic frequency is given by
623: $f_*=c/(2\pi L)$ with $L$ being the arm-length of detector.
624: With the arm-length $L=5\times10^6$ km, the characteristic frequency of
625: LISA becomes $f_*\simeq 9.54$ mHz.
626: In what follows, we use the analytic expressions for equal arm-length
627: limit to demonstrate the reconstruction of GWB skymap.
628: This is sufficient for the present purpose, because we are concerned
629: with a fundamental theoretical basis to map-making capability of GWBs.
630: The idea provided in the present paper would allow us to examine a more
631: realistic situation and the same strategy can be applied to an extended
632: analysis in the same manner.
633:
634:
635:
636: Finally, we note that the antenna pattern functions for the optimal
637: combinations of TDI have several distinctive features in angular sensitivity.
638: In the low frequency limit ($\hat{f} \ll 1$), the antenna pattern
639: functions for the self-correlations and the cross-correlations are expanded as
640: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
641: \begin{equation}
642: \begin{aligned}
643: {\mathcal F}_{AA} (f,\,\mathbf{\Omega}) =~
644: & {\mathcal F}_{AA}^{(2)}(\mathbf{\Omega})\, \hat{f}^2 &
645: &&
646: +~ &~{\mathcal F}_{AA}^{(4)}(\mathbf{\Omega})\, \hat{f}^4 &
647: + ~O(\hat{f}^6)
648: \\
649: %\nonumber
650: {\mathcal F}_{TT} (f,\,\mathbf{\Omega}) =~
651: & &
652: & &
653: &~{\mathcal F}_{TT}^{(4)}(\mathbf{\Omega})\, \hat{f}^4 &
654: +~ O(\hat{f}^6)
655: \\
656: %\nonumber
657: {\mathcal F}_{AE} (f,\,\mathbf{\Omega}) = ~
658: &{\mathcal F}_{AE}^{(2)}(\mathbf{\Omega})\, \hat{f}^2&
659: +~ &~{\mathcal F}_{AE}^{(3)}(\mathbf{\Omega})\, \hat{f}^3&
660: +~ &~{\mathcal F}_{AE}^{(4)}(\mathbf{\Omega})\, \hat{f}^4&
661: +~ O(f^5)
662: \\
663: {\mathcal F}_{AT} (f,\,\mathbf{\Omega}) =~
664: & &
665: &~{\mathcal F}_{AT}^{(3)}(\mathbf{\Omega})\, \hat{f}^3 &
666: +~ &~{\mathcal F}_{AT}^{(4)}(\mathbf{\Omega})\, \hat{f}^4&
667: + ~O(\hat{f}^5)
668: \end{aligned}
669: \label{eq:expand_antenna}
670: \end{equation}
671: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
672: The frequency dependence of ${\mathcal F}_{EE}$ and ${\mathcal F}_{ET}$
673: are the same as for ${\mathcal F}_{AA}$ and ${\mathcal F}_{AT}$, respectively.
674: Since the leading term of the TT-correlation is $O(\hat{f}^4)$, the
675: $TT$-correlation becomes insensitive to the gravitational waves in the
676: low-frequency regime. We will use all correlated data set except for the
677: $TT$-correlation. In Table \ref{tab:summay_antenna}, we summarize the
678: important properties for the multipole moments of antenna pattern functions.
679: Also, in Appendix \ref{appendix:Mulipole coefficients}, employing the
680: perturbative approach based on the low-frequency approximation
681: $\hat{f}\ll1$, the spherical harmonic expansion for antenna pattern
682: functions are analytically computed, which will be useful in subsequent analysis.
683: %
684: %
685: %
686: %
687: %
688: %
689: %
690: %
691: %
692: %
693: % ==============================================================%
694: % ==============================================================%
695: \section{Perturbative reconstruction method
696: for GWB skymap}
697: \label{sec:method}
698: % ==============================================================%
699: % ==============================================================%
700: %
701: %
702: %
703: %
704: %
705: %
706: %
707: %
708: %
709: We are in position to discuss the methodology to reconstruct a skymap of
710: GWB based on the expression (\ref{eq:deconvolution}) (or
711: (\ref{eq:def_of_C(f)})).
712: Since we are specifically concerned with low-frequency sources observed
713: via LISA, it would be helpful to employ a perturbative approach using
714: the low-frequency expansion of antenna pattern function.
715: In Sec.\ref{subsec:general_scheme}, owing to the expression
716: (\ref{eq:deconvolution}), a perturbative reconstruction method is presented.
717: Sec.\ref{subsec:time-domain} discusses alternative reconstruction method
718: based on the time-series representation (\ref{eq:def_of_C(f)}).
719:
720:
721:
722: % ///////////////////////////////////////////////////////////// %
723: \subsection{General scheme}
724: \label{subsec:general_scheme}
725: % ///////////////////////////////////////////////////////////// %
726:
727:
728:
729: Hereafter, we focus on the GWBs in the low-frequency band of the
730: detector, the wavelength of which is typically longer than the
731: arm-length of the gravitational detector, i.e., $\hat{f}\lesssim 1$.
732: Without loss of generality, we restrict our attention to the
733: reconstruction of a GWB skymap in a certain narrow frequency range,
734: $f\sim f+ \Delta f$, within which a separable form of the GWB spectrum
735: is a good assumption:
736: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
737: \begin{equation}
738: S_h(f,\mathbf{\Omega}) = H(f)\,\,P(\mathbf{\Omega}).
739: \label{eq:S_h_separable}
740: \end{equation}
741: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
742: Further, for our interest of the narrow bandwidth, it is reasonable to
743: assume that the spectral density $H(f)$ is described by a power-law
744: form as $H(f)= {\mathcal N} \, f^{\alpha}$.
745: In the reconstruction analysis discussed below, the spectral index
746: $\alpha$ is assumed to be determined beforehand from theoretical
747: prediction and/or experimental constraint
748: \footnote{In our general scheme, we do not assume a priori the
749: normalization factor ${\mathcal N}$ in the function $H(f)$, which
750: should be simultaneously determined with the reconstruction of an
751: intensity distribution $P(\mathbf{\Omega})$.
752: In the low-frequency approximation, $\hat{f}\ll1$, however, there
753: exists a degeneracy between the monopole and the quadrupole
754: components and one cannot correctly determine the normalization
755: ${\mathcal N}$. (See Appendix.\ref{appendix:on_the_degeneracy}.) }.
756:
757:
758:
759: In the low frequency approximation up to the order
760: $\mathcal{O}(\hat{f}^3)$, we have five output signals which
761: respond to the GWBs, i.e., $AA$-, $EE$-, $AE$-, $AT$- and
762: $ET$-correlations, each output of which is represented by equation
763: (\ref{eq:deconvolution}).
764: Collecting these linear equations and arraying them appropriately,
765: the linear algebraic equations can be symbolically written in a matrix form as
766: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
767: \begin{eqnarray}
768: {\mathbf {c}}(f) = {\mathbf {A}}(f) \cdot {\mathbf {p}}.
769: \label{eq: c=Ap}
770: \end{eqnarray}
771: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
772: In the above expression, while the vector ${\mathbf {p}}$ represents
773: the unknowns consisting of the multipole coefficients of GWBs
774: $p_{\ell m}$, the vector ${\mathbf {c}}(f)$ contains the correlation signals
775: $\tilde{C}_{k}(f)$.
776: Here, the frequency dependence of the function $H(f)$ has been already
777: factorized and thereby the vector ${\mathbf {p}}$ only contains the
778: information about the angular distribution.
779: The matrix ${\mathbf {A}}(f)$ is the known quantity consisting of the
780: multipole coefficients of each antenna pattern and the Wigner $D$ matrices
781: (see Appendix \ref{appendix:SVD} for an explicit example).
782:
783:
784: As we have already mentioned, the linear systems (\ref{eq: c=Ap})
785: become either over-determined or under-determined system.
786: In most of our treatment in the low-frequency regime, the linear systems
787: (\ref{eq: c=Ap}) tend to become over-determined, but this is not always
788: correct depending on the amplitude of GWB spectrum
789: $H(f)$ (see Sec.\ref{subsec:on_snr}).
790: In any case, the matrix $\mathbf{A}$ would not be a square matrix and
791: the number of components of the vector $\mathbf{c}$ does not coincide
792: with the one in the vector $\mathbf{p}$.
793: In the over-determined case, while there is a hope to get a unique solution
794: $\mathbf{p}$ to produce the gravitational-wave signal
795: $\mathbf{c}$, it seems practically difficult due to the errors
796: associated with the instrumental noise and/or numerical analysis.
797: Hence, instead of pursuit of a rigorous solution, it would be better to
798: focus on the issue how to get an approximate solution by a simple and
799: systematic method.
800:
801:
802: In the case of our linear system, the approximate solution for the
803: multipole coefficient $\mathbf{p}_{\rm approx}$ can be obtained from
804: the least-squares method in the following form:
805: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
806: \begin{equation}
807: \mathbf{p}_{\rm\scriptscriptstyle approx}
808: = \mathbf{A}^+ \cdot \mathbf{c}.
809: \label{eq:p=AC}
810: \end{equation}
811: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
812: The matrix $\mathbf{A}^+$ is called the pseudo-inverse matrix of
813: Moore-Penrose type, whose explicit expression can be uniquely determined
814: from the singular value decomposition (SVD)
815: \cite{Press:NRC++}. According to the theorem of linear algebra, the
816: matrix $\mathbf{A}$, whose number of rows is greater than or equal to
817: its number of columns, can be generally written as
818: $\mathbf{A}= U^T\cdot \mbox{diag}[w_i]\cdot V $. Here, the matrices $U$ and
819: $V$ are orthonormal matrices which satisfy
820: $U^{\dagger}\,U=V^{\dagger}\,V=\mathbf{1}$,
821: where the quantity with subscript $^{\dagger}$ represents the Hermite
822: conjugate variable. The quantity $\mbox{diag}[w_i]$ represents an
823: diagonal matrix with singular values $w_i$ with respect to the matrix
824: $\mathbf{A}$. Then the pseudo-inverse matrix becomes
825: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
826: \begin{equation}
827: \mathbf{A}^+ = V^T\cdot \mbox{diag}[w_i^{-1}]\cdot U.
828: \label{eq:pseudo-inverse_A}
829: \end{equation}
830: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
831: It should be stressed that the explicit form of the pseudo-inverse matrix
832: $\mathbf{A}^+$ is characterized only by the angular dependence of
833: antenna pattern functions.
834:
835:
836: In principle, the least-squares method by SVD can work well and all the
837: accessible multipole moments of GWB would be obtained as long as the
838: antenna pattern functions have the corresponding sensitivity to each
839: detectable multipole moment. As we mentioned in
840: Sec.\ref{subsec:antenna_pattern}, however, the angular power of antenna
841: pattern function depends sensitively on the frequency. In the
842: low-frequency regime, the frequency dependence of the non-vanishing
843: multipole moments appears at
844: $\mathcal{O}(\hat{f}^2)$ for
845: $\ell=0,\,\,2$ and $4$, and $\mathcal{O}(\hat{f}^3)$ for
846: $\ell=1,\,\,3$ and $5$ (see Table \ref{tab:summay_antenna}).
847: In this respect, by a naive application of the least-squares method,
848: it is difficult to extract the information about $\ell=$odd modes of
849: GWBs because the singular values of the matrix $\mathbf{A}$ are
850: dominated by the lowest-order contribution of the antenna pattern functions.
851:
852:
853: For a practical and a reliable estimate of the odd multipoles in
854: low-frequency regime, the least-squares method should be applied
855: combining with the perturbative scheme described below. Let us recall
856: from the expression (\ref{eq:expand_antenna}) that the matrix
857: $\mathbf{A}$ can be expanded as
858: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
859: \begin{eqnarray}
860: \mathbf{A} = \hat{f}^2\,\,\mathbf{A}^{(2)} +
861: \hat{f}^4\,\,\mathbf{A}^{(4)} + \cdots
862: \label{eq:expand_mat_A1}
863: \end{eqnarray}
864: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
865: for a matrix consisting of the self-correlation signals
866: $\mathcal{F}_{AA}$ and $\mathcal{F}_{EE}$,
867: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
868: \begin{eqnarray}
869: \mathbf{A} = \hat{f}^2\,\,\mathbf{A}^{(2)} + \hat{f}^3\,\,\mathbf{A}^{(3)} +
870: \hat{f}^4\,\,\mathbf{A}^{(4)} + \cdots
871: \label{eq:expand_mat_A2}
872: \end{eqnarray}
873: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
874: for a matrix consisting of the cross-correlation signal
875: $\mathcal{F}_{AE}$, and
876: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
877: \begin{eqnarray}
878: \mathbf{A} = \hat{f}^3\,\,\mathbf{A}^{(3)} +
879: \hat{f}^4\,\,\mathbf{A}^{(4)} + \cdots
880: \label{eq:expand_mat_A3}
881: \end{eqnarray}
882: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
883: for a matrix consisting of the cross-correlation signals
884: $\mathcal{F}_{AT}$ and $\mathcal{F}_{ET}$.
885: Then the resultant matrices $\mathbf{A}^{(i)}$ become independent of the
886: frequency. The above perturbative expansion implies that the output
887: signal $\mathbf{c}$ is also expanded in powers of $\hat{f}$. Since the
888: frequency dependence of the function
889: $H(f)$ has been already factorized in equation (\ref{eq: c=Ap}), we have
890: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
891: \begin{eqnarray}
892: \mathbf{c}(f) =
893: \left\{
894: \begin{array}{lll}
895: & \hat{f}^2 ~ \mathbf{c}^{(2)} +
896: \hat{f}^4 \mathbf{c}^{(4)} + \cdots,
897: \quad\, & \text{for AA-, EE-correlations}
898: \\
899: \\
900: & \hat{f}^2 ~ \mathbf{c}^{(2)} +
901: \hat{f}^3 ~ \mathbf{c}^{(3)} +
902: \hat{f}^4 \mathbf{c}^{(4)} + \cdots,
903: \quad\, & \text{for AE-correlation}
904: \\
905: \\
906: & \hat{f}^3 ~ \mathbf{c}^{(3)} +
907: \hat{f}^4 \mathbf{c}^{(4)} + \cdots,
908: \quad\, &
909: \text{for AT-,ET-correlations}
910: \end{array}
911: \right.
912: \label{eq:expand_vec_c}
913: \end{eqnarray}
914: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
915: Substituting the terms (\ref{eq:expand_mat_A1})-(\ref{eq:expand_mat_A3}) and
916: (\ref{eq:expand_vec_c}) into the expression (\ref{eq: c=Ap}) and
917: collecting the terms of each order of $\hat{f}$, we have
918: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
919: \begin{equation}
920: \mathbf{c}^{(i)} = \mathbf{A}^{(i)}\,\cdot\,\mathbf{p}^{(i)},
921: \quad \quad (i=2,\,\,3,\,\,4,\,\,\cdots)
922: \label{eq:perturbative_c=Ap}
923: \end{equation}
924: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
925: where the subscript $^{(i)}$ means the quantity consisting of the order
926: $\mathcal{O}(\hat{f}^i)$ terms.
927: The vector $\mathbf{p}^{(i)}$ represents the accessible multipole moments
928: $p_{\ell m}$ to which the antenna pattern functions become sensitive
929: in this order. For example, the vector $\mathbf{p}^{(2)}$ contains the
930: $\ell=0,\,2$ and $4$ modes, while the vector $\mathbf{p}^{(3)}$ have
931: the multipole moments with $\ell=1,\,3$ and $5$. Since the expression
932: (\ref{eq:perturbative_c=Ap}) has no explicit frequency dependence, we
933: can reliably apply the least-squares solution by the SVD to reconstruct
934: the odd modes of GWBs, as well as the even modes:
935: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
936: \begin{equation}
937: \mathbf{p}^{(i)}_{\rm\scriptscriptstyle approx }
938: = [\mathbf{A}^{(i)}]^+\,\cdot\,\mathbf{c}^{(i)}.
939: \label{eq:perturbative_p=AC}
940: \end{equation}
941: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
942: Each step of the perturbative reconstruction scheme is summarized in Fig.
943: \ref{fig:flowchart} (see Appendix \ref{appendix:SVD} in more details).
944:
945:
946: %%%%%%%%%%%%%%%%%%%%%% Figure %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
947: \begin{figure}[htbp]
948: \begin{center}
949: \includegraphics[width=13cm,angle=0,clip]{figs/flowchart/flowchart.eps}
950: \end{center}
951: \caption{Flowchart of perturbative reconstruction scheme based on
952: the harmonic-Fourier representation.
953: }
954: \label{fig:flowchart}
955: \end{figure}
956: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
957:
958:
959: Finally, we note that the low-frequency reconstruction method presented
960: here assumes the perturbative expansion form of the output signal $\mathbf{c}$.
961: To determine the coefficients $\mathbf{c}^{(i)}$ in equation
962: (\ref{eq:expand_vec_c}), one must know the frequency dependence of the
963: vector $\mathbf{c}$ in the narrow bandwidth $f\sim f+\Delta f$, which
964: can be achieved by analyzing the multi-frequency data.
965: One important remark is that the signal-to-noise ratio in each
966: reconstructed multipoles might be influenced by the sampling frequencies
967: and/or the data analysis strategy. This point will be discussed later in Sec.
968: \ref{subsec:reconstruction}.
969: %
970: %
971: %
972: %
973: %
974: %
975: %
976: %
977: %
978: %
979: % ///////////////////////////////////////////////////////////// %
980: \subsection{Reconstruction based on the time-series representation}
981: \label{subsec:time-domain}
982: % ///////////////////////////////////////////////////////////// %
983: %
984: %
985: %
986: %
987: %
988: So far, we have discussed the reconstruction method based on the
989: harmonic-Fourier representation (\ref{eq:deconvolution}).
990: The main advantage of the harmonic-Fourier representation is that
991: it gives a simple algebraic equation suitable for applying the
992: least-squares solution by SVD.
993: However, the harmonic-Fourier representation implicitly assumes that
994: the space interferometer orbits around the sun under keeping their
995: configuration rigidly.
996: In reality, rigid adiabatic treatment of the spacecraft motion is no
997: longer valid due to the intrinsic variation of arm-length caused by the
998: Keplerian motion of three space crafts, as well as the tidal
999: perturbation by the gravitational force of solar system planets
1000: \cite{Cornish:2003tz,Tinto:2003vj,Shaddock:2003bc}.
1001: In a rigorous sense, the time dependence of antenna pattern
1002: function cannot be described by the Euler rotation of antenna pattern
1003: function at rest frame (see footnote \ref{footnote:orbital motion}).
1004: Although the influence of arm-length variation is expected to be small
1005: in the low-frequency regime, alternative approach based on the other
1006: representation would be helpful to consider the GWB skymap beyond the
1007: low-frequency regime.
1008:
1009:
1010: Here, we briefly discuss the reconstruction method based on the
1011: time-series representation (\ref{eq:def_of_C(f)}).
1012: The time-series representation is mathematically equivalent to the
1013: harmonic-Fourier representation under the rigid adiabatic treatment,
1014: but in general, no additional assumption for space craft configuration
1015: is invoked to derive the expression
1016: (\ref{eq:def_of_C(f)}), except for the premise that the time-dependence
1017: of the antenna pattern functions, i.e., the motion of each spacecraft,
1018: is well-known theoretically and/or observationally.
1019: Moreover, as we see below, the method based on the time-series
1020: representation allows us to reconstruct directly the skymap $S_h$,
1021: without going through intermediate variables, e.g. $p_{\ell m}$.
1022: In this sense, the time-series representation would be superior to the
1023: harmonic-Fourier representation for practical purpose, although there
1024: still remains the same problem as discussed in Appendix
1025: \ref{appendix:on_the_degeneracy} concerning the {\it{degeneracy}} of the
1026: multipole coefficients.
1027:
1028:
1029: In principle, the perturbative reconstruction scheme given in
1030: Sec.\ref{subsec:general_scheme} can be applicable to the map-making
1031: problem based on the time-series representation.
1032: To apply this, we discretize the integral expression
1033: (\ref{eq:def_of_C(f)}) so as to reduce it to a matrix form like
1034: equation (\ref{eq: c=Ap}). For example, we discretize the celestial sphere
1035: $(\theta, \,\phi)$ into a regular $N$ meshes and also the continuous
1036: time-series into the regular $M$ grids. Then we have
1037: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1038: \begin{equation}
1039: \widetilde{C}(t_i,\,f)= \sum_{j=1}^{N}\,\,
1040: S_h(|f|,\,\,\mathbf{\Omega}_j)\,\,
1041: \mathcal{F}^E(f,\,\mathbf{\Omega}_j;\,t_i)\,
1042: \frac{\Delta\mathbf{\Omega}_j}{4\pi}, \quad
1043: (i=1,2,\cdots,\,M),
1044: \label{eq:discrete_eq}
1045: \end{equation}
1046: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1047: where $\Delta\mathbf{\Omega}_i=\sin\theta_i\,\Delta\theta_i\,\Delta\phi_i$.
1048: Here, we have ignored the noise contribution and dropped the subscript
1049: $_{IJ}$. For a reconstruction of low-frequency skymap, a large number
1050: of mesh and/or grid are not necessary. Typically, for the angular
1051: resolution up to $\ell=5$, it is sufficient to set
1052: $M=16$ and $N=16\time32$ (see Sec.\ref{subsec:reconstruction}).
1053: The above discretization procedure is repeated for the five output
1054: data of the correlation signals.
1055: Then, simply following the same procedure as in the case of the
1056: harmonic-Fourier representation, the least-squares method by SVD
1057: is applied to get the luminosity distribution of GWBs.
1058: Note that the meanings of the matrix $\mathbf{A}^{(i)}$, the vectors
1059: $\mathbf{c}^{(i)}$ and $\mathbf{p}^{(i)}$ in the perturbative expansion
1060: are appropriately changed as follows. For the matrix $\mathbf{A}^{(i)}$, we have
1061: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1062: \begin{equation}
1063: \mathbf{A}^{(i)} = \frac{1}{4\pi}
1064: \left(
1065: \begin{array}{cccc}
1066: \mathcal{F}^{(i)}(\mathbf{\Omega}_1, t_1)\,\Delta\mathbf{\Omega}_1 &
1067: \mathcal{F}^{(i)}(\mathbf{\Omega}_2, t_1)\,\Delta\mathbf{\Omega}_2 & \cdots &
1068: \mathcal{F}^{(i)}(\mathbf{\Omega}_N, t_1)\,\Delta\mathbf{\Omega}_N \\
1069: \mathcal{F}^{(i)}(\mathbf{\Omega}_1, t_2)\,\Delta\mathbf{\Omega}_1 &
1070: \mathcal{F}^{(i)}(\mathbf{\Omega}_2, t_2)\,\Delta\mathbf{\Omega}_2 & \cdots &
1071: \mathcal{F}^{(i)}(\mathbf{\Omega}_N, t_2)\,\Delta\mathbf{\Omega}_N \\
1072: \vdots & \vdots & \ddots &\vdots \\
1073: \mathcal{F}^{(i)}(\mathbf{\Omega}_1, t_M)\,\Delta\mathbf{\Omega}_1 &
1074: \mathcal{F}^{(i)}(\mathbf{\Omega}_2, t_M)\,\Delta\mathbf{\Omega}_2 & \cdots &
1075: \mathcal{F}^{(i)}(\mathbf{\Omega}_N, t_M)\,\Delta\mathbf{\Omega}_N \\
1076: \end{array}
1077: \right), \quad\quad (i=2,\,3,\,\cdots)
1078: \end{equation}
1079: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1080: The corresponding vectors $\mathbf{c}_{(i)}$ and $\mathbf{p}_{(i)}$ becomes
1081: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1082: \begin{equation}
1083: \mathbf{c}^{(i)} =
1084: \left(
1085: \begin{array}{c}
1086: \widetilde{C}^{(i)}(t_1) \\
1087: \widetilde{C}^{(i)}(t_2) \\
1088: \vdots \\
1089: \widetilde{C}^{(i)}(t_M)
1090: \end{array}
1091: \right),
1092: %% \quad\quad(i=2,\,3,\,\cdots)
1093: \qquad
1094: \mathbf{p}^{(i)} =
1095: \left(
1096: \begin{array}{c}
1097: S_h^{(i)}(\Omega_1) \\
1098: S_h^{(i)}(\Omega_2) \\
1099: \vdots \\
1100: S_h^{(i)}(\Omega_N) \\
1101: \end{array}
1102: \right),
1103: \end{equation}
1104: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1105: where $\widetilde{C}^{(i)}(t)$ and $S_h^{(i)}(\Omega)$ the perturbative
1106: coefficients of $\widetilde{C}(t,f)$ and $S_h (|f|, \Omega)$ in power
1107: of $\hat{f}^i$, respectively. Using these expressions, the
1108: least-squares solution is constructed as
1109: $\mathbf{p}_{\rm approx}=
1110: \mathbf{p}_{\rm approx}^{(2)}+\mathbf{p}_{\rm approx}^{(3)}+\cdots$,
1111: with a help of equation (\ref{eq:perturbative_p=AC}).
1112: Then, the resultant expression $\mathbf{p}_{\rm approx}$ directly gives a
1113: GWB skymap in the ecliptic coordinate, i.e.,
1114: $S_h(|f|,\,\mathbf{\Omega})$, not the multipole coefficients.
1115:
1116:
1117:
1118:
1119:
1120: %
1121: % ==============================================================%
1122: % ==============================================================%
1123: \section{Demonstration: skymap of
1124: Galactic background}
1125: \label{sec:demonstration}
1126: % ==============================================================%
1127: % ==============================================================%
1128: %
1129:
1130:
1131: Perturbative reconstruction scheme presented in the previous section is
1132: applicable to a general map-making problem for any kind of GWB sources.
1133: In this section, to see how our general scheme works in practice, we
1134: demonstrate the reconstruction of a GWB skymap focusing on a specific
1135: source of anisotropic
1136: GWB. An interesting example for LISA detector is a confusion-noise
1137: background produced by the Galactic population of unresolved binaries.
1138: After describing a model of Galactic GWB in Sec.\ref{subsec:GWB_model},
1139: the expected signals for time-modulation data of self- and
1140: cross-correlated TDIs are calculated in Sec.\ref{subsec:on_snr}.
1141: Based on these, the detectable Fourier components for time-modulation
1142: signals are discussed evaluating the signal-to-noise ratios.
1143: In Sec.\ref{subsec:reconstruction}, a reconstruction of the GWB skymap
1144: is performed based on the harmonic-Fourier representation and the
1145: time-series representation. The resultant values of the multipole
1146: coefficients for Galactic GWBs are compared with those from the
1147: full-resolution skymap, taking account of the influence of the noises.
1148:
1149:
1150:
1151:
1152: % ///////////////////////////////////////////////////////////// %
1153: \subsection{A model of Galactic GWB}
1154: \label{subsec:GWB_model}
1155: % ///////////////////////////////////////////////////////////// %
1156:
1157:
1158: For our interest of GWBs in the low-frequency regime with
1159: $f\lesssim f_* \simeq 9.54$ mHz, it is reasonable to assume that the
1160: anisotropic GWB spectrum $S_h(f,\mathbf{\Omega})$ is separately treated as
1161: $S_h(f,\mathbf{\Omega})=H(f)\,P(\mathbf{\Omega})$, as we mentioned.
1162: The power spectral density $H(f)$ is approximated by a power-law function like
1163: $H(f)= {\mathrm N } \,f^{\alpha}$, whose amplitude will be explicitly
1164: given later. For illustrative purpose, we consider the simplest model of
1165: luminosity distribution $P(\mathbf{\Omega})$, in which the Galactic GWB
1166: is described by an incoherent superposition of gravitational waves
1167: produced by compact binaries whose spatial structure just traces the
1168: Galactic stellar distribution observed via infrared photometry.
1169: We use the fitting model of Galactic stellar distribution given in
1170: \cite{Binney:1996sv}, which consists of triaxial bulge and disk components
1171: (see also \cite{Seto:2004ji}).
1172: The explicit functional form of the density distribution
1173: $\rho(\mathbf{x})$ written in the Galactic coordinate system becomes
1174: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1175: \begin{eqnarray}
1176: \label{eq:rho_GWB}
1177: && \rho(\mathbf{x}) = \rho_{\rm disk}(\mathbf{x}) +
1178: \rho_{\rm bulge}(\mathbf{x})\,\,;
1179: \label{eq:BGS_model} \\
1180: &&\quad \rho_{\rm bulge} = \frac{\rho_0}{(1+a/a_0)^{1.8}}\,\,
1181: e^{-(a/a_{\rm m})^2},
1182: \nonumber \\
1183: &&\quad \rho_{\rm disk} = \left(\frac{e^{-|z|/z_0}}{z_0}
1184: + \alpha \,
1185: \frac{e^{-|z|/z_1}}{z_1}\right) \,R_s\, e^{-R/R_{\rm s}}
1186: \nonumber
1187: \end{eqnarray}
1188: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1189: with the quantities $a$ and $R$ defined by
1190: $a\equiv [x^2+(y/\eta)^2+(z/\zeta)^2 ]^{1/2}$ and
1191: $R\equiv (x^2 + y^2)^{1/2}$. Note that the $z$-axis is oriented to the north
1192: Galactic pole and the direction of the $x$-axis is $20^{\circ}$
1193: different from the Sun-center line.
1194: Here the parameters are given as follows: $\rho_0=624$,
1195: $a_{\rm m} = 1.9$ kpc, $a_0=100$ pc, $R_{\rm s}=2.5$ kpc, $z_0=210$ pc,
1196: $z_1=42$ pc, $\alpha = 0.27$, $\eta=0.5$ and $\zeta=0.6$.
1197: Provided the three-dimensional structure of stellar distribution
1198: $\rho(\mathbf{x})$, the angular distribution $P(\mathbf{\Omega})$ is
1199: obtained by projecting it onto the sphere in observed frame, i.e.,
1200: ecliptic coordinate:
1201: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1202: \begin{equation}
1203: P(\mathbf{\Omega}) = C\,\int dr\,\,4\pi\,r^2\,\,
1204: \frac{\rho(r,\mathbf{\Omega})}{r^2},
1205: \end{equation}
1206: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1207: where $C$ is a numerical constant normalized by
1208: $\int d\mathbf{\Omega}\,P(\mathbf{\Omega})=1$.
1209: The integration over $r$ is performed along a line-of-sight direction
1210: from a location of space interferometer to infinity.
1211:
1212:
1213: In Fig.\ref{fig:full-resolution_skymap}, the projected intensity
1214: distribution of GWB is numerically obtained specifically in ecliptic
1215: coordinate and it is then transformed into Galactic coordinate, shown
1216: as the Hammer-Aitoff map. A strong intensity peak is found around the
1217: Galactic center and the disk-like structure can be clearly seen.
1218: Notice that while the result depicted in
1219: Fig.\ref{fig:full-resolution_skymap} represents a full-resolution
1220: skymap, it cannot be attained from the perturbative reconstruction
1221: scheme in low-frequency regime. For the antenna pattern functions of
1222: cross-correlated TDI signals up to the order $\mathcal{O}(\hat{f}^3)$,
1223: the detectable multipole moments of GWB are limited to $\ell\leq5$.
1224: Moreover, in the low-frequency limit $\mathcal{O}(\hat{f}^2)$, only
1225: the even modes $\ell=0$, $2$ and $4$ are measurable.
1226: Thus, the reconstructed skymap would be rather miserable compared to
1227: the full skymap which contains the higher multipoles $\ell \gtrsim 30$
1228: (see Fig.\ref{fig:sigma_gwb} in Appendix). Taking account of these
1229: facts, in right panel of Fig.\ref{fig:expected_skymap}, we plot the
1230: low-resolution skymap, which was obtained from the full-resolution
1231: skymap just dropping the higher multipole moments with $\ell>5$.
1232: In Appendix \ref{appendix:multipole of GWB}, with a help of the
1233: Fortran package of spherical harmonic analysis (\cite{Adams:2003}, see
1234: Appendix \ref{appendix:multipole of GWB}), the numerical values of
1235: the multipole coefficients $p_{\ell m}$ up to $\ell=5$ are computed
1236: and are summarized in Table \ref{tab:summay_multipole}.
1237: Also in the left panel, the odd modes are further subtracted and the
1238: remaining multipole moments are only $\ell=0$, $2$ and $4$.
1239: As a result, the fine structure around the bulge and the disk components
1240: is coarse-grained and the intensity of the GWB diminishes.
1241: It also shows some fake patterns with negative intensity.
1242: Nevertheless, one can clearly see the anisotropic structure of GWB
1243: , which is mainly contributed from the gravitational-wave sources
1244: around the Galactic disk. With a perturbative reconstruction of
1245: low-frequency up to $\mathcal{O}(\hat{f}^3)$, one can roughly infer
1246: that the main sources of Galactic GWB comes from the Galactic center.
1247:
1248:
1249:
1250: %%%%%%%%%%%%%%%%%%%%%% Figure %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1251: \begin{figure}[t]
1252: \begin{center}
1253: \includegraphics[width=10cm,angle=0,clip]{figs/map_GC_JpegEPS/gwbdist.eps}
1254: \label{fig:fill_skymap}
1255: \end{center}
1256: \caption{Full-resolution
1257: skymap of the Galactic gravitational-wave background shown
1258: as Hammer-Aitoff map in Galactic coordinate.
1259: }
1260: \label{fig:full-resolution_skymap}
1261: \end{figure}
1262: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1263: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1264: \begin{figure}[t]
1265: \begin{center}
1266: \includegraphics[width=8.3cm,clip]{figs/map_GC_JpegEPS/gwbdist_cut_l0-2-4.eps}
1267: \hspace*{0.3cm}
1268: \includegraphics[width=8.3cm,clip]{figs/map_GC_JpegEPS/gwbdist_cut_l0-5.eps}
1269: \end{center}
1270: \caption{Expected images of GWB skymap by LISA based on
1271: the low-frequency reconstruction scheme, which are both depicted in the
1272: Galactic coordinate. {\it Left}: low-resolution skymap
1273: created by truncating all the multipoles except for
1274: $\ell=0$, $2$ and $4$ from the original full-resolution skymap.
1275: {\it Right:} skymap created by truncating the higher multipoles
1276: $\ell>6$.
1277: }
1278: \label{fig:expected_skymap}
1279: \end{figure}
1280: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1281: %
1282: %
1283: %
1284: %
1285: % ///////////////////////////////////////////////////////////// %
1286: \subsection{Time-modulation signals and signal-to-noise ratio}
1287: \label{subsec:on_snr}
1288: % ///////////////////////////////////////////////////////////// %
1289: %
1290: %
1291: %
1292: %
1293: %
1294: %
1295: Given the intensity distribution of GWB, one can calculate the
1296: cross-correlation signals observed via LISA, which are inherently
1297: time-dependent due to the orbital motion of the LISA detector.
1298: The effect of the annual modulation of the Galactic binary confusion
1299: noise on the LISA data analysis had been previously studied in
1300: \cite{Seto:2004ji} in the low-frequency limit $\hat{f}\ll1$.
1301: Recently, Monte Carlo simulations of Galactic GWB were carried out by
1302: several groups and the annual modulation of GWB intensity has been
1303: confirmed in a realistic setup with specific detector output
1304: \cite{Benacquista:2003th,Edlund:2005ye,Timpano:2005gm}.
1305:
1306:
1307:
1308: Owing to the expression (\ref{eq:def_of_C(f)}), the time-modulation
1309: signals $\widetilde{C}(t,f)$ neglecting the instrumental noises are
1310: computed for optimal TDIs at the frequency $\hat{f}=0.1$, i.e.,
1311: $f\simeq1 \mathrm{mHz}$ using the full expression of antenna pattern function
1312: (\ref{eq:def_of_antenna}). The results are then plotted as function of
1313: orbital phase. In Fig. \ref{fig:annual data}, six outputs of the
1314: self- and cross-correlation signals normalized to its time-averaged value,
1315: $\widetilde{C}(t,f)/|\widetilde{C}_0(f)|$ are shown.
1316: The solid and dashed lines denote the real and imaginary parts of
1317: the correlation signals, respectively.
1318:
1319:
1320: The time modulation of these outputs basically reflects the spatial
1321: structure of GWB. As LISA orbits around the Sun, the direction normal
1322: to LISA's detector plane, which is the direction sensitive to the
1323: gravitational waves, sweeps across the Galactic plane.
1324: Since the response of the LISA detector to the gravitational
1325: waves along the $\mathbf{\hat{\Omega}}$ direction give the same
1326: response to the waves along the
1327: $-\mathbf{\hat{\Omega}}$ direction
1328: (see Eq.(\ref{eq:Y_lm expansion of S and F}) and the brief comment
1329: there), the time-modulation signal is expected to have a bimodal
1330: structure, like $AA$- and $EE$-correlations.
1331: However, the actual modulation signals are more complicated, depending
1332: on the angular sensitivity of their antenna pattern functions, as well
1333: as the observed frequency. Further, the cross-correlation data can be
1334: generally complex variables, whose behaviors are different between
1335: real- and imaginary-parts.
1336:
1337:
1338: %%%%%%%%%%%%%%%%%%%%% Figure %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1339: \begin{figure}[t]
1340: \begin{center}
1341: \includegraphics[width=5.7cm,clip]{figs/signal/signal_AA_f01.eps}
1342: \hspace*{0cm}
1343: \includegraphics[width=5.7cm,clip]{figs/signal/signal_EE_f01.eps}
1344: \hspace*{0cm}
1345: \includegraphics[width=5.7cm,clip]{figs/signal/signal_TT_f01.eps}
1346: \end{center}
1347: \begin{center}
1348: \includegraphics[width=5.7cm,clip]{figs/signal/signal_AE_f01.eps}
1349: \hspace*{0cm}
1350: \includegraphics[width=5.7cm,clip]{figs/signal/signal_AT_f01.eps}
1351: \hspace*{0cm}
1352: \includegraphics[width=5.7cm,clip]{figs/signal/signal_ET_f01.eps}
1353: \end{center}
1354: \caption{
1355: Annual time-modulation signals of self- and cross-correlation data
1356: as function of orbital phase assuming the observed frequency
1357: $\hat{f}=0.1$, i.e., $f\simeq1 \mathrm{mHz}$.
1358: Here, the correlation signals
1359: $\widetilde{C}_{IJ}(t,f)$ are plotted normalizing
1360: with $k=0$ component of correlation signal $\widetilde{C}_{IJ,0}(f)$.
1361: Solid and dashed lines represent the real and the imaginary
1362: part of correlation signal, respectively.
1363: }
1364: \label{fig:annual data}
1365: \end{figure}
1366: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1367:
1368:
1369:
1370: Notice that the time-modulation signals presented in Fig.
1371: \ref{fig:annual data} are the results in an idealistic situation free
1372: from the noise contribution. In presence of the random noise, some of
1373: the correlation data which contain gravitational wave signals are
1374: overwhelmed by noise, which cannot be used for the reconstruction of
1375: GWB skymap. Hence, one must consider the signal-to-noise ratio (SNR)
1376: to discriminate available correlation data.
1377: To evaluate this, the output signals depicted in
1378: Fig.\ref{fig:annual data} are first transformed to its Fourier counterpart,
1379: $\widetilde{C}_k(f)$. The resultant Fourier components for each signal
1380: are then compared to the noise contributions.
1381: The SNR for each component is expressed in the following form
1382: (\cite{Seto:2004np}, see also
1383: \cite{Giampieri:1997,Ungarelli:2001xu}):
1384: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1385: \begin{equation}
1386: \left(\frac{S}{N}\right)_k =\sqrt{2\,\,\Delta f\,\,T}\,\,
1387: \frac{|\widetilde{C}_k|}{N_k}.
1388: \label{eq:def_SNR}
1389: \end{equation}
1390: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1391: Here we set the observational time to $T=10^8$sec and the bandwidth to
1392: $\Delta f=10^{-3}$Hz. The quantity $N_k$ represents the noise
1393: contribution. The important remark is that the noise contribution
1394: comes from not only the detector noise but also the randomness of
1395: the signal itself. The details of the analytic expression for $N_k$
1396: will be given elsewhere \cite{Kudoh:2005inprep}.
1397: Here, as a crude estimate, we evaluate the noise contribution $N_k$ as
1398: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1399: \begin{equation}
1400: N_k= \sqrt{2\,\,\Delta f\,\,T}\,\,S_n^{II}, \quad(I= A,~E,~T)
1401: \end{equation}
1402: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1403: for $k=0$ component of self-correlation signals and
1404: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1405: \begin{equation}
1406: N_k= \sqrt{\mbox{max}\left(\widetilde{C}_{II,0}\widetilde{C}_{JJ,0},
1407: ~\widetilde{C}_{II,0}S_n^{JJ},~
1408: \widetilde{C}_{JJ,0}S_n^{II},~S_n^{II}S_n^{JJ}\right) },
1409: \quad (I,J= A,~E,~T)
1410: \end{equation}
1411: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1412: for cross-correlation signals and $k\neq0$ component of self-correlation signals.
1413: To estimate SNR, one further needs the spectral density for instrumental
1414: noise, i.e., $S_n^{\scriptscriptstyle\rm AA}(f)$,
1415: $S_n^{\scriptscriptstyle\rm EE}(f)$ and $S_n^{\scriptscriptstyle\rm
1416: TT}(f)$. We use the expressions given in equation (58) of Paper I
1417: (see also \cite{Cornish:2001bb,Prince:2002hp})
1418: \footnote{In equation (58) of Paper I, there are some typos in the
1419: numerical values of $S_{\rm shot}(f)$ and $S_{\rm accel}(f)$.
1420: In the present paper, we adopt $S_{\rm shot}(f)=1.6\times10^{-41}$Hz$^{-1}$
1421: and $S_{\rm accel}(f)=2.31\times10^{-41}(\mbox{mHz}/f)^4$Hz$^{-1}$
1422: according to Refs.\cite{Bender:1998,Cornish:2001bb}.}.
1423: Note that all the cross-correlated noise spectra such as
1424: $S_n^{\scriptscriptstyle\rm AE}(f)$ and $S_n^{\scriptscriptstyle\rm AT}(f)$
1425: are exactly canceled.
1426:
1427:
1428:
1429:
1430: In Fig.\ref{fig:SNR}, the SNRs for six output data are evaluated and
1431: are shown in the histogram as function of Fourier component, $k$.
1432: In these panels, thick-dotted lines show the detection limit of
1433: $(S/N)_k=5$, while the thin-dotted lines mean $(S/N)_k=1$. When
1434: evaluating the SNR, we specifically consider the two cases:
1435: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1436: \begin{description}
1437: \item[Case A:]\quad realistic case in which
1438: the rms amplitude of GWB spectrum is given by
1439: $S_h^{1/2}=5\times10^{-19}$Hz$^{-1/2}$ at the frequency
1440: $f=1$mHz \footnote{The amplitude of GWB spectrum might be reduced
1441: by a factor of $2\sim5$ according to the recent numerical simulations
1442: \cite{Edlund:2005ye,Timpano:2005gm}.
1443: In case A, however, the noise contributions in SNR (\ref{eq:def_SNR})
1444: used for reconstruction analysis are basically determined by the $k=0$
1445: components of self-correlation signals, i.e.,
1446: $N_k=\sqrt{\widetilde{C}_{II,0}\widetilde{C}_{JJ,0}}$.
1447: Hence, a slight change of the GWB amplitude would not alter the final results.}.
1448: \item[Case B:]\quad optimistic case in which the rms amplitude of GWB
1449: spectrum is ten times larger than that in the realistic case, i.e.,
1450: $S_h^{1/2}=5\times10^{-18}$Hz$^{-1/2}$ at $f=1$mHz.
1451: \end{description}
1452: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1453: The results of SNR are then shown in solid (case A) and dashed lines
1454: (case B), respectively.
1455:
1456:
1457:
1458: As anticipated from the sensitivity of antenna pattern function in Table
1459: \ref{tab:summay_antenna}, the SNR for $TT$-correlation is much less than unity.
1460: Even in the optimistic case, the SNR is about ten times smaller than
1461: unity and thus the $TT$-correlation data cannot be used for
1462: reconstruction of the GWB skymap. Apart from this, the SNRs for the
1463: self-correlation signals $AA$ and $EE$ as well as for the
1464: cross-correlation signal $AE$ are generally good compared to the
1465: cross-correlation data $AT$ and $ET$. With sufficient higher SNR of
1466: $(S/N)_k\geq5$, the available Fourier components of $AA$-, $EE$- and
1467: $AE$-correlations become
1468: $k=-2,\,-1,\,0,\,+1,\,+2$ in both realistic and optimistic cases.
1469: This is consistent with the previous estimates \cite{Seto:2004np}.
1470: \footnote{The precise numerical values of the SNR for $AA$- and
1471: $EE$-correlations slightly differ from Ref.~\cite{Seto:2004np}.
1472: This is mainly because the Euler rotation angles for the LISA orbital
1473: motion given in Sec.\ref{subsec:correlation} are different from
1474: those in \cite{Seto:2004np}. }
1475: On the other hand, with a large amplitude of GWB spectrum (case B),
1476: only the $k=0$ component is accessible in the signal combinations of
1477: $AT$ and $ET$. This is mainly due to the fact that the sensitivity of
1478: the $T$-variable is poor at low-frequency and thereby the noise
1479: contribution $N_k$ becomes
1480: $\{\widetilde{C}_{AA,0}S_n^{\rm\scriptscriptstyle TT}\}^{1/2}$ or
1481: $\{\widetilde{C}_{EE,0}S_n^{\rm\scriptscriptstyle TT}\}^{1/2}$,
1482: which is much larger than
1483: $\{\widetilde{C}_{AA,0}\widetilde{C}_{TT,0}\}^{1/2}$ or
1484: $\{\widetilde{C}_{EE,0}\widetilde{C}_{TT,0}\}^{1/2}$.
1485:
1486:
1487: Thus, the available Fourier components of correlation data used for
1488: the reconstruction of the skymap would be severely restricted in practice.
1489: Under such restricted situation, the deconvolution problem of the linear system
1490: (\ref{eq:deconvolution}) tends to be \textit{under-determined}.
1491: Nevertheless, as it will be shown below, one can determine the $\ell=0$,
1492: $2$ and $4$ modes of multipole coefficients of the Galactic GWB with
1493: sufficiently small errors.
1494: In addition, with the $k=0$ components of $AT$- and $ET$-correlations,
1495: the odd moments $\ell=1$ and $3$ can be recovered.
1496:
1497:
1498: %%%%%%%%%%%%%%%%%%%%% Figure %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1499: \begin{figure}[t]
1500: \begin{center}
1501: \includegraphics[width=5.7cm,clip]{figs/snr/snr_AA_f01.eps}
1502: \hspace*{0cm}
1503: \includegraphics[width=5.7cm,clip]{figs/snr/snr_EE_f01.eps}
1504: \hspace*{0cm}
1505: \includegraphics[width=5.7cm,clip]{figs/snr/snr_TT_f01.eps}
1506: \end{center}
1507: \begin{center}
1508: \includegraphics[width=5.7cm,clip]{figs/snr/snr_AE_f01.eps}
1509: \hspace*{0cm}
1510: \includegraphics[width=5.7cm,clip]{figs/snr/snr_AT_f01.eps}
1511: \hspace*{0cm}
1512: \includegraphics[width=5.7cm,clip]{figs/snr/snr_ET_f01.eps}
1513: \end{center}
1514: \caption{
1515: Signal-to-noise ratio for each Fourier component of
1516: correlation signals, $(S/N)_k$ estimated at
1517: $\hat{f}=0.1$ ($f\simeq1$mHz). The histogram depicted in solid line
1518: indicates the case with the amplitude of Galactic GWB given by
1519: $S_h^{1/2}=5\times10^{-19}$Hz$^{-1/2}$ at $f=1$mHz (case A),
1520: while the histogram in dashed
1521: line represents the signal-to-noise ratio for the case with
1522: $S_h^{1/2}=5\times10^{-18}$Hz$^{-1/2}$ (case B). The thin-dashed,
1523: and thick-dashed lines respectively denote
1524: the lines of $(S/N)_k=1$ and $5$.
1525: }
1526: \label{fig:SNR}
1527: \end{figure}
1528: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1529: %
1530: %
1531: %
1532: %
1533: %
1534: %
1535: %
1536: % ///////////////////////////////////////////////////////////// %
1537: \subsection{Reconstruction of a skymap}
1538: \label{subsec:reconstruction}
1539: % ///////////////////////////////////////////////////////////// %
1540:
1541:
1542: Keeping the remarks on the SNR estimation in previous subsection in
1543: mind, we now proceed to a reconstruction analysis and test the validity
1544: of perturbative reconstruction method presented in Sec. \ref{sec:method}.
1545: For this purpose, in addition to the analysis in the under-constrained cases
1546: (case A and B) mentioned above, we also consider the over-constrained
1547: case as an illustrative example.
1548:
1549:
1550: %%%%%%%%%%%%%%%%%%%%%%%%%%% TABLE %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1551: \begin{table}[b]
1552: \caption{
1553: Components of $\widetilde{C}_k$ used for reconstruction analysis based
1554: on the harmonic-Fourier representation}
1555: \begin{ruledtabular}
1556: \begin{tabular}{c|cccccc}
1557: & AA & EE & TT & AE & AT & ET
1558: \\
1559: \hline
1560: Over-determined case & $-2\leq k\leq +2$ & $-2\leq k \leq +2$ & none
1561: & $-8 \leq k \leq +8$ & $-4 \leq k \leq +4$ & $-4 \leq k \leq +4$
1562: \\
1563: Under-determined case & & & & & & \\
1564: case A & $-2\leq k\leq +2$ & $-2\leq k \leq +2$ & none
1565: & $-2 \leq k \leq +2$ & none & none
1566: \\
1567: case B & $-2\leq k\leq +2$ & $-2\leq k \leq +2$ & none
1568: & $-2 \leq k \leq +2$ & $k=0$ & $k=0$
1569: \end{tabular}
1570: \label{tab:components_used}
1571: \end{ruledtabular}
1572: \end{table}
1573: %%%%%%%%%%%%%%%%%%%%%%%%%%% TABLE %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1574:
1575:
1576: % ------------------------------------------------------------- %
1577: \subsubsection{Over-determined case}
1578: \label{subsubsec:over-determined}
1579: % ------------------------------------------------------------- %
1580:
1581: \textit{(I) Harmonic-Fourier representation} \quad
1582: Let us first focus on a very idealistic situation that the noise
1583: contributions are entirely neglected. In such a case, all the components
1584: of self- and cross-correlation data $\widetilde{C}_k$ are available to
1585: the reconstruction analysis.
1586: In practice, however, it is sufficient to consider some restricted
1587: components among all available data.
1588: Here, to make a skymap with multipoles of $\ell\leq5$, we use the
1589: $k=-2\sim+2$ components of self-correlation signals
1590: $AA$ and $EE$, the $k=-4\sim+4$ components of cross-correlation signals
1591: $AT$ and $ET$, and $k=-8\sim+8$ components of $AE$-signal
1592: (see Table \ref{tab:components_used}). Even in this case, the linear
1593: system (\ref{eq:deconvolution}) is still over-determined and the
1594: least-squares approximation by SVD is potentially powerful to obtain
1595: the multipole coefficients of anisotropic GWB.
1596: The procedure of reconstruction analysis is the same one as presented in
1597: Fig.\ref{fig:flowchart}.
1598: For the output signals of harmonic-Fourier representation, we use the
1599: data $\widetilde{C}_k$ observed at the frequencies
1600: $\hat{f}=0.05$ and $0.15$, in addition to the data for our interest at
1601: $\hat{f}=0.1$.
1602: Collecting these multi-frequency data, the perturbative expansion form
1603: of the vector $\mathbf{c}(f)$ is specified up to the third order in
1604: $\hat{f}$ and the coefficients $\mathbf{c}^{(i)}$ are determined
1605: (Appendix \ref{appendix:SVD}).
1606:
1607:
1608:
1609: In Fig.\ref{fig:reconstructed_skymap1}, the reconstructed results of
1610: multipole coefficients $p_{\ell m}$ are converted to the projected
1611: intensity distribution
1612: (\ref{eq:Y_lm expansion of S and F}) and are shown as Hammer-Aitoff map
1613: in Galactic coordinate.
1614: Left panel shows the skymap reconstructed from the lowest-order signals
1615: of self- and cross-correlation data
1616: $\mathbf{c}^{(2)}$, which only includes the $\ell=0$, $2$ and $4$ modes,
1617: while right panel is the result taking account of the leading-order correction
1618: $\mathbf{c}^{(3)}$. Comparing those with the expected skymap shown in
1619: Fig.\ref{fig:expected_skymap}, the reconstruction seems almost perfect.
1620: In Fig. \ref{fig:reconstructed_plm1}, numerical values of the
1621: reconstructed multipole coefficients are compared with the true values
1622: listed in Table \ref{tab:summay_multipole}. The agreement between the
1623: reconstruction results ({\it open circles}) and the true values
1624: ({\it crosses}) is quite good and the fractional errors are well within
1625: a few percent except for $p_{20}$ and $p_{50}$.
1626: A remarkable fact is that the monopole and the quadrupole values can be
1627: reproduced reasonably well despite the presence of the degeneracy
1628: mentioned in Appendix.\ref{appendix:on_the_degeneracy}.
1629: This is just an accidental result. The (small) discrepancy in the ``recovered''
1630: $p_{20}$ can be ascribed to the expression (\ref{eq:fake_solution}).
1631: After all, the least-squares method by SVD provides a robust
1632: reconstruction method when the linear system (\ref{eq:deconvolution})
1633: becomes over-determined.
1634:
1635:
1636: %%%%%%%%%%%%%%%%%%%%% Figure %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1637: \begin{figure}[t]
1638: \begin{center}
1639: \includegraphics[width=8.3cm,clip]{figs/map_GC_JpegEPS/gwb_reconst_l0-2-4.eps}
1640: \hspace{0.3cm}
1641: \includegraphics[width=8.3cm,clip]{figs/map_GC_JpegEPS/gwb_reconst_l0-5.eps}
1642: \end{center}
1643: \caption{Reconstructed skymap from the time-modulation signals in
1644: the {\it{over-determined}} case. The results were obtained
1645: by the method based on harmonic-Fourier representation and are
1646: plotted in the Galactic coordinate.
1647: The left panel shows the leading-order result,
1648: where $\ell=0$, $2$ and $4$ modes are only reconstructed, while
1649: the right panel represents the result including all reconstructed
1650: multipoles ($\ell\leq5$).
1651: }
1652: \label{fig:reconstructed_skymap1}
1653: \end{figure}
1654: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1655: %%%%%%%%%%%%%%%%%%%%% Figure %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1656: \begin{figure}[tp]
1657: \begin{center}
1658: \includegraphics[width=8.3cm,clip]{figs/error/plm.eps}
1659: \end{center}
1660: \caption{
1661: Reconstructed values of multipole coefficients $|p_{\ell m}|$
1662: in over-determined case. These numerical values are
1663: evaluated in the ecliptic frame.
1664: The open circles represent the reconstruction results, while the
1665: crosses mean the true values, which are obtained from the spherical
1666: harmonic expansion of full-resolution skymap
1667: (see Table \ref{tab:summay_multipole}).}
1668: \label{fig:reconstructed_plm1}
1669: \end{figure}
1670: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1671:
1672:
1673: \textit{(II) Time-series representation} \quad
1674: The successful reconstruction of the GWB skymap can also be achieved
1675: by the alternative approach based on the time-series representation
1676: (\ref{eq:def_of_C(f)}). Following the procedure presented in
1677: Sec.\ref{subsec:time-domain}, the intensity skymap of the GWB is
1678: directly obtained and the results taking account of the lowest-order
1679: and the leading-order contributions to the antenna pattern functions
1680: are shown in left and right panels in
1681: Fig.\ref{fig:reconstructed skymap2}, respectively.
1682: Here, to create the discretized data set (\ref{eq:discrete_eq}), the
1683: number of grid and/or mesh was specified as $M=16$ in time and
1684: $N=17\times32$ in spherical coordinate. The time-series data of antenna
1685: pattern functions were numerically generated based on the full analytic
1686: expressions given in Sec.\ref{subsec:antenna_pattern} under assuming
1687: that the arm-length of the three space crafts are rigidly kept fixed.
1688: The reconstructed skymap reasonably agrees with
1689: Fig.\ref{fig:reconstructed_skymap1} as well as Fig.\ref{fig:expected_skymap}.
1690: Although the situation considered here is very idealistic and thus the
1691: results in Fig.\ref{fig:reconstructed skymap2} should be regarded as
1692: just a preliminary one, one expects that the methodology based on the
1693: time-series representation is potentially powerful even when the rigid
1694: adiabatic treatment of space craft motion becomes inadequate.
1695: To discuss its effectiveness, a further investigation is needed.
1696: The details of the analysis including the effects of arm-length
1697: variation will be presented elsewhere.
1698:
1699:
1700:
1701:
1702: %%%%%%%%%%%%%%%%%%%%% Figure %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1703: \begin{figure}[t]
1704: \begin{center}
1705: \includegraphics[width=8cm,clip]{figs/map_GC_JpegEPS/gwb_reconst_l0-2-4_time.eps}
1706: \includegraphics[width=8cm,clip]{figs/map_GC_JpegEPS/gwb_reconst_l0-5_time.eps}
1707: \end{center}
1708: \caption{Reconstructed skymap from the time-modulation signals
1709: based on the time-series representation. Here,
1710: to perform the
1711: reconstruction analysis, the number of grid and/or mesh
1712: for the discretized data set (\ref{eq:discrete_eq})
1713: was specified as $M=16$ in time and $N=17\times32$ in
1714: spherical coordinate.
1715: The left panel shows the lowest-order result
1716: in which only the $\ell=0$, $2$ and $4$ modes are included,
1717: while the right panel represents the result taking account of
1718: the leading-order correction to the antenna pattern functions,
1719: which includes the multipoles, $\ell \leq5$.
1720: }
1721: \label{fig:reconstructed skymap2}
1722: \end{figure}
1723: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1724:
1725:
1726: % ------------------------------------------------------------- %
1727: \subsubsection{Under-determined case}
1728: \label{subsubsec:under-determined}
1729: % ------------------------------------------------------------- %
1730:
1731:
1732:
1733: Turning to focus on the analysis based on the harmonic-Fourier
1734: representation, we next consider the under-constrained case in which
1735: the number of available Fourier components is restricted due to the
1736: noises (case A and B discussed in Sec.\ref{subsec:on_snr}).
1737: Fig.\ref{fig:reconstructed skymap3} shows the reconstructed images of
1738: GWB intensity map in Galactic coordinate, free from the noises but ]
1739: restricting the number of Fourier components according to Table
1740: \ref{tab:components_used}.
1741: The top panel is the intensity map obtained from the lowest-order signals
1742: $\mathbf{c}^{(2)}$.
1743: Since the accessible Fourier components are basically the same in
1744: the lowest-order analysis, the same results are obtained between case A and B.
1745: On the other hand, the bottom panels of Fig.\ref{fig:reconstructed
1746: skymap3} represent the skymap reconstructed from both
1747: $\mathbf{c}^{(2)}$ and $\mathbf{c}^{(3)}$, which indicate that the
1748: different images of GWB skymap are obtained depending on the available
1749: number of Fourier components (or the amplitude of GWB spectrum);
1750: case A ({\it left}) and case B ({\it right}).
1751: In Fig.\ref{fig:reconstructed_plm3}, numerical values of the
1752: reconstructed multipoles $p_{\ell m}$ in ecliptic coordinate are
1753: summarized together with the statistical errors.
1754: The statistical errors were roughly estimated according to the
1755: discussion in Appendix \ref{appendix:statistical_error} based on the SNR
1756: [Eq. (\ref{eq:def_SNR})].
1757:
1758:
1759: It turns out that the case A fails to reconstruct all the dipole
1760: moments, while they can be somehow reproduced in the case B.
1761: This is because the cross-correlation data $AT$ and $ET$ which include
1762: the information about $\ell=1$ modes were not used in the reconstruction
1763: analysis in case A. Although the $\ell=5$ modes of multipole
1764: coefficients
1765: were not reproduced well in both cases, their contributions to the
1766: intensity distribution are not large. Hence, the reconstructed GWB
1767: skymap in case B roughly matches the expected intensity map in
1768: Fig.\ref{fig:expected_skymap} and the visual impression becomes better
1769: than case A. This readily implies that large amplitude of the GWB
1770: spectrum is required for a correct reconstruction of a skymap,
1771: in practice. However, it should be emphasized that the present
1772: reconstruction technique can work well in the under-determined cases.
1773: Even in the realistic situation with smaller amplitude of GWB
1774: spectrum (case A), the reconstructed skymap including the multipoles
1775: $\ell<5$ still shows a disk-like structure, which may provide an
1776: important clue to discriminate between the Galactic and the
1777: extragalactic GWBs.
1778:
1779:
1780:
1781: %%%%%%%%%%%%%%%%%%%%% Figure %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1782: \begin{figure}[t]
1783: \begin{center}
1784: \includegraphics[width=8.3cm,clip]{figs/map_GC_JpegEPS/gwb_reconst_restricted_l0-2-4.eps}
1785: \end{center}
1786: \begin{center}
1787: \includegraphics[width=8.3cm,clip]{figs/map_GC_JpegEPS/gwb_reconst_restricted_l0-5.eps}
1788: \hspace*{0.3cm}
1789: \includegraphics[width=8.3cm,clip]{figs/map_GC_JpegEPS/gwb_reconst_restricted2_l0-5.eps}
1790: \end{center}
1791: \caption{Reconstructed skymap from the time-modulation signals
1792: using the restricted Fourier components in {\it{under-determined}} cases
1793: (case A, B). These are all plotted in Galactic coordinate.
1794: The upper panel shows the result from lowest-order
1795: analysis in which only the multipole coefficients of
1796: $\ell=0,\,2$ and $4$ modes are considered.
1797: The bottom panels are the intensity map taking account of the
1798: leading-order correction including the multipoles, $\ell<5$. Left and
1799: right panels respectively show the results obtained from the case A and B.
1800: Note that the available number of Fourier components was restricted in
1801: the reconstruction analysis according to Table \ref{tab:components_used}. }
1802: \label{fig:reconstructed skymap3}
1803: \end{figure}
1804: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1805: %%%%%%%%%%%%%%%%%%%%% Figure %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1806: \begin{figure}[ht]
1807: \begin{center}
1808: \includegraphics[width=8.3cm,clip]{figs/error/plm_restricted1.eps}
1809: \hspace*{0.3cm}
1810: \includegraphics[width=8.3cm,clip]{figs/error/plm_restricted2.eps}
1811: \end{center}
1812: \caption{Reconstructed values of the multipole coefficients
1813: $|p_{\ell m}|$ in the under-constrained cases.
1814: The numerical values are evaluated in the ecliptic coordinates.
1815: Left (right) panel shows the result in case A (case B). The open
1816: circles with error bar represent the reconstruction results, while
1817: the crosses mean the true values. The error bars represent the
1818: statistical error in the signal processing. The evaluation of the error
1819: is discussed in Appendix \ref{appendix:statistical_error}.
1820: }
1821: \label{fig:reconstructed_plm3}
1822: \end{figure}
1823: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1824: %
1825: %
1826: %
1827: %
1828: %
1829: %
1830: %
1831: %
1832: %
1833: %
1834: %
1835: %
1836: % ==============================================================%
1837: % ==============================================================%
1838: \section{Conclusion and Discussion}
1839: \label{sec:conclusion}
1840: % ==============================================================%
1841: % ==============================================================%
1842: %
1843: %
1844: %
1845: %
1846: In this paper, we have presented a perturbative reconstruction method to
1847: make a skymap of GWB observed via space interferometer.
1848: The orbital motion of the detector makes the output signals of GWB
1849: time-dependent due to the presence of anisotropies of GWB.
1850: Since the output signals of GWB are obtained through an all-sky integral
1851: of primary signals convolving with an antenna pattern function of
1852: gravitational-wave detectors, the time dependence of output data can
1853: be used to reconstruct the luminosity distribution of GWBs under full
1854: knowledge of detector's antenna pattern functions.
1855: Focusing on the low-frequency regime, we have explicitly given a
1856: non-parametric reconstruction method based on both the
1857: \textit{harmonic-Fourier} and the \textit{time-series} representation.
1858: With a help of low-frequency expansion of the antenna pattern functions,
1859: the least-squares approximation by SVD enables us to obtain the
1860: multipole coefficients of GWB or direct intensity map even when the
1861: system becomes over-determined or under-determined.
1862: For illustrative purpose, the reconstruction analysis of the GWB skymap
1863: has been demonstrated for the confusion-noise background of Galactic
1864: binaries around the low-frequency $f=1$mHz. It then turned out that the
1865: space interferometer LISA free from the noises is capable of making a
1866: skymap of Galactic GWB with angular resolution up to the multipoles,
1867: $\ell=5$. For more realistic case based on the estimation of
1868: signal-to-noise ratios, the system tends to become under-determined
1869: and the number of Fourier components used in reconstruction analysis
1870: would be severely restricted.
1871: Nevertheless, the resultant skymap still contains the information
1872: of the multipoles up to $\ell<5$, from which one can infer that
1873: the main sources of GWB come from the Galactic center.
1874:
1875:
1876: Although the present paper focuses on the reconstruction of the GWB
1877: skymap from the LISA, the methodology discussed in Sec.\ref{sec:method}
1878: is quite general and is also applicable to the reconstruction of
1879: low-frequency skymap obtained from the ground detectors as well as
1880: the other space interferometers, provided their antenna pattern functions.
1881: Despite a wide applicability of the present method, however, accessible
1882: multipole moments of low-frequency GWB are generally restricted to lower
1883: multipoles due to the properties of antenna pattern functions.
1884: This would be generally true in any gravitational-wave detectors
1885: at the frequencies $f<f_*=c/(2\pi L)$, where $L$ is arm-length of
1886: single detector or separation between the two detectors.
1887: Hence, with the limited angular resolution, only the present methodology
1888: may not give a powerful constraint on the luminosity distribution of
1889: GWBs. In this respect, we need to combine the other techniques such as
1890: the parametric reconstruction method, in which we specifically assume an
1891: explicit functional form of the luminosity distribution characterized by
1892: the finite number of model parameters and determine them through the
1893: likelihood analysis. The data analysis strategy to give a tight
1894: constraint on the luminosity distribution is definitely a very important
1895: issue and it must be considered urgently.
1896: Nevertheless, we note that detectable multipole moments depend
1897: practically on the signal-to-noise ratios. Since the signal-to-noise
1898: ratios are determined both from the amplitude of GWB and detector's
1899: intrinsic noises, a further feasibility study is necessary in order
1900: to clarify the detectable multipole moments correctly.
1901: With improved data analysis strategy, it might be even possible that
1902: the angular resolution of GWB skymap becomes better than that of LISA
1903: considered in the present paper.
1904:
1905:
1906: Another important issue on the map-making problem is to consider the
1907: reconstruction of skymap beyond the low-frequency regime, where the
1908: antenna pattern functions give a complicated response to the anisotropic
1909: GWBs and thereby the angular resolution of GWB map can be improved
1910: \cite{Kudoh:2004he}.
1911: Since the low-frequency expansion cannot be used there, a new
1912: reconstruction technique should be devised to extract the information
1913: of anisotropic GWBs.
1914: Further, in the case of LISA, the effect of arm-length variation becomes
1915: important and the rigid adiabatic treatment of the detector response
1916: cannot be validated. Hence, as emphasized in
1917: Sec.\ref{subsec:time-domain}, it would become essential to consider the
1918: reconstruction method based on the time-series representation, in which
1919: no additional assumptions for spacecraft configuration and/or motion are
1920: needed to compute the antenna pattern functions. The analysis concerning
1921: this issue will be presented elsewhere.
1922: %
1923: %
1924: %
1925: %
1926: %
1927: %
1928: %
1929: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1930: \begin{acknowledgments}
1931: We would like to thank N. Seto for valuable comments on the estimation of
1932: signal-to-noise ratios.
1933: We also thank Y. Himemoto and T. Hiramatsu for discussions and comments,
1934: T. Takiwaki and K. Yahata for the technical support to plot the skymap.
1935: The work of H.K. is supported by the
1936: Grant-in-Aid for Scientific Research of Japan Society for Promotion of Science.
1937: \end{acknowledgments}
1938: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1939: %
1940: %
1941: %
1942: %
1943: %
1944: %
1945: \appendix
1946: % -------------------------------------------------------------%
1947: % -------------------------------------------------------------%
1948: \section{Multipole coefficients of antenna
1949: pattern functions in low-frequency regime}
1950: \label{appendix:Mulipole coefficients}
1951: % -------------------------------------------------------------%
1952: % -------------------------------------------------------------%
1953: %
1954: %
1955: %
1956: %
1957: %
1958: In this appendix, based on the analytic expression (21) of paper I,
1959: the multipole moments for antenna pattern functions of optimal TDIs
1960: are calculated at detector's rest frame.
1961: Using the low-frequency approximation $\hat{f}\ll1$, we present the
1962: perturbative expressions up to the forth order in $\hat{f}$.
1963:
1964:
1965: To evaluate the antenna pattern functions, we must first specify the
1966: directional unit vectors $\mathbf{a}$, $\mathbf{b}$ and $\mathbf{c}$,
1967: which connect respective spacecrafts.
1968: Here, we specifically choose (Fig. 1 and Eq.(28) of Paper I):
1969: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1970: \begin{eqnarray}
1971: \mathbf{a} =
1972: -\frac{\sqrt{3}}{2}\,\mathbf{x} + \frac{1}{2}\,\mathbf{y} ,
1973: \quad\quad
1974: \mathbf{b} = -\mathbf{y} ,
1975: \quad\quad
1976: \mathbf{c} =
1977: \frac{\sqrt{3}}{2}\,\mathbf{x} + \frac{1}{2}\,\mathbf{y} ,
1978: \nonumber
1979: \end{eqnarray}
1980: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1981: where the vectors $\mathbf{x}$ and $\mathbf{y}$ respectively denote
1982: the unit vectors parallel to the $x$- and $y$-axes in detector's rest
1983: frame (see Fig.1 of Paper I).
1984: Then, based on the expression (\ref{eq:def AET mode}), the antenna
1985: pattern functions of the optimal TDIs, $\mathcal{F}_{IJ}$
1986: $(I,J=A,E,T)$ are analytically computed at detector's rest frame.
1987: Their multipole moments become
1988: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1989: \begin{equation}
1990: a_{\ell m}(\hat{f}) = \int_0^{\pi}d\theta\,\int_0^{2\pi}\,d\phi\,
1991: \sin\theta\,\,Y_{\ell m}^*(\theta,\phi)\,\,\mathcal{F}(\hat{f},\theta,\phi),
1992: \end{equation}
1993: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1994: which are expressed as a function of normalized frequency
1995: $\hat{f}=f/f_*$.
1996: Here, for definiteness, we write down the explicit form of the harmonic
1997: functions $Y_{\ell m}$:
1998: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1999: \begin{eqnarray}
2000: Y_{\ell}^m (\theta,\phi)
2001: = \sqrt{\frac{2\ell +1}{4\pi}\,\frac{(\ell-m)!}{(\ell+m)!}} \,\,
2002: P_{\ell}^m(\cos\theta)\,\, e^{i m \phi}.
2003: \label{eq:harmonic_Y_lm}
2004: \end{eqnarray}
2005: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2006:
2007:
2008: The analytic expressions for the multipole moments $a_{\ell m}$ are
2009: generally intractable due to the complicated form of the antenna
2010: pattern functions. In the low-frequency regime, however, the
2011: perturbative treatment regarding $\hat{f}$ as a small expansion
2012: parameter is applied to derive an analytic expression of multipole
2013: moments. Below, we summarize the perturbation results up to the fourth
2014: order in $\hat{f}$:
2015: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2016: \begin{eqnarray}
2017: {\mathcal F}_{AA} :
2018: &&
2019: a_{00}= \frac{2\sqrt{\pi}}{5}\,\hat{f}^2 -
2020: \frac{211\sqrt{\pi}}{1260}\,\hat{f}^4,
2021: \quad
2022: a_{20}=\frac{4}{7}\sqrt{\frac{\pi}{5}}\,\hat{f}^2-
2023: \frac{16}{63}\sqrt{\frac{\pi}{5}}\,\hat{f}^4
2024: \quad
2025: a_{22}= - \frac{1-i \,\sqrt{3}}{756}
2026: {\sqrt{\frac{15\pi }{2}}}\,{\hat f}^4,
2027: \cr
2028: &&
2029: a_{40}= \frac{\sqrt{\pi}}{105}\,{\hat f}^2 +
2030: \frac{13\sqrt{\pi}}{9240}\,{\hat f}^4,
2031: \quad
2032: a_{42}= \frac{13 (1-i{\sqrt{3}})}{5544} \sqrt{\frac{\pi}{10}}\,{\hat f}^4,
2033: \quad
2034: a_{44}=-\frac{1+i \sqrt{3}}{6}\sqrt{\frac{\pi}{70}}\,{\hat f}^2
2035: +\frac{13(1+i \sqrt{3})}{792}\sqrt{\frac{5\pi}{14}}\,{\hat f}^4
2036: \cr
2037: &&
2038: a_{60}= \frac{1}{1848}\sqrt{\frac{\pi}{13}} \,{\hat f}^4 ,
2039: \qquad
2040: a_{62}= -\frac{(1 -i\,\sqrt{3} )}{4752}
2041: \sqrt{\frac{3\pi }{455}}\,{\hat f}^4 ,
2042: \qquad
2043: a_{64}= -\frac{1+i\sqrt{3}}{396} \sqrt{\frac{\pi }{182}}\,{\hat f}^4,
2044: \end{eqnarray}
2045: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2046: for self-correlated antenna pattern, $\mathcal{F}_{AA}$. Note that
2047: the multipole moments of ${\mathcal F}_{EE}$ are related to those of
2048: $\mathcal{F}_{AA}$ through the relations,
2049: $a_{\ell m}^{\scriptscriptstyle\rm EE}=
2050: a_{\ell m}^{\scriptscriptstyle\rm AA}$ for $m=0,\pm6,\pm12,\pm18,\cdots$ and
2051: $a_{\ell m}^{\scriptscriptstyle\rm EE}=
2052: -a_{\ell m}^{\scriptscriptstyle\rm AA}$ for $m=\pm2,\pm4,\pm 8,\cdots$
2053: (see Paper I). The multipole moments of cross-correlation signals are
2054: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2055: \begin{eqnarray}
2056: {\mathcal F}_{AE} :
2057: &&
2058: a_{22}= \frac{3+ i\sqrt{3}}{756} \sqrt{\frac{5\pi}{2}}\,{\hat f}^4,
2059: \quad
2060: a_{33}=\frac{1}{18} \sqrt{\frac{7\pi}{15}}\,{\hat f}^3 ,
2061: \quad
2062: a_{42}= - \frac{ 13(\sqrt{3}+i)}{5544} \sqrt{\frac{\pi }{10}}\,{\hat f}^4,
2063: \quad
2064: \cr
2065: &&
2066: a_{44}= \frac{\sqrt{3} -i}{6}\sqrt{\frac{\pi}{70}}\,\hat{f}^2
2067: -\frac{13(\sqrt{3} -i)}{792}\sqrt{\frac{5\pi}{14}}\,\hat{f}^4,
2068: \quad
2069: a_{53}= \frac{1}{18} \sqrt{\frac{\pi }{1155}}\,{\hat f}^3,
2070: \quad
2071: a_{62}= \frac{\sqrt{3} + i}{4752}\sqrt{ \frac{3\pi}{455} }\,{\hat f}^4,
2072: \cr
2073: &&
2074: a_{64}= \frac{ \sqrt{3}-i }{396} \sqrt{\frac{\pi }{182}}\,{\hat f}^4,
2075: \end{eqnarray}
2076: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2077:
2078: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2079: \begin{eqnarray}
2080: {\mathcal F}_{AT } :
2081: &&
2082: a_{11}= \frac{ ( 1 +i \sqrt{3} ) \sqrt{\pi}}{168}\,{\hat f}^3,
2083: \quad
2084: a_{22}= -\frac{\sqrt{3}+i}{864}\sqrt{\frac{3\pi }{5}}\,{\hat f}^4,
2085: \quad
2086: a_{31}= \frac{1+i\sqrt{3}}{72}{\sqrt{\frac{\pi }{14}}}\, {\hat f}^3 ,
2087: \quad
2088: \cr
2089: &&
2090: a_{42}= \frac{ \sqrt{3}+i}{3168}{\sqrt{\frac{\pi }{5}}}\,{\hat f}^4 ,
2091: \quad
2092: a_{44}= \frac{17(\sqrt{3}-i)}{3168}\sqrt{\frac{\pi}{35}}\,{\hat f}^4,
2093: \quad
2094: a_{51}= \frac{ 1 +i \sqrt{3}}{1008}{\sqrt{\frac{\pi }{55}}}\,{\hat f}^3 ,
2095: \cr
2096: &&
2097: a_{55}= \frac{1 -i\sqrt{3}}{72}{\sqrt{\frac{3\pi }{154}}}\,{\hat f}^3,
2098: \quad
2099: a_{62}= \frac{\sqrt{3}+i}{4752}{\sqrt{\frac{3\pi }{910}}}\,{\hat f}^4,
2100: \quad
2101: a_{64}= \frac{\sqrt{3} - i}{3168}{\sqrt{\frac{\pi }{91}}}\,{\hat f}^4.
2102: \end{eqnarray}
2103: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2104: The multipole moments of antenna pattern function $\mathcal{F}_{ET}$ are
2105: also obtained from those of $\mathcal{F}_{AT}$ using the relation,
2106: $ a_{\ell m}^{ET} = - (i/\sqrt{3})\, \tan \left(m\pi/3\right) a_{\ell m
2107: }^{AT}$, as shown in Paper I.
2108:
2109:
2110: Finally, we also list the multipole moments of self-correlated antenna
2111: pattern $\mathcal{F}_{TT}$, which are all higher-order contribution with
2112: $\mathcal{O}(\hat{f}^4)$:
2113: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2114: \begin{eqnarray}
2115: {\mathcal F}_{TT} :
2116: &&
2117: a_{00}= \frac{\sqrt{\pi} }{504}\,{\hat f}^4 ,
2118: \qquad
2119: a_{40}= - \frac{\sqrt{\pi} }{1584}\,{\hat f}^4 ,
2120: \qquad
2121: a_{60}= - \frac{1}{11088} \sqrt{\frac{\pi}{13}}\,{\hat f}^4,
2122: \qquad
2123: a_{66}= - \frac{1}{48} \sqrt{\frac{\pi }{3003}}\,{\hat f}^4 .
2124: \end{eqnarray}
2125: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2126: %
2127: %
2128: %
2129: %
2130: %
2131: %
2132: % ///////////////////////////////////////////////////////////// %
2133: \section{Remarks on the degeneracy between
2134: monopole and quadrupole moments for LISA measurement of
2135: GWB anisotropy}
2136: \label{appendix:on_the_degeneracy}
2137: % ///////////////////////////////////////////////////////////// %
2138: %
2139: %
2140: %
2141: %
2142: %
2143: %
2144: %
2145: %
2146: %
2147: The reconstruction method based on the harmonic-Fourier representation
2148: (\ref{eq:deconvolution}) presented in Sec.\ref{subsec:general_scheme}
2149: has been first discussed by Cornish
2150: \cite{Cornish:2001hg,Cornish:2002bh}. Later, Seto \& Cooray
2151: \cite{Seto:2004np} considered the reconstruction of a skymap in
2152: the low-frequency \textit{limit} using the optimal set of TDI
2153: signals $A$ and $E$. In this case, the accessible multipole moments
2154: $p_{\ell m}$ are $\ell=0,\,2$ and $4$ (see Table \ref{tab:summay_antenna}).
2155: Seto \& Cooray explicitly wrote down the expressions for linear equation
2156: (\ref{eq:deconvolution}) in the case of the self-correlation signals
2157: (i.e., AA-, EE-correlations) and showed that the output data with
2158: sufficient statistical significance are only $\widetilde{C}_0$,
2159: $\widetilde{C}_{\pm 1}$ and $\widetilde{C}_{\pm 2}$.
2160: Further, they found that the multipole coefficients
2161: $p_{00}$ and $p_{20}$ cannot be separately determined due to the
2162: degeneracy associated with a specific combination between the
2163: Wigner $D$ matrices and the antenna pattern functions.
2164:
2165:
2166:
2167: Here, we explicitly point out the origin of this degeneracy and
2168: discuss its influences on the reconstruction of GWB skymap.
2169: Using the multipole coefficients of antenna patterns in
2170: Appendix \ref{appendix:Mulipole coefficients}, the lowest-order contribution
2171: to the linear system
2172: (\ref{eq:deconvolution}) for AA- and EE-correlations becomes
2173: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2174: \begin{eqnarray}
2175: \widetilde{C}^{(2)}_{AA,0}
2176: &=&\frac{1}{4\sqrt{\pi}}\left\{
2177: \frac{2}{5}\,p_{00} - \frac{1}{14\sqrt{5}}\,p_{20}
2178: -\frac{37}{13440}\,p_{40}
2179: - \frac{27}{512\sqrt{70}}\sum_{m=-4,4}
2180: b_{4m}^{\scriptscriptstyle ({\mathrm{AA}})}
2181: \,p_{4m} \right\},
2182: \label{eq:degenerate_CAA0}
2183: \\
2184: \widetilde{C}^{(2)}_{EE,0}
2185: &=&\frac{1}{4\sqrt{\pi}}\left\{
2186: \frac{2}{5}\,p_{00} - \frac{1}{14\sqrt{5}}\,p_{20}
2187: -\frac{37}{13440}\,p_{40}
2188: + \frac{27}{512\sqrt{70}}\sum_{m=-4,4}
2189: b_{4m}^{\scriptscriptstyle ({\mathrm{AA}})}
2190: \,p_{4m} \right\}.
2191: \label{eq:degenerate_CEE0}
2192: \end{eqnarray}
2193: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2194: with $b^{\scriptscriptstyle ({\mathrm{AA}})}_{44}= 1-i\sqrt{3}=
2195: [b^{\scriptscriptstyle ( {\mathrm{AA}} )}_{4,-4}]^*$.
2196: Here, we only show the relevant components of $\widetilde{C}_k$ which
2197: contains the multipole coefficients $p_{00}$ and $p_{20}$.
2198: The above expressions include the multipole coefficients of
2199: $\ell=4$, which are all determined separately from the lowest-order
2200: contribution of $AE$-correlation, $\widetilde{C}_{AE,k}^{(2)}$.
2201: Thus, apart from the octupole moments, equations
2202: (\ref{eq:degenerate_CAA0}) and (\ref{eq:degenerate_CEE0}) clearly show
2203: the presence of degeneracy between the remaining multipole coefficients,
2204: $p_{00}$ and $p_{20}$.
2205:
2206:
2207: In a language of the least-squares method by SVD, this degeneracy
2208: implies that the sub-system in the matrix equation
2209: (\ref{eq:perturbative_c=Ap})
2210: containing the coefficients $p_{00}$ and $p_{20}$ becomes under-determined.
2211: In this case, the least-squares method by SVD cannot correctly produce
2212: the approximate solutions for $p_{00}$ and $p_{20}$, although it still
2213: provides some ``approximate'' solution.
2214: From equations (\ref{eq:degenerate_CAA0}) and
2215: (\ref{eq:degenerate_CEE0}),
2216: the only meaningful equation for $p_{00}$ and $p_{20}$ is now reduced to
2217: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2218: \begin{equation}
2219: \widetilde{C}=\frac{1}{4\sqrt{\pi}}\left(
2220: \frac{2}{5}\,p_{00} - \frac{1}{14\sqrt{5}}\,p_{20}\right),
2221: \end{equation}
2222: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2223: where the numerical constant $\widetilde{C}$ represents a collection of
2224: the
2225: irrelevant terms of $\ell=4$ modes, which are separately determined from
2226: the linear system in $AE$-correlation. If one naively applies the
2227: least-squares method to the above system, a very tight relation is obtained:
2228: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2229: \begin{eqnarray}
2230: p_{00} = \frac{7840\sqrt{\pi}}{789}\,\, \widetilde{C},
2231: \quad \quad
2232: p_{20} = \frac{280\sqrt{5\pi}}{789}\,\, \widetilde{C}.
2233: \label{eq:fake_solution}
2234: \end{eqnarray}
2235: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2236:
2237:
2238: The presence of degenerate coefficients may be a big obstacle in
2239: constructing the skymap as well as in determining the normalization
2240: factor of GWB spectrum. In principle, this degeneracy can be broken
2241: when we consider the higher-order terms of $\mathcal{O}(\hat{f}^4)$.
2242: However, these terms are generally small and irrelevant for the
2243: reconstruction analysis in the low-frequency regime.
2244: In this sense, the reconstruction of low-frequency skymap is, in nature,
2245: incomplete and the other additional information for monopole or
2246: quadrupole moment is required to make a full skymap. Nevertheless, it
2247: should be emphasized that with a high signal-to-noise ratio, the other
2248: remaining multipole coefficients can be all determined by the
2249: least-squares solution by SVD, independently of the above degeneracy.
2250: Moreover, in cases with $p_{00} \gg p_{20}$, which is usually
2251: satisfied, the least-squares solution (\ref{eq:fake_solution}) provides
2252: a modest estimate of the degenerate coefficients $p_{00}$ and $p_{20}$
2253: because of the hierarchy of the coefficients in Eq. (\ref{eq:fake_solution}).
2254: This point has been explicitly demonstrated in Sec.\ref{sec:demonstration}.
2255: %
2256: %
2257: %
2258:
2259: %
2260: %
2261: %
2262: %==============================================================%
2263: \section{Data sets and least-squares method}
2264: \label{appendix:SVD}
2265: %==============================================================%
2266:
2267: In this appendix, we describe in more details how to obtain the
2268: multipole moments of GWBs from many data streams by applying the
2269: least-squares method.
2270: Here we specifically focus on the situation considered in
2271: Sec. \ref{subsubsec:over-determined}. That is, the data sets that we use are
2272: (i) the $k=-2 \sim+2$ components of self-correlation signals $AA$ and $EE$,
2273: (ii) $k=-8\sim+8$ components of $AE$-signal, and (iii) the $k=-4\sim+4$
2274: components of cross-correlation signals $AT$ and $ET$. According to the
2275: general strategy given in Sec. \ref{subsec:general_scheme} (see Eq.
2276: (\ref{eq:perturbative_c=Ap})), we first combine the leading data streams
2277: of (i) and (ii) which correspond to $i=2$ in Eq. (\ref{eq:expand_vec_c}).
2278: The combined data sets consist of the following matrices.
2279: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2280: \begin{equation}
2281: %%%%%%%%%%%%%%%%%%
2282: \left(
2283: \begin{array}{c}
2284: \tilde{C}_{+2 ,AA}^{(2)} \\
2285: \tilde{C}_{+1 ,AA}^{(2)} \\
2286: \tilde{C}_{-1 ,AA}^{(2)} \\
2287: \tilde{C}_{-2 ,AA}^{(2)} \\
2288: \tilde{C}_{+2 ,EE}^{(2)} \\
2289: \vdots \\
2290: \tilde{C}_{-2 ,EE}^{(2)} \\
2291: \end{array}
2292: \right)
2293: %%%%%%%%%%%%%%%%%
2294: =
2295: \left(
2296: \begin{array}{ccccccccccccc}
2297: 0 &0 &0 & * & \\
2298: 0 &0 & * & 0 & \\
2299: 0 &* &0 & 0 & \\
2300: * &0 &0 & 0 & \\
2301: 0 &0 &0 & * & \\
2302: 0 &0 &* & 0 & \\
2303: 0 &* &0 & 0 & \\
2304: * &0 &0 & 0 & \\
2305: \end{array}
2306: \right)
2307: %%%%%%%%%%%%%%%%%
2308: \left(
2309: \begin{array}{c}
2310: p_{2 2} \\
2311: p_{2 1} \\
2312: p_{2,-1} \\
2313: p_{2,-2} \\
2314: \end{array}
2315: \right),
2316: %%%%%%%%%%%%%%%%%%
2317: \quad
2318: %%%%%%%%%%%%%%%%%%
2319: \left(
2320: \begin{array}{c}
2321: \tilde{C}_{+8 ,AE}^{(2)} \\
2322: \tilde{C}_{+7 ,AE}^{(2)} \\
2323: \vdots \\
2324: \tilde{C}_{0 ,AE}^{(2)} \\
2325: \vdots \\
2326: \tilde{C}_{-7 ,AE}^{(2)} \\
2327: \tilde{C}_{-8 ,AE}^{(2)} \\
2328: \end{array}
2329: \right)
2330: %%%%%%%%%%%%%%%%%
2331: =
2332: \left(
2333: \begin{array}{ccccccccccccc}
2334: 0 & && & * & \\
2335: & &&\Rdots& & \\
2336: &\Rdots&& & & \\
2337: * & && & * & \\
2338: & &&\Rdots& & \\
2339: &\Rdots&& & & \\
2340: * & && & 0 & \\
2341: \end{array}
2342: \right)
2343: %%%%%%%%%%%%%%%%%
2344: \left(
2345: \begin{array}{c}
2346: p_{4 4} \\
2347: p_{4 3} \\
2348: \vdots \\
2349: p_{4 0} \\
2350: \vdots \\
2351: p_{4,-3} \\
2352: p_{4,-4} \\
2353: \end{array}
2354: \right),
2355: %%%%%%%%%%%%%%%%%%
2356: \end{equation}
2357: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2358: where $*$ represents a non-vanishing complex number. Note that for
2359: simplicity we have not included the data $\tilde{C}_{0 ,AA}^{(2)}$
2360: and $\tilde{C}_{0 ,EE}^{(2)}$, which contain degeneracy between
2361: the multipole moments (see Sec. \ref{appendix:on_the_degeneracy}).
2362: The matrices in the right hand side correspond to $\mathbf{A}^{(i)}$
2363: in Eq. (\ref{eq:expand_vec_c}). These sparse forms of matrix are typical
2364: for the present problem.
2365: We perform the singular value decomposition with respect to these sparse
2366: matrices, and construct pseudo-inverse matrices which give the
2367: least-squares solutions of $p_{\ell m}$.
2368: In a similar way the least-squares solutions of $\ell=\mathrm{odd}$
2369: modes are obtained from the following matrix equation,
2370: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2371: \begin{equation}
2372: \left(
2373: \begin{array}{c}
2374: \tilde{C}_{+8 ,AE}^{(3)} \\
2375: \vdots \\
2376: \tilde{C}_{-8 ,AE}^{(3)} \\
2377: \tilde{C}_{4 ,AT}^{(3)} \\
2378: \vdots \\
2379: \tilde{C}_{-4 ,AT}^{(3)} \\
2380: \tilde{C}_{4 ,ET}^{(3)} \\
2381: \vdots \\
2382: \tilde{C}_{-4 ,ET}^{(3)} \\
2383: \end{array}
2384: \right)
2385: %%%%%%%%%%%%%%%%%
2386: =
2387: \left(
2388: \begin{array}{ccccccccccccc}
2389: & & & & \\
2390: & & \mathbf{A}_{\mathrm{AE}}^{(3)} & & \\
2391: & & & & \\
2392: & & \mathbf{A}_{\mathrm{AT}}^{(3)} & & \\
2393: & & & & \\
2394: & & \mathbf{A}_{\mathrm{ET}}^{(3)} & & \\
2395: & & & & \\
2396: \end{array}
2397: \right)
2398: %%%%%%%%%%%%%%%%%
2399: \left(
2400: \begin{array}{c}
2401: p_{11} \\
2402: p_{10} \\
2403: p_{1,-1} \\
2404: p_{33} \\
2405: \vdots \\
2406: p_{3,-3} \\
2407: p_{5 5} \\
2408: \vdots \\
2409: p_{5, -5} \\
2410: \end{array}
2411: \right),
2412: \end{equation}
2413: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2414: where we have symbolically written the matrix ${\mathbf{A}}^{(3)}$ due
2415: to space limitation.
2416: %
2417: %
2418: %
2419: %
2420: %
2421: %==============================================================%
2422: %==============================================================%
2423: \section{Multipole coefficients for gravitational-wave
2424: backgrounds}
2425: \label{appendix:multipole of GWB}
2426: %==============================================================%
2427: %==============================================================%
2428: %
2429: %
2430: %
2431: %
2432: Here, we give the multipole coefficients for normalized intensity
2433: distribution $P(\mathbf{\Omega})$ of galactic GWB presented in
2434: Sec.\ref{subsec:GWB_model}.
2435: To evaluate this, we first create the intensity map $P(\mathbf{\Omega})$
2436: with $129\times256$ regular grids of celestial sphere $(\theta,\,\phi)$.
2437: Then, the spherical harmonic expansion of the intensity map is
2438: numerically carried out using the SPHEREPACK 3.1 package
2439: \cite{Adams:2003}.
2440: Table \ref{tab:summay_multipole} summarizes the multipole coefficients
2441: $p_{\ell m}$ up to $\ell=5$, which are relevant to the analysis in
2442: Sec.\ref{sec:demonstration}. Note also that
2443: $p_{\ell,-m}=(-1)^m p_{\ell m}^*$.
2444:
2445:
2446: To characterize the contribution of the $\ell$-th moment to the galactic
2447: GWB, we introduce the angular power defined by
2448: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2449: \begin{eqnarray}
2450: \sigma_{\ell}=\left\{ \frac{1}{2\ell+1}\,\,
2451: \sum_{m=-\ell}^{\ell} |p_{\ell m}|^2 \right\}^{1/2},
2452: \end{eqnarray}
2453: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2454: which is rotationally invariant \cite{Kudoh:2004he}.
2455: Figure \ref{fig:sigma_gwb} shows the normalized angular power,
2456: $\sigma_{\ell}/\sigma_0$ up to $\ell=30$.
2457: The dominant contribution to the intensity of galactic GWB comes from
2458: the multipoles with $\ell\lesssim 4$, however, the asymptotic behavior
2459: at higher multipoles is very slow and can be fitted by
2460: $\sigma_{\ell}/\sigma_{0} \sim 1.85\, e^{-0.005\ell}/\ell$
2461: (dotted line in Fig.\ref{fig:sigma_gwb}), which turns out to be a good
2462: approximation even to much higher multipoles, $\ell\sim100$.
2463:
2464: %%%%%%%%%%%%%%%%%%%%%%%% Figure %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2465: \begin{figure}[ht]
2466: \begin{center}
2467: \includegraphics[width=8.3cm,clip]{figs/gwbdist/sigma_gwb.eps}
2468: \end{center}
2469: \caption{Normalized angular power of galactic GWB anisotropy,
2470: $\sigma_{\ell}/\sigma_0$ as a function of multipole $\ell$.
2471: \label{fig:sigma_gwb}}
2472: \end{figure}
2473: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2474:
2475:
2476: %%%%%%%%%%%%%%%%%%%%%%%%%%% TABLE %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2477: \begin{table}[t]
2478: %\begin{ruledtabular}
2479: \caption{
2480: Multipole coefficients for galactic GWB.
2481: $p_{\ell m}$ with $m<0$ is given by
2482: $p_{\ell m}= (-1)^m p^*_{\ell,-m}$.
2483: }
2484: \begin{tabular}{cc|rrr}
2485: $\ell$ & $m$ & Re[$p_{\ell m}$] & Im[$p_{\ell m}$] \\
2486: \hline\hline
2487: 0 & 0 & 0.282 & 0 \\
2488: \hline
2489: 1 & 0 & -0.0215 & 0 \\
2490: 1 & 1 & 0.00876& -0.156 \\
2491: \hline
2492: 2 & 0 & -0.0681 & 0 \\
2493: 2 & 1 & -0.0828 & 0.0216 \\
2494: 2 & 2 & -0.180 & -0.0124 \\
2495: \hline
2496: 3 & 0 & 0.0231 & 0 \\
2497: 3 & 1 & 0.00419 & 0.0648 \\
2498: 3 & 2 & 0.0212 & 0.0544 \\
2499: 3 & 3 & -0.0170 & 0.135 \\
2500: \hline
2501: 4 & 0 & 0.0120 & 0 \\
2502: 4 & 1 & -0.0205 & -0.0224 \\
2503: 4 & 2 & 0.0807 & -0.00274 \\
2504: 4 & 3 & 0.0936 & -0.0198 \\
2505: 4 & 4 & 0.139 & 0.0194 \\
2506: \hline
2507: 5 & 0 & -0.0172 & 0 \\
2508: 5 & 1 & 0.000230 & -0.0272 \\
2509: 5 & 2 & -0.0250 & -0.00640 \\
2510: 5 & 3 & -0.00336 & -0.0693 \\
2511: 5 & 4 & -0.0177 & -0.0719 \\
2512: 5 & 5 & 0.0222 & -0.1.13 \\
2513: \hline
2514: \hline
2515: \end{tabular}
2516: \label{tab:summay_multipole}
2517: %\end{ruledtabular}
2518: \end{table}
2519: %%%%%%%%%%%%%%%%%%%%%%%%%%% TABLE %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2520: %
2521: %
2522: %
2523: %
2524: %
2525: %
2526: %
2527: %
2528: %==============================================================%
2529: %==============================================================%
2530: \section{Estimation of statistical error
2531: in reconstruction analysis}
2532: \label{appendix:statistical_error}
2533: %==============================================================%
2534: %==============================================================%
2535:
2536:
2537: In the reconstruction analysis based on the harmonic-Fourier
2538: representation, the statistical errors plotted in
2539: Fig.\ref{fig:reconstructed_plm3} are roughly estimated as follows.
2540: In the presence of the noises, the least-squares solution given
2541: in equation (\ref{eq:perturbative_p=AC}) becomes
2542: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2543: \begin{equation}
2544: \mathbf{p}^{(i)}_{\rm approx} =\left[\mathbf{A}^{(i)}\right]^+ \cdot
2545: \left\{\mathbf{c}^{(i)}+\mathbf{s}_{\rm n}^{(i)}\right\},
2546: \end{equation}
2547: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2548: where the additional term $\mathbf{s}_{\rm n}^{(i)}$ represents the
2549: noise contributions to the $i$-th order coefficient of perturbative
2550: expansion for $\mathbf{c}(f)$.
2551: The root-mean-square amplitude of the error $\Delta\mathbf{p}^{(i)}$
2552: is then defined by taking the ensemble average of the random noises as
2553: $\Delta\mathbf{p}^{(i)}
2554: \equiv \left\langle|\mathbf{p}^{(i)}
2555: -\langle\mathbf{p}^{(i)}\rangle|^2\right\rangle$, which gives the
2556: errors in multipole coefficient $\Delta p_{\ell m}$.
2557: The $j$-th components of the vector $\Delta\mathbf{p}^{(i)}$ becomes
2558: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2559: \begin{equation}
2560: |\Delta\mathbf{p}^{(i)}_{{\rm approx},j}|^2 =
2561: \left[\mathbf{A}^{(i)}\right]^+_{jk} \left[\mathbf{A}^{(i)}\right]^{+*}_{jk} \,
2562: \langle |\mathbf{s}_{{\rm n},k}^{(i)}|^2\rangle.
2563: \end{equation}
2564: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2565: Here, the variance $\langle |\mathbf{s}_{\rm n,k}^{(i)}|^2\rangle$
2566: roughly corresponds to the quadrature of the vector $\mathbf{c}^{(i)}$
2567: divided by the signal-to-noise ratio:
2568: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2569: \begin{equation}
2570: \mathbf{S}_{{\rm n},k}^{(i)}= \left\{\alpha\,\,
2571: \frac{\displaystyle \mathbf{c}_k^{(i)}}
2572: {\displaystyle \mathbf{(s/r)}_k}\right\}^2.
2573: \label{eq:error}
2574: \end{equation}
2575: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2576: In the above expression, the vector $\mathbf{(s/r)}$ represents the SNR,
2577: each component of which is the quantity $(S/N)_k$ defined in
2578: (\ref{eq:def_SNR}) just for the same component of the vector
2579: $\mathbf{c}^{(i)}$.
2580: Notice that the factor $\alpha$ is multiplied in equation
2581: (\ref{eq:error}) by hand.
2582: The reason why we have introduced the factor $\alpha$ is as follows.
2583: First note that the quantity $(S/N)_k$ basically reflects the
2584: signal-to-noise ratio for the most dominant term in the perturbative
2585: expansion for the signal
2586: $\tilde{C}_k(f)$ (see Eq.(\ref{eq:expand_vec_c})).
2587: For the $AE$-correlation, $(S/N)_k$ gives the signal-to-noise
2588: ratio for second-order term, i.e., $\mathbf{c}^{(2)}$.
2589: For the $AT$-correlation, $(S/N)_k$ reflects the signal-to-noise ratio for
2590: $\mathbf{c}^{(3)}$.
2591: In our reconstruction analysis, the higher-order contributions to the
2592: $AE$-correlation, $\mathbf{c}^{(3)}$ is used to make the skymap
2593: with multipoles $\ell\leq5$, which can be estimated by analyzing the
2594: multi-frequency data. In general, the signal from higher-order
2595: contribution is weaker than the lowest-order term, leading to the
2596: calibration error. The significance of this error would be enhanced
2597: by the factor roughly proportional to $\hat{f}^{-(i-2)}$ for the $i$-th
2598: order terms of $AE$-correlation. Hence, in order to mimic this, the
2599: factor $\alpha$ is multiplied and set to $\alpha=\hat{f}^{-(i-2)}$.
2600: In the case examined in Fig.\ref{fig:reconstructed_plm3}, we adopt
2601: $\alpha=\hat{f}^{-1}=10$ for third-order cross-correlation data
2602: $\tilde{C}_{{\rm AE},k}^{(3)}$.
2603: Otherwise we set $\alpha=1$. As a result, statistical errors of $\ell=3$
2604: modes become larger than those in $\ell=$even modes
2605: (see Fig.\ref{fig:reconstructed_plm3}). Note that the multipole
2606: coefficients with $\ell=1$ in case A and with $\ell=5$ in both case
2607: are completely degenerate and cannot be recovered by reconstruction
2608: analysis. Hence, the statistical errors were not evaluated.
2609:
2610:
2611:
2612: %==============================================================%
2613: %==============================================================%
2614: \section{Computational method of multipole coefficients}
2615: %==============================================================%
2616: %==============================================================%
2617:
2618:
2619:
2620: In this Appendix, we give a brief description of the numerics of
2621: calculating the multipole coefficients using the SPHEREPACK 3.1
2622: package \cite{Adams:2003}.
2623: The traditional spherical harmonic transform of a scalar function is
2624: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2625: \begin{eqnarray}
2626: \Psi (\theta,\phi) = \sum_{\ell,m} p_{\ell m} Y_{\ell m} (\theta,\phi).
2627: \label{eq:Psi}
2628: \end{eqnarray}
2629: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2630: The spherical harmonics are given by equation
2631: (\ref{eq:harmonic_Y_lm}) and the associated Legendre polynomials are
2632: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2633: \begin{eqnarray}
2634: P^m_{\ell} (x) = \frac{(-1)^m}{2^\ell \ell!} (1-x^2)^{m/2} \frac{d^{m+\ell}}
2635: {dx^{m+\ell}}(x^2-1)^\ell.
2636: \label{eq:legendre}
2637: \end{eqnarray}
2638: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2639: On the other hand, numerical computation of spherical harmonic
2640: expansion is performed with the subroutine \verb|shaec| in
2641: SPHEREPACK 3.1 package, which directly gives the following expansion form:
2642: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2643: \begin{eqnarray}
2644: \Psi(\theta,\phi) &=& \sqrt{ \frac{\pi}{2} }
2645: \sum_{\ell=0}^\infty \sum_{m=0}^{\infty}{}'
2646: P^m_\ell (\cos\theta) \gamma_{\ell m}
2647: \left[ \alpha_{\ell m} \cos m \phi - \beta_{\ell m} \sin m \phi \right],
2648: \label{eq:spherepack expansion}
2649: \end{eqnarray}
2650: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2651: where the prime notation on the sum indicates that the first term
2652: corresponding to $m=0$ is multiplied by the factor $1/2$.
2653: To relate the coefficients $\alpha_{\ell m}$ and $\beta_{\ell m}$ with
2654: $p_{\ell m}$, the expansion (\ref{eq:Psi}) is compared with the
2655: expression (\ref{eq:spherepack expansion}), leading to
2656: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2657: \begin{eqnarray}
2658: p_{\ell m } &=& \sqrt{ \frac{\pi}{2} }
2659: (\alpha_{\ell m} + i \beta_{\ell m} ),
2660: \cr
2661: (-1)^m p_{\ell,- m} &=& \sqrt{ \frac{\pi}{2} } (\alpha_{\ell m} - i
2662: \beta_{\ell m}),
2663: \label{eq:alpha_ellm to p_ellm}
2664: \end{eqnarray}
2665: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2666: for $m >0$, and
2667: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2668: \begin{equation}
2669: p_{\ell 0} = \sqrt{ \frac{\pi}{2} } \alpha_{\ell 0},
2670: \label{eq:alpha_ellm to p_ellm2}
2671: \end{equation}
2672: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2673: for $m=0$. Hence, with a help of these expressions, one can read
2674: off the multipole coefficients $p_{\ell m}$ from the expansion
2675: formula (\ref{eq:spherepack expansion}).
2676: Notice that the associated Legendre polynomials used in the
2677: SPHEREPACK 3.1 package differ from (\ref{eq:legendre}) by a
2678: factor $(-1)^m$ so that one must multiply this factor to the
2679: transformation law (\ref{eq:alpha_ellm to p_ellm}) and
2680: (\ref{eq:alpha_ellm to p_ellm2}).
2681:
2682:
2683:
2684: %%*****************************************************************
2685: %%*****************************************************************
2686: %%*****************************************************************
2687: %
2688: %
2689: %
2690: %
2691: %
2692: %
2693: %\clearpage
2694: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2695: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2696: %%%%%%%%%%%%%%%%%%%%%%%% REFERENCES %%%%%%%%%%%%%%%%%%%%%%%%%
2697: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2698: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2699:
2700:
2701:
2702: % \bibliographystyle{plain} %% a standard style
2703: %\bibliographystyle{apsrev} %% APS RevTex Style
2704:
2705:
2706: %\bibliography{GWs}
2707: \input{GWs.tex}
2708:
2709:
2710: %==============================================================%
2711: \end{document}
2712: %==============================================================%
2713:
2714:
2715: