gr-qc0504112/ms.tex
1: \documentclass[preprint,showpacs,preprintnumbers,amsmath,amssymb,nofootinbib]{revtex4}
2: \usepackage{graphicx}
3: 
4: \usepackage{dcolumn}
5: \usepackage{bm}
6: 
7: \begin{document}
8: \title{The White Dwarf --- White Dwarf galactic background in the LISA
9:   data.}
10: 
11: \author{Jeffrey A. Edlund}
12: \email{Jeffrey.A.Edlund@jpl.nasa.gov}
13: \affiliation{Jet Propulsion Laboratory, California Institute of
14:              Technology, Pasadena, CA 91109}
15: 
16: \author{Massimo Tinto}
17: \email{Massimo.Tinto@jpl.nasa.gov}
18: \affiliation{Jet Propulsion Laboratory, California Institute of
19:              Technology, Pasadena, CA 91109}
20: 
21: \author{Andrzej Kr\'olak}
22: \email{krolan@aei.mpg.de} \altaffiliation[also at: ]{Institute of
23: Mathematics, Polish Academy of Sciences, Warsaw, Poland}
24: \affiliation{Max-Planck-Institute for Gravitational Physics,
25: Albert Einstein Institute, D-14476 Golm, Germany}
26: 
27: \author{Gijs Nelemans} 
28: \email{nelemans@astro.ru.nl}
29: \affiliation{Department of Astrophysics, IMAPP, Radboud University
30:   Nijmegen, The Netherlands}
31: 
32: \date{\today}
33: %
34: \begin{abstract}
35:   LISA (Laser Interferometer Space Antenna) is a proposed space
36:   mission, which will use coherent laser beams exchanged between three
37:   remote spacecraft to detect and study low-frequency cosmic
38:   gravitational radiation. In the low-part of its frequency band, the
39:   LISA strain sensitivity will be dominated by the incoherent
40:   superposition of hundreds of millions of gravitational wave signals
41:   radiated by inspiraling white-dwarf binaries present in our own
42:   galaxy. In order to estimate the magnitude of the LISA response to
43:   this background, we have simulated a synthesized population that
44:   recently appeared in the literature. Our approach relies on entirely
45:   analytic expressions of the LISA Time-Delay Interferometric
46:   responses to the gravitational radiation emitted by such systems,
47:   which allows us to implement a computationally efficient and
48:   accurate simulation of the background in the LISA data. We find the
49:   amplitude of the galactic white-dwarf binary background in the LISA
50:   data to be modulated in time, reaching a minimum equal to about
51:   twice that of the LISA noise for a period of about two months around
52:   the time when the Sun-LISA direction is roughly oriented towards the
53:   Autumn equinox.  This suggests that, during this time period, LISA
54:   could search for other gravitational wave signals incoming from
55:   directions that are away from the galactic plane.  Since the
56:   galactic white-dwarfs background will be observed by LISA not as a
57:   stationary but rather as a cyclostationary random process with a
58:   period of one year, we summarize the theory of cyclostationary
59:   random processes, present the corresponding generalized spectral
60:   method needed to characterize such process, and make a comparison
61:   between our analytic results and those obtained by applying our
62:   method to the simulated data. We find that, by measuring the
63:   generalized spectral components of the white-dwarf background, LISA
64:   will be able to infer properties of the distribution of the
65:   white-dwarfs binary systems present in our Galaxy.
66: \end{abstract}
67: 
68: \pacs{04.80.Nn, 95.55.Ym, 07.60.Ly}
69: \maketitle
70: 
71: \section{Introduction}
72: \label{intro}
73: 
74: The Laser Interferometric Space Antenna (LISA) is a space mission
75: jointly proposed to the National Aeronautics and Space Administration
76: (NASA) and the European Space Agency (ESA). Its aim is to detect and
77: study gravitational waves (GW) in the millihertz frequency band. It
78: will use coherent laser beams exchanged between three identical
79: spacecraft forming a giant (almost) equilateral triangle of side $5
80: \times 10^6$ kilometers. By monitoring the relative phase changes of
81: the light beams exchanged between the spacecraft, it will extract the
82: information about the gravitational waves it will observe at
83: unprecedented sensitivities  \cite{PPA98}.
84: 
85: The astrophysical sources that LISA is expected to observe within its
86: operational frequency band ($10^{-4} - 1$ Hz) include extra-galactic
87: super-massive black-hole coalescing binaries, stochastic gravitational
88: wave background from the early universe, and galactic and
89: extra-galactic coalescing binary systems containing white dwarfs and
90: neutron stars.
91: 
92: Recent surveys have uniquely identified twenty binary systems emitting
93: gravitational radiation within the LISA band, while population studies
94: have concluded that the large number of binaries present in our own
95: galaxy should produce a stochastic background that will lie
96: significantly above the LISA instrumental noise in the low-part of its
97: frequency band.  It has been shown in the literature (see \cite{NYP01}
98: for a recent study and \cite{HBW90,EIS87} for earlier investigations)
99: that these sources will be dominated by detached white-dwarf ---
100: white-dwarf (WD-WD) binaries, with $1.1 \ \times 10^8$ of such systems
101: in our Galaxy. The detached WD-WD binaries evolve by
102: gravitational-radiation reaction and the number of such sources
103: rapidly decreases with increasing orbital frequency. Although it is
104: expected that, above a certain frequency cut-off ($1 - 2 \ {\rm
105:   mHz}$), we will be able to resolve individual signals and remove
106: them from the LISA data, it is still not clear how to further improve
107: the LISA sensitivity to other gravitational wave signals in the region
108: of the frequency band below the WD-WD background frequency cut-off.
109: Although two promising data analysis procedures have been proposed
110: \cite{CL03,KT03} for attempting to subtract the galactic background,
111: considerable work still needs to be done to verify their
112: effectiveness. In this context, simulating the LISA response to the
113: WD-WD background will be particularly useful for verifying present and
114: future data analysis ``cleaning'' algorithms.  A realistic simulation
115: will also quantify the effects of the LISA motion around the Sun on
116: the overall amplitude and phase of the GW signal generated by the
117: background in the LISA data.  The directional properties of the LISA
118: response and its time dependence introduced by the motion of LISA
119: around the Sun, together with the non-isotropic and non-homogeneous
120: distribution of the WD-WD binary systems within the galactic disk as
121: seen by LISA, imply that the magnitude of the background observed by
122: LISA will not be a stationary random process. As a consequence of the
123: one-year periodicity of the LISA motion around the Sun, there exist
124: relatively long ($\approx 2$ months) stretches of data during which
125: the magnitude of the LISA response to the background will reach an
126: absolute minimum \cite{Seto04}. Our simulation shows this minimum to
127: be less than a factor of two larger than the level identified by the
128: LISA secondary noises, suggesting the possibility of performing
129: searches for gravitational radiation from other sources located in
130: regions of the sky that are away from the galactic plane.  The LISA
131: sensitivity to such signal in fact will be less limited by the WD-WD
132: background during these periods of observation.
133: 
134: This paper is organized as follows. In Section \ref{LISA_RES} we
135: provide the analytic expression of one of the LISA Time-Delay
136: Interferometric (TDI) responses to a signal radiated by a binary
137: system. Although all the TDI responses to binary signals were first
138: given in their closed analytic form in \cite{KTV04}, in what follows
139: we will focus our attention only on the unequal-arm Michelson
140: combination, $X$. In Section \ref{WD_POP} we give a summary of how the
141: WD-WD binary population was obtained, and a description of our
142: numerical simulation of the $X$ response to it.  In Section \ref{Simu}
143: we describe the numerical implementation of our simulation of the LISA
144: $X$ response to the WD-WD background, and summarize our results.  In
145: particular, in agreement with the results by Seto \cite{Seto04}, we
146: find the amplitude of the galactic WD-WD background in the LISA
147: $X$-combination to be modulated in time, reaching a minimum when the
148: Sun-LISA direction is roughly oriented towards the Autumn equinox.
149: Furthermore, we show the amplitude of the background at its minimum to
150: be a factor less than two larger than the level identified by the LISA
151: noise for a time period of about two months, suggesting that LISA
152: could search (during this time period) for other gravitational wave
153: signals incoming from regions of the sky that are away from the
154: galactic plane.
155: 
156: The time-dependence and periodicity of the magnitude of the WD-WD
157: galactic background in the LISA data implies that it is not a
158: stationary but rather a cyclostationary random process of period one
159: year.  After providing a brief summary of the theory of
160: cyclostationary random processes relevant to the LISA detection of the
161: WD-WD galactic background, we apply it to three years worth of
162: simulated LISA $X$ data. We find that, by measuring the generalized
163: spectral components of such cyclostationary random process, LISA will
164: be able to infer key-properties of the distribution of the WD-WD
165: binary systems present in our own Galaxy.
166: 
167: \section{The LISA response to signals from binary systems}
168: \label{LISA_RES}
169: 
170: The overall LISA geometry is shown in Figure (\ref{Fig1}).
171: \begin{figure}
172: \centering
173: \includegraphics[width=2.5in, angle=-90]{Figure1.ps}
174: \caption{Schematic LISA configuration. Each spacecraft is 
175:   equidistant from point $o$, with unit vectors $\hat p_i$ indicating
176:   directions to the three spacecraft. Unit vectors $\hat n_i$ point
177:   between spacecraft pairs with the indicated orientation. \label{Fig1}}
178: \end{figure}
179: There are six beams exchanged between the LISA spacecraft, 
180: with the six Doppler measurements $y_{ij}$ ($i,j = 1, 2, 3$) recorded
181: when each received beam is mixed with the laser light of the
182: receiving optical bench.  The frequency fluctuations from the six
183: lasers, which enter in each of the six Doppler measurements, need to
184: be suppressed to a level smaller than that identified by the secondary
185:  (proof mass and optical path) noises  \cite{TEA02} in order to detect
186: and study gravitational radiation at the predicted amplitudes. 
187: 
188: Since the LISA triangular array has systematic motions, the two
189: one-way light times between any spacecraft pair are not the same
190:  \cite{S04}.  Delay times for light travel between
191: the spacecraft must be accounted for depending on the sense of light
192: propagation along each link when combining these data as a consequence
193: of the rotation of the array.  Following  \cite{TEA04}, the arms are
194: labeled with single numbers given by the opposite spacecraft; e.g.,
195: arm $2$ (or $2^{'}$) is opposite spacecraft $2$, where primed delays are
196: used to distinguish light-times taken in the counter-clockwise sense
197: and unprimed delays for the clockwise light times (see Figure (\ref{Fig2})).
198: Also the following labeling convention of the Doppler data will be
199: used.  Explicitly: $y_{23}$ is the one-way Doppler shift measured at
200: spacecraft $3$, coming from spacecraft $2$, along arm $1$.  Similarly,
201: $y_{32}$ is the Doppler shift measured on arrival at spacecraft $2$
202: along arm $1'$ of a signal transmitted from spacecraft $3$.  Due to the
203: relative motion between spacecraft, $L_1 \neq L_1^{'}$ in general. As
204: in  \cite{ETA00,TEA02}, we denote six further data streams,
205: $z_{ij}$ ($i,j = 1, 2, 3$), as the intra-spacecraft metrology data
206: used to monitor the motion of the two optical benches and the relative
207: phase fluctuations of the two lasers on each of the three spacecraft.
208: \begin{figure}
209: \centering
210: \includegraphics[width=3.0 in, angle=0.0]{Figure2.eps}
211: \caption{Schematic diagram of LISA configurations involving
212:   six laser beams. Optical path delays taken in the counter-clockwise
213:   sense are denoted with a prime, while unprimed delays are in the
214:   clockwise sense. 
215: \label{Fig2}}
216: \end{figure}
217: The frequency fluctuations introduced by the lasers, by the optical
218: benches, by the proof masses, by the fiber optics, and by the
219: measurements themselves at the photo-detectors (i.e.\ the shot-noise
220: fluctuations) enter the Doppler observables $y_{ij}$, $z_{ij}$ with
221: specific time signatures; see Refs.  \cite{ETA00,TEA02} for a detailed
222: discussion.  The contribution $y^\mathrm{GW}_{ij}$ due to GW signals
223: was derived in Ref.  \cite{AET99} in the case of a stationary array,
224: and further extended to the realistic configuration \cite{KTV04} of
225: the LISA array orbiting around the Sun.
226: 
227: Let us consider for instance the ``second generation'' unequal-arm
228: Michelson TDI observables, ($X_1, X_2, X_3$). Their expressions, in
229: terms of the Doppler measurements $y_{ij}$, $z_{ij}$, are as follows \cite{STEA03}
230: \begin{eqnarray}
231: X_1 & = & [{(y_{31} + y_{13;2}) + (y_{21} + y_{12;3'})}_{;2'2}
232: + {(y_{21} + y_{12;3'})}_{;33'2'2}
233: + {(y_{31} + y_{13;2})}_{;33'33'2'2}]
234: \nonumber
235: \\
236: & & -
237: [(y_{21} + y_{12;3'})
238: + {(y_{31} + y_{13;2})}_{;33'}
239: + {(y_{31} + y_{13;2})}_{2'233'} +
240: {(y_{21} + y_{12;3'})}_{;2'22'233'}]
241: \nonumber
242: \\
243: & & +  \frac{1}{2} \ [(z_{21} - z_{31}) - {(z_{21} -
244: z_{31})}_{;33'} - {(z_{21} - z_{31})}_{;2'2}
245: + {(z_{21} - z_{31})}_{;33'33'2'2}
246: \nonumber
247: \\
248: & & + {(z_{21} - z_{31})}_{;2'22'233'}
249: - {(z_{21} - z_{31})}_{;2'233'33'2'2}] \ ,
250: \label{eq:X1}
251: \end{eqnarray}
252: with $X_2$, $X_3$ following from Eq. (\ref{eq:X1}) by permutations of
253: the spacecraft indices. The semicolon notation shown in equation
254: (\ref{eq:X1}) emphasizes that the operation of sequentially applying
255: two or more delays to a given measurement is non-commutative as
256: consequence of the time dependence of the light-times $L_i$ and
257: $L_i^{'}$ ($i = 1, 2, 3$), and a specific order has to be adopted to
258: adequately suppress the laser noises \cite{TEA04,CH03,STEA03}.
259: Specifically: $y_{ij;kl} \equiv y_{ij} (t - L_l (t) - L_k (t - L_l))
260: \ne y_{ij;lk}$ (units in which the speed of light $c=1$).
261: 
262: The expressions of the gravitational wave signal and the secondary
263: noise sources entering into $X_1$ will in general be different from
264: those entering into $X$, the corresponding ``first generation''
265: unequal-arm Michelson observable derived under the assumption of a
266: stationary LISA array \cite{AET99,ETA00}.  However, the magnitude
267: of the corrections introduced by the motion of the array are
268: proportional to the product between the time derivative of the GW
269: amplitude and the difference between the actual light travel times and
270: those valid for a stationary array.  At $1$ Hz, for instance, the
271: larger correction to the signal (due to the difference between the
272: co-rotating and counter-rotating light travel times) is two orders of
273: magnitude smaller than the main signal. Since the amplitude of this
274: correction scales linearly with the Fourier frequency, we can
275: completely disregard this effect (and the weaker effect due to the
276: time dependence of the light travel times) over the entire LISA band
277:  \cite{TEA04}.  Furthermore, since along the LISA orbit the three
278: armlengths will differ at most by $\sim$ 1\%--2\%, the degradation in
279: signal-to-noise ratio introduced by adopting signal templates that
280: neglect the inequality of the armlengths will be of only a few
281: percent.  For these reasons, in what follows we will focus on the
282: expressions of the GW responses of various second-generation Time-Delay
283: Interferometry (TDI) observables by disregarding the differences in
284: the delay times experienced by light propagating clockwise and
285: counterclockwise, and by assuming the three LISA armlengths to be
286: constant and equal to $L = 5 \times 10^6 \, \mbox{km} \simeq 16.67 \,
287: \mbox{s}$ \cite{PPA98}.  These approximations, together with the
288: treatment of the moving-LISA GW response discussed in \cite{KTV04} are
289: essentially equivalent to the \emph{rigid adiabatic approximation\/} of
290: Ref. \cite{RCP04}, and to the formalism of Ref. \cite{Seto04}.
291: 
292: These considerations imply that the second generation TDI expressions
293: for the gravitational wave signal and the secondary noises can be
294: expressed in terms of the corresponding first generation TDIs.  For
295: instance, the gravitational wave signal entering into the second
296: generation unequal-arm Michelson combination, $X^{\rm GW}_1$, can be
297: written in terms of the gravitational wave response of the
298: corresponding first generation unequal-arm Michelson combination,
299: $X^{\rm GW} (t)$, in the following manner \cite{TL04}
300: \begin{equation}
301: X^{\rm GW}_1 (t) = X^{\rm GW} (t) - X^{\rm GW} (t - 4L)
302: \label{X1fromX}
303: \end{equation}
304: Equation (\ref{X1fromX}) implies that any data analysis procedure and
305: algorithm that will be implemented for the second generation TDI
306: combinations can actually be derived by considering the corresponding
307: first generation TDI expressions. For this reason, from now on we
308: will focus our attention on the gravitational wave responses of the
309: first generation combinations.
310: 
311: The gravitational wave response $X^{\rm GW} (t)$ of the unequal-arm
312: Michelson TDI combination to a signal from a binary system has been
313: derived in \cite{KTV04}, and it can be written in the following form
314: %
315: \begin{equation}
316: X^\mathrm{GW} (t) = \Re \left[A(x, t) \ e^{-i \phi (t)}\right] \ ,
317: \label{X}
318: \end{equation}
319: where $x = \omega_s L$ ($\omega_s$ being the angular frequency of the
320: GW signal in the source reference frame), and the expressions for the
321: complex amplitude $A (x, t)$ and the real phase $\phi (t)$
322: are
323: \begin{eqnarray}
324: A(x, t) & = & 2 \, x \, \sin(x) \left\{ \left[ sinc[(1+c_2(t))\frac{x}{2}] \ e^{i x(\frac{3}{2} +
325:   d_2 (t))} 
326: + sinc[(1-c_2(t))\frac{x}{2}] \ e^{i x(\frac{5}{2} +  d_2 (t))}
327:   \right] \right. \ {\cal B}_2 (t) 
328: \nonumber
329: \\
330: & - & \left. \left[ sinc[(1-c_3(t))\frac{x}{2}] \ e^{i x(\frac{3}{2} +
331:   d_3 (t))} 
332: + sinc[(1+c_3(t))\frac{x}{2}] \ e^{i x(\frac{5}{2} +  d_3 (t))}
333:   \right] \ {\cal B}_3 (t) \right\} \ ,
334: \label{A}
335: \end{eqnarray}
336: \begin{equation}
337: \phi(t) = \omega_s t + \omega_s \ R \ \cos \beta \ \cos(\omega_s t +
338: \eta_0 - \lambda) \ .
339: \label{phi}
340: \end{equation}
341: In equation (\ref{phi}) $R$ is the distance of the guiding center of
342: the LISA array, $o$, from the Solar System Barycenter, ($\beta,
343: \lambda$) are the ecliptic latitude and longitude respectively of the
344: source location in the sky, $\Omega = 2 \pi/{\rm year}$, and
345: $\eta_0$ defines the position of the LISA guiding center in the
346: ecliptic plane at time $t = 0$. Note that the functions $c_k (t)$,
347: $d_k (t)$, and ${\cal B}_k (t)$ ($k = 2, 3$) do not depend on $x$. The
348: analytic expressions for $c_k (t)$, and $d_k (t)$ are the same as
349: those given in equations (46,47) of reference \cite{KTV04}, while the
350: functions ${\cal B}_k (t)$ ($k = 2, 3$) are equal to
351: \begin{equation}
352: {\cal B}_k (t) = (a^{(1)} + i \ a^{(3)}) \ u_k (t) + (a^{(2)} + i \ a^{(4)}) \ v_k
353: (t) \ .
354: \label{B}
355: \end{equation}
356: The coefficients ($a^{(1)}, a^{(2)}, a^{(3)}, a^{(4)}$) depend only on
357: the two independent amplitudes of the gravitational wave signal,
358: ($h_+$, $h_\times$), the polarization angle, $\psi$, and an arbitrary
359: phase, $\phi_0$, that the signal has at time $t = 0$. Their analytic
360: expressions are given in equations (41--44) of reference \cite{KTV04},
361: while the functions $u_k (t)$, and $v_k (t)$ ($k=2, 3$) are given in
362: equations (27,28) in the same reference.
363: 
364: Since most of the gravitational wave energy radiated by the galactic
365: WD-WD binaries will be present in the lower part of the LISA
366: sensitivity frequency band, say between $10^{-4} - 10^{-3}$ Hz, it is
367: useful to provide an expression for the Taylor expansion of the $X$
368: response in the long-wavelength limit (LWL), i.e. when the wavelength
369: of the gravitational wave signal is much larger than the LISA
370: armlength ($x << 1$). As it will be shown in the following sections,
371: the LWL expression will allow us to analytically describe the general
372: features of the white dwarfs background in the $X$-combination, and
373: derive computationally efficient algorithms for numerically simulating
374: the WD-WD background in the LISA data.
375: 
376: The nth-order truncation, $X^{\rm GW}_{(n)} (t)$, of the Taylor
377: expansion of $X^{\rm GW}(t)$ in power series of $x$ can be written in
378: the following form
379: \begin{equation}
380: X^{\rm GW}_{(n)} (t) = Re \sum_{k=0}^n A^{(k)} (t) \ x^{k+2} \ e^{-i \phi (t)}
381: \ ,
382: \label{Xn}
383: \end{equation}
384: where the first three functions of time $A^{(k)} (t), \ \ k \le 2$
385: are equal to
386: \begin{eqnarray}
387: A^{(0)}  & = &  4 \ [{\cal B}_2 - {\cal B}_3 ] \ ,
388: \nonumber
389: \\
390: A^{(1)}  & = & 4 i \ [(d_2+2) \ {\cal B}_2  - (d_3+2) {\cal B}_3 ] \ ,
391: \nonumber
392: \\
393: A^{(2)}  & = & [2{d_3}^2  + 8 d_3 + \frac{28}{3} + \frac{1}{6}
394: {c_3}^2] \ {\cal B}_3 - [2{d_2}^2  + 8 d_2 + \frac{28}{3} + \frac{1}{6}
395: {c_2}^2] \ {\cal B}_2 \ .
396: \label{An}
397: \end{eqnarray}
398: Note that the form we adopted for $X^{\rm GW} (t)$ (equation \ref{X})
399: makes the derivation of the functions $A^{(k)} (t)$ particularly easy
400: since the dependence on $x$ in $A (x,t)$ is now limited only to the
401: coefficients in front of the two functions ${\cal B}_2 (t)$ and ${\cal B}_3 (t)$
402: (see equation (\ref{A})). 
403: 
404: Although it is generally believed that the lowest order
405: long-wavelength expansion of the $X$ combination, $X^{\rm GW}_{(0)}$,
406: is sufficiently accurate in representing a gravitational wave signal
407: in the low-part of the LISA frequency band, there has not been in the
408: literature any quantitative analysis of the error introduced by
409: relying on such a zero-order approximation.  Since any TDI combination
410: will contain a linear superposition of tens of millions of signals, it
411: is crucial to estimate such an error as a function of the order of the
412: approximation, $n$. In order to determine how many terms we need to
413: use for a given signal angular frequency, $\omega_s$, we will rely on
414: the following `` matching function''
415: \begin{equation}
416: M(X^{\rm GW}, X^{\rm GW}_{(n)}) \equiv  \sqrt{\frac{\int_{0}^{T} {[X^{\rm GW} (t) - X^{\rm
417:     GW}_{(n)} (t)]}^2 dt}{\int_{0}^{T} {[X^{\rm GW} (t)]}^2 dt}} \ .
418: \label{M}
419: \end{equation}
420: Equation (\ref{M}) estimates the percent root-mean-squared error
421: implied by using the $n^{\rm th}$ order LWL approximation. In Figure
422: (\ref{error}) we plot $M$ as a function of the signal frequency,
423: $f_s$ ($=\omega_s/2\pi$), for $n=0, 1, 2$. At $5 \times 10^{-4}$ Hz, for
424: instance, the zero-order LWL approximation ($n = 0$) of the $X$
425: combination shows an r.m.s.\ deviation from the exact response equal
426: to about $10$ percent. As expected, this inaccuracy increases for
427: signals of higher frequencies, becoming equal to $40$ percent at $2
428: \times 10^{-3}$ Hz. With $n=1$ the accuracy improves showing that the
429: $X^{\rm GW}_{(1)}$ response deviates from the exact one with an r.m.s.
430: error smaller than $10$ percent in the frequency band ($10^{-4} - 2
431: \times 10^{-3}$) Hz. In our simulation we have actually implemented
432: the $n=2$ LWL expansion because it was possible and easy to do.
433: \begin{figure}
434: \includegraphics[width=5in]{Figure3.ps}
435: \caption{Plots of the percentage root-mean-squared errors, $
436:   M(X^{\rm GW}, X^{\rm GW}_{(n)})$, associated with the
437:   long-wavelength expansion index $n$, as functions of the
438:   gravitational wave frequency, $f_s$. The source location has been
439:   assumed to be in the center of our galaxy.}
440: \label{error}
441: \end{figure}
442: 
443: \section{White Dwarf binary population distribution}
444: \label{WD_POP}
445: 
446: The gravitational wave signal radiated by a WD-WD binary system
447: depends on eight parameters, ($\phi_o, \iota, \psi, D, \beta, \lambda,
448: {\cal M}_c, \omega_s$), which are the constant phase of the signal
449:  ($\phi_o$) at the starting time of the observation, the inclination
450: angle ($\iota$) of the angular momentum of the binary system relative
451: to the line of sight, the polarization angle ($\psi$) describing the
452: orientation of the wave polarization axes, the distance ($D$) to the
453: binary, the angles ($\lambda, \beta$) describing the location of the
454: source in the sky relative to the ecliptic plane, the chirp mass
455:  (${\cal M}_c$), and the angular frequency ($\omega_s$) in the source
456: reference frame respectively.  Since it can safely be assumed that the
457: chirp mass ${\cal M}_c$ and the angular frequency $\omega_s$ are
458: independent of the source location \cite{NYP01} and of the remaining
459: angular parameters $\phi_o, \iota, \psi$, and because there are no
460: physical arguments for preferred values of the constant phase $\phi_o$
461: and the orientation of the binary given by the angles $\iota$ and
462: $\psi$, it follows that the joint probability distribution, $P(\phi_o,
463: \iota, \psi, D, \beta, \lambda, {\cal M}_c, \omega_s)$, can be
464: rewritten in the following form
465: \begin{equation}
466: P(\phi_o, \iota, \psi, D, \beta, \lambda, {\cal M}_c, \omega_s) = 
467: P_1 (\phi_o) P_2 (\iota) P_3 (\psi) P_4 (D, \beta, \lambda) 
468: P_5({\cal M}_c, \omega_s) \ . 
469: \end{equation}
470: In the implementation of our simulation we have assumed the angles
471: $\phi_o$ and $\psi$ to be uniformly distributed in the interval $[0, 2
472: \pi)$, and $\cos\iota$ uniformly distributed in the interval $[-1,1]$.
473: We further assumed the binary systems to be randomly distributed in
474: the Galactic disc according to the following axially symmetric
475: distribution ${\cal P}_4 (R, z)$ (see \cite{NYP01} Eq. (5))
476: \begin{equation}
477: {\cal P}_4 (R,z) = \frac{e^{-R/H} \ sech^2(z/z_o)}{4 \pi z_o H^2}  \ ,
478: \label{eq:Galdis}
479: \end{equation}
480: where $(R,z)$ are cylindrical coordinates with origin at the galactic
481: center, $H = 2.5$ kpc, and $z_o = 200$ pc, and it is proportional to
482: $P_4 (D, \lambda, \beta)$ through the Jacobian of the coordinate
483: transformation.  Note that the position of the Sun in this coordinate
484: system is given by $R_{\odot} = 8.5 \ {\rm kpc}$ and $z_{\odot} = -30
485: \ {\rm pc}$. We then generate the positions of the sources from the
486: distribution given by Eq.  (\ref{eq:Galdis}) and map them to their
487: corresponding ecliptic coordinates $(D, \beta,\lambda)$.
488: 
489: The physical properties of the WD-WD population (${\cal M}_c \equiv
490: {(m_1 m_2)}^{3/ 5}/{(m_1 + m_2)}^{1/5}$, with $m_1$, $m_2$ being the
491: masses of the two stars, and $\omega_s = 2 \pi f_s = 4 \pi/{\rm orbital
492:   period}$) are taken from the binary population synthesis simulation
493: discussed in \cite{NYP04}. For details on this simulation we refer the
494: reader to \cite{NYP04}, and for earlier work to
495: \cite{HBW90,EIS87,HB90,HB00,BH97,NYP01}. The basic ingredient for
496: these simulations is an approximate binary evolution code. A
497: representation of the complete Galactic population of binaries is
498: produced by evolving a large (typically $10^6$) number of binaries
499: from their formation to the current time, where the distributions of
500: the masses and separations of the initial binaries are estimated from
501: the observed properties of local binaries.  This initial-to-final
502: parameter mapping is then convolved with an estimate of the binary
503: formation rate in the history of the Galaxy to obtain the total
504: Galactic population of binaries at the present time.  From these the
505: binaries of interest can then be selected.  In principle this
506: technique is very powerful, although the results can be limited by the
507: limited knowledge we have on many aspects of binary evolution. For
508: WD-WD binaries, the situation is better than for many other
509: populations, since the observed population of WD-WD binaries allows us
510: to gauge the models (e.g. \cite{NYP01}).
511: 
512: We also include the population of semi-detached WD-WD binaries
513: (usually referred to as AM CVn systems) that are discussed in detail
514: in \cite{NYP04}.  In these binaries one white dwarf transfers its
515: outer layers onto a companion white dwarf. Due to the redistribution
516: of mass in the system, the orbital period of these binaries increases
517: in time, even though the angular momentum of the binary orbit still
518: decreases due to gravitational wave losses. The formation of these
519: systems is very uncertain, mainly due to questions concerning the
520: stability of the mass transfer (e.g. \cite{MNS04})
521: 
522: From the models of the Galactic population of the detached WD-WD
523: binaries and AM CVn systems two dimensional histograms were created,
524: giving the expected number of both WD-WD binaries and AM CVn systems
525: currently present in the Galaxy as function of the $log$ of the GW
526: radiation frequency, $f_s (= \omega_s/2 \pi)$ and chirp mass,
527: $\mathcal{M}_c$.  In the case of the detached WD-WD binaries, the
528: ($\log f_s, \mathcal{M}_c)$) space was defined over the set
529: $\mathcal{M}_c \in (0, 1.5]$, $\log f_s \in [-6, -1]$, and contained
530: $30 \ \times \ 50$ grid points, while in the case of the AM CVn
531: systems the region is intrinsically smaller, $\mathcal{M}_c \in (0,
532: 1.2]$, $\log f_s \in [-4, -1.5]$, containing only $24 \ \times \ 25$
533: grid points.
534: 
535: Figure (\ref{fig:wdwd_nel1}) shows the distribution of the number of
536: detached WD-WD binaries as a function of the chirp mass and signal
537: frequency in the form of a contour plot. This distribution reaches its
538: maximum within the LISA frequency band when the chirp mass is equal to
539: $\simeq 0.25 \ {\cal M}_\odot$, and it monotonically decreases as a
540: function of the signal frequency.  The distribution of the number of
541: AM CVn systems has instead a rather different shape, as shown by the
542: contour plot given in Figure (\ref{fig:amcvn_nel1}). The region of the
543: ($\mathcal{M}_c, \log f_s$) space over which the distribution is
544: non-zero is equal to ($0, 0.07$) $\times$ ($-3.4, -2.2$), and it
545: reaches its maximum at the point ($0.03, -3.35$).
546: 
547: \begin{figure}
548: \includegraphics[width=10cm]{Figure4.eps}
549: \caption{The distribution of detached white-dwarf --- white-dwarf
550:   binaries in our galaxy as a function of the gravitational wave
551:   frequency, $f_s$, and chirp mass, ${\mathcal M}_c$.}
552: \label{fig:wdwd_nel1}
553: \end{figure}
554: \begin{figure}
555: \includegraphics[width=10cm]{Figure5.eps}
556: \caption{The distribution of AM CVn binary systems in our galaxy as a
557:     function of the gravitational wave frequency, $f_s$, and chirp
558:     mass, ${\mathcal M}_c$.}
559: \label{fig:amcvn_nel1}
560: \end{figure}
561: 
562: \section{Simulation of the background signal in the LISA data}
563: \label{Simu}
564: 
565: In order to simulate the LISA $X$ response to the population of WD-WD
566: binaries derived in Section \ref{WD_POP} one needs to coherently add
567: the LISA response to each individual signal. Although this could
568: naturally be done in the time domain, the actual CPU time required to
569: successfully perform such a simulation would be unacceptably long. The
570: generation in the time domain of one year of $X^{\rm GW} (t)$ response
571: to a single signal sampled at a rate of $16$ seconds would require
572: about $1$ second with an optimized C++ code running on a Pentium IV
573: $3.2$ GHz processor.  Since the number of signals from the background
574: is of the order $10^{8}$, it is clear that a different algorithm is
575: needed for simulating the background in the LISA data within a
576: reasonable amount of time. We were able to derive and implement
577: numerically an analytic formula of the Fourier transform of each
578: binary signal, which has allowed us to reduce the computational time
579: by almost a factor $100$.  Furthermore, we have run our code on the
580: Jet Propulsion Laboratory (JPL) supercomputer system, which includes
581: $64$ Intel Itanium2 processors each with a clock speed of 900 MHz.
582: 
583: \subsection{The Fourier transform of a binary signal}
584: \label{X_FOURIER}
585: 
586: The expression of the Fourier transform of the TDI response
587: $X^\mathrm{GW}$ to a single binary signal (Eqs. (\ref{X}, \ref{A},
588: \ref{phi}), cannot be written (to our knowledge) in closed analytic
589: form.  However, by using the LWL expansion of the $X^{\rm
590:   GW}$-response, it is possible to obtain a closed-form expression of
591: its Fourier transform.  Since the WD-WD binary background has a
592: natural frequency cut-off that is between $1$ and $2$ millihertz, the
593: LWL expansion of the $X^{\rm GW}$ response (Eq \ref{Xn}), truncated at
594: $n=2$, can be used for accurately representing the gravitational wave
595: response of each binary signal, as discussed in Section \ref{LISA_RES}.
596: 
597: In order to derive the Fourier transform of $X^{\rm GW}_{(2)} (t)$, we
598: use the following expansion of the function $e^{- i \phi (t)}$ in terms of
599: the Bessel functions of the first kind, $J_q$, \cite{AS72}
600: \begin{equation}
601: e^{-i \ \phi (t)} =  \sum_{q=-\infty}^{\infty} J_q (\omega_s R \cos\beta) \ e^{- i[\omega_s t \ + \ q \
602:   (\Omega t \ + \ \eta_0 \ - \ \lambda \ + \  \frac{\pi}{2})]} \ .
603: \label{Bessel}
604: \end{equation}
605: Since the Bessel functions $|J_q(\omega_s R \cos\beta)|$ are much
606: smaller than unity when $|q| >> |\omega_s R \cos\beta|$, the expansion
607: given by equation (\ref{Bessel}) can be truncated at a finite index
608: $Q$, providing an accurate numerical estimation of the function $e^{-i
609:   \ \phi (t)}$.  These considerations allow us to write the following
610: expression of the Fourier transform of $X^{\rm GW}_{(n)} (t)$
611: \begin{eqnarray} 
612: \widetilde{X}^{\rm GW}_{(n)} (\omega) & = & \mathcal{F} 
613: \left( \Re \sum_{k=0}^n \sum_{q= - Q}^{Q}  A^{(k)} (t) \ x^{k+2} \ J_q(\omega_s R \cos\beta) \ e^{- i[\omega_s t \ + \ q \
614:   (\Omega t \ + \ \eta_0 \ - \ \lambda \ + \  \frac{\pi}{2})]} \right) 
615: \nonumber
616: \\ 
617: & = &  \pi \sum_{q=- Q}^{Q}  J_q(\omega_s R \cos\beta) \sum_{k=0}^n x^{k+2} \left\{
618: \mathcal{F}\left(\Re [A^{(k)} (t)] \right) * \left[\delta(\omega +
619: \omega_s + q \Omega)  e^{i q (-\eta_0+\lambda-\pi/2)} \right. \right.
620: \nonumber
621: \\
622: && + \left. \left. \delta(-\omega+\omega_s + q \Omega) e^{i q(
623:       \eta_0- \lambda+\pi/2)} \right] +
624: i \mathcal{F} \left(\Im [A^{(k)} (t)] \right) * \left[\delta(\omega +
625: \omega_s + q \Omega) e^{i q (-\eta_0+\lambda-\pi/2)} \right. \right.
626: \nonumber
627: \\
628: && - \left. \left. \delta(-\omega+\omega_s + q \Omega) e^{i q( \eta_0-
629:     \lambda+\pi/2)} \right] \right\} \ ,
630: \label{XF}
631: \end{eqnarray}
632: where $\mathcal{F}$ is the Fourier transform operator, the symbol $*$
633: between two expressions means their convolution, $\omega$ is the
634: Fourier angular frequency, and $A^{(k)} (t)$ are defined in Eq. (\ref{Xn})
635: and given in Eq. (\ref{An}) with $k= 0, 1, 2$.
636: 
637: As an example application of this general formula for the Fourier
638: transform of the $X^{\rm GW}_{(n)} (t)$ response, let us apply it to
639: the lowest order LWL expansion ($n=0$)
640: \begin{eqnarray} 
641: \widetilde{X}^{\rm GW}_{(0)} (\omega) & = & 4 \pi \sum_{q = - Q}^{Q}
642: J_q(\omega_s R \cos\beta) x^{2} \left\{ 
643: \left[a_1 (\tilde{u}_2(\omega)-\tilde{u}_3(\omega)) + a_2
644:   (\tilde{v}_2(\omega)-\tilde{v}_3(\omega)) \right] \right.
645: \nonumber
646: \\
647: && \left. * \left[\delta(\omega + \omega_s + q \Omega) 
648: e^{i q (-\eta_0+\lambda-\pi/2)} + \delta(-\omega+\omega_s + q \Omega) e^{i q( \eta_0-
649:       \lambda+\pi/2)} \right] \right.
650: \nonumber
651: \\
652: && \left. + i \left[a_3 (\tilde{u}_2(\omega)-\tilde{u}_3(\omega)) + a_4
653:   (\tilde{v}_2(\omega)-\tilde{v}_3(\omega))\right] \right.
654: \nonumber
655: \\
656: && \left. * \left[\delta(\omega + \omega_s + q \Omega) e^{i q
657:     (-\eta_0+\lambda-\pi/2)} - \delta(-\omega+\omega_s + q \Omega) e^{i q( \eta_0-
658:   \lambda+\pi/2)} \right] \right\} \ .
659: \label{XF0}
660: \end{eqnarray}
661: Since the Fourier transforms of $u$ and $v$ are both linear
662: combinations of nine Dirac delta functions centered on the frequencies
663: $\pm l \ \Omega \ , \ l=0, 1, 2, 3, 4$ (see equations (27--30) in
664: reference \cite{KTV04} for the expressions of $u$ and $v$), it follows
665: that $\tilde{X}^{\rm GW}_{(0)} (\omega)$ is also a linear combination
666: of Dirac delta functions. In particular, in the limit of negligible
667: Doppler modulation, the resulting expression (\ref{XF0}) reduces as
668: expected to that of a purely amplitude modulated sinusoidal signal
669: with central frequency equal to $\omega_s$ and upper and lower
670: band-limits given by $\omega_s + 4 \Omega$ and $\omega_s - 4 \Omega$
671: respectively \cite{GP97}.
672: 
673: The actual expression of the Fourier transform we implemented in our
674: simulation of the WD-WD background used Eq. (\ref{XF}) with $n=2$, and
675: maximum value of the index of the Bessel expansion, $Q$, equal to $
676: |\omega_s R \cos \beta| + 20$ in order to make negligible the error
677: associated with the truncation of the expansion itself.
678: 
679: One extra mathematical detail that we need to include is that the
680: Fourier transform of $X^{\rm GW}_{(n)} (t)$ is performed over a finite
681: integration time, $T$, while the expression in Eq. (\ref{XF})
682: corresponds to an infinite-time Fourier transform. In order to account
683: for this discrepancy we convolved the analytic Fourier transform of
684: the signal given in equation (\ref{XF}) with the Fourier transform of
685: a window function with an integration time $T$.  To avoid leakage
686: introduced by using a simple rectangular window, we have used instead
687: the Nuttall's modified Blackman-Harris window \cite{NBH81}. Although
688: this window is characterized by having the main lobe of its Fourier
689: transform slightly wider than that of the rectangular window, the
690: maximum of its side-lobes are about four orders of magnitude lower
691: than those of the rectangular window, reducing leakage significantly.
692: The expression of its Fourier transform, ${\widetilde {W}} (\omega)$,
693: is equal to
694: \begin{eqnarray}
695: {\widetilde {W}} (\omega) & = & n_0
696: [Sinc(\omega T) + i Cosc(\omega T)]
697: \label{WF}
698: \\
699: & - & n_1 \ 
700: [Sinc(\omega T + 2 \pi) + Sinc(\omega T - 2 \pi) 
701: + i \ (Cosc(\omega T + 2 \pi) + Cosc(\omega T - 2 \pi))] 
702: \nonumber
703: \\
704: & - & n_2 \ [Sinc(\omega T + 4 \pi) + Sinc(\omega T - 4
705: \pi) 
706: + i \ (Cosc(\omega T + 4 \pi) + Cosc(\omega T - 4 \pi))]
707: \nonumber
708: \\
709: & - & n_3 \ [Sinc(\omega T + 6 \pi) + Sinc(\omega T - 6 \pi)
710: + i \ (Cosc(\omega T + 6 \pi) + Cosc(\omega T - 6 \pi))] \ ,
711: \nonumber
712: \end{eqnarray}
713: where the functions $Sinc (.)$ and $Cosc (.)$ are defined as follows
714: \begin{equation}
715: Sinc(.) \equiv \frac{\sin(.)}{.} \ \ , \ \ Cosc(.) \equiv
716: \frac{\cos(.) - 1}{.} \ ,
717: \end{equation}
718: and  the coefficients $n_r \ , \ r=0, 1, 2, 3$
719: have the following numerical values
720: \begin{equation}
721: n_0 = 0.3635819 \ \ , \ \ 
722: n_1 = 0.24458875 \ \ , \ \ 
723: n_2 = 0.06829975\ \ , \ \ 
724: n_3 = 0.00532055\ .
725: \end{equation}
726: 
727: \subsection{Generation of the signal parameters}
728: 
729: We used the distributions given in Section \ref{WD_POP} to randomly
730: generate the parameters $\phi_o$, $\iota$, $\psi$, $D$, $\beta,
731: \lambda$, while the values of the chirp mass, ${\cal M}_c$, and the
732: logarithm of the frequency of the signal, $\log(f_s)$, were obtained
733: by further processing the numeric distribution function (given in
734: Section \ref{WD_POP}) of the number of sources. To derive the
735: distribution function for the variables (${\cal M}_c, \log (f_s)$)
736: within each grid-rectangle of our numerical distribution we proceeded
737: in the following way \cite{ValPriv04}.  Let us consider the number of
738: sources $N (x_1, x_2)$ as a function of two coordinates ($x_1, x_2$)
739: of a point in the $({\cal M}_c, \log (f_s))$ plane within a specified
740: grid-rectangle of the numerical distribution.  This function can be
741: approximated there by the following quadratic polynomial
742: \begin{equation}
743: N (x_1,x_2)= n_{00} (1-x_1) (1-x_2) + n_{11} x_1 x_2 + 
744: n_{01} (1-x_1) x_2 + n_{10} x_1 (1-x_2) \ ,
745: \end{equation}
746: where $n_{00}$, $n_{11}$, $n_{01}, n_{10}$ are equal to the number of
747: signals at the ``corners'' of the considered grid-rectangle and are
748: obtained by interpolation; ($x_1, x_2$) are therefore two real numbers
749: defined in the range $[0, 1]$.
750: 
751: If we integrate out the $x_2$-dependence in $N(x_1,x_2)$ we obtain
752: \begin{equation}
753: N_{x_2} (x_1) = \int_0^1 N (x_1,x_2) dx_2 = 
754: \frac{(-n_{00}+n_{11}-n_{01}+n_{10}) x_1 + n_{00} + n_{01} }{2} \ ,
755: \end{equation}
756: which defines the total number of sources within that grid-rectangle
757: having chirp mass equal to $x_1$. In order to derive the probability
758: distribution function of $x_1$ within that grid-rectangle we can
759: define the following mapping between a uniformly distributed random
760: variable, say $z_1$, and the random variable $x_1$ 
761: \begin{equation}
762: z_1 =  \frac{\int_0^{x_1} N_{x_2}(x_1') dx_1'}{\int_0^1 N_{x_2}(x_1') dx_1'} = 
763: \frac{(-n_{00}+n_{11}-n_{01}+n_{10}) x_1^2 + 2 (n_{00} + n_{01}) x_1 }
764: {n_{00}+n_{11}+n_{01}+n_{10}} \ .
765: \end{equation}
766: By solving the above non-linear equation for
767: every uniformly sampled $z_1$ we obtain
768: \begin{equation}
769: x_1 = \frac{n_{00}+n_{01} - \sqrt{n_{00}^2+2 n_{00} n_{01}+n_{01}^2 + 
770: (- n_{00}^2 - 2 n_{00} n_{01} + n_{11}^2 + 2 n_{11} n_{10}-n_{01}^2+
771: n_{10}^2) z_1 }}{n_{00}-n_{11}+n_{01}-n_{10}} \ ,
772: \label{x1}
773: \end{equation}
774: where the branch ``-'' has been chosen such that $x_1$ remains in the range
775: $[0, 1]$.  If $n_{00}-n_{11}+n_{01}-n_{10} = 0$ then equation
776:  (\ref{x1}) is no longer valid, and we have instead $x_1 = z_1$. 
777: 
778: A similar procedure can be implemented for calculating $x_2$. By
779: integrating $N (x_1, x_2')$ with respect to $x_2'$ over the
780: range ($0, x_2$), we can establish the following relationship between
781: another uniformly distributed random variable, say $z_2$, and $x_2$
782: \begin{equation}
783: z_2 = \frac{\int_0^{x_2} N(x_1,x_2') dx_2'}{\int_0^{1} N(x_1,x_2') dx_2'} = 
784: \frac{[(n_{00}+n_{11}-n_{01}-n_{10}) x_1-n_{00}+n_{01}] x_2^2+
785: 2 [(-n_{00}+n_{10}) x_1+n_{00}] x_2}{(-n_{00}+ n_{11}- n_{01} + n_{10})
786:   x_1+n_{00}+n_{01}} \ .
787: \end{equation}
788: After some simple algebra we can finally solve for $x_2$ in terms of
789: $x_1$ (itself a function of the uniformly distributed random variable
790: $z_1$) and $z_2$
791: \begin{equation}
792: x_2 = \frac{n_{00}(x_1 -1) -n_{10} x_1 +
793: \sqrt{F(x_1) z_2 
794: +(n_{00}^2-2 n_{00} n_{10}+n_{10}^2) x_1^2+ 2 (-n_{00}^2+n_{00}
795: n_{10}) x_1+n_{00}^2}}
796: {(n_{11}-n_{01}-n_{10}+n_{00}) x_1 - n_{00}+n_{01}} \ ,
797: \end{equation}
798: where now we have chosen the ``+'' branch so $x_2 \in [0, 1]$ range,
799: and the function $F(x_1)$ is equal to
800: \begin{equation}
801: F(x_1) \equiv (n_{11}^2 + n_{01}^2 - n_{00}^2 - 2 n_{01} n_{11} + 2 n_{00}
802:   n_{10} - n_{10}^2) x_1^2 +2 (-n_{00} n_{10} +
803:   n_{00}^2-n_{01}^2+n_{01} n_{11}) x_1-n_{00}^2+n_{01}^2 \ .
804: \end{equation}
805: Note that the equation for $x_2$ above is no longer valid when
806: $(n_{11}-n_{01}-n_{10}+n_{00}) x_1 - n_{00}+n_{01} = 0$, in which case
807: $x_2 = z_2$.  Once $x_1$ and $x_2$ are calculated, they can be
808: converted into the physical parameters ${\cal M}_c$ and $\omega_s$
809: according to the following relationships
810: \begin{eqnarray}
811: {\cal M}_c & = & {{\cal M}_c}_{00} +  x_1 \ \Delta_{{\cal M}_c} \\
812: \omega_s & = & 2 \pi 10^{\log ({f_s}_{00}) + x_2 \ \Delta_{\log (f_s)}}
813: \end{eqnarray}
814: where (${{\cal M}_c}_{00}, \log ({f_s}_{00})$) are the coordinates of
815: the ``lower-left-hand-corner'' of the considered grid-rectangle, and
816: $\Delta_{{\cal M}_c}$, $\Delta_{\log ({f_s})}$ are the lengths of the
817: sides of the grid-rectangle.
818: 
819: \subsection{Results of the Numerical Simulation}
820: \label{NumSim}
821: 
822: The expression for the finite-time Fourier transform of each WD-WD
823: signal in the $\widetilde{X}^{\rm GW}$ response given in Section
824: \ref{X_FOURIER} allows us to coherently add in the Fourier domain all
825: the signals radiated by the WD-WD galactic binary population described
826: in Section \ref{WD_POP}.  After inverse Fourier transforming the
827: synthesized response and removing the window from it, we finally
828: obtain the time-domain representation of the background as it will be
829: seen in the LISA TDI combination $X$.  This is shown in Figure
830: (\ref{time_results}), where we plot three years worth of simulated
831: $X^{\rm GW} (t)$, and include the LISA noise \cite{PPA98}. The
832: one-year periodicity induced by the motion of LISA around the Sun is
833: clearly noticeable.  One other interesting feature shown by Figure
834: (\ref{time_results}) is that the amplitude response reaches absolute
835: minima when the Sun-LISA direction is roughly oriented towards the
836: Autumn equinox, while the absolute maxima take place when the Sun-LISA
837: direction is oriented roughly towards the Galactic center
838: \cite{Seto04}. This fact can easily be understood by looking at Figure
839: (\ref{Galaxy_Ecliptic}). Since the ecliptic plane is not parallel to
840: the galactic plane, and because our own solar system is about $8.5 \ 
841: {\rm kpc}$ away from the galactic center (where most of the of WD-WD
842: binaries are concentrated ) it follows that the LISA $X^{\rm GW}$
843: response does not have a six-months periodicity.
844: 
845: \begin{figure}
846: \includegraphics[width=6in]{Figure6.eps}
847: \caption{Two snapshots of LISA along its trajectory as recorded six
848:   months apart by an observer in the ecliptic plane. They correspond
849:   to the times when the LISA response to the background achieves a local
850:   maximum.  The magnitudes of these maxima are not equal due to the
851:   relative disposition of the ecliptic plane with respect to the
852:   galactic plane. At time $t_0$ the angle $\theta_0$ between the
853:   normal to the plane of LISA and a vector pointing to the galactic
854:   center is equal to $24.47^\circ$.  Six months later, when the LISA
855:   response is also at a maximum, the angle $\theta_1$ is equal to
856:   $35.53^\circ$ which results in a smaller maximum.}
857: \label{Galaxy_Ecliptic}
858: \end{figure}
859: 
860: Note also that, for a time period of about $2$ months, the absolute
861: minima reached by the amplitude of the LISA response to the WD-WD
862: background is only a factor less than $2$ larger than the level of the
863: instrumental noise.  This implies that during these observation times
864: LISA should be able to search for other sources of gravitational
865: radiation that are not located in the galactic plane. This might turn
866: out to be the easiest way to mitigate the detrimental effects of the
867: WD-WD background when searching for other sources of gravitational
868: radiation.  We will quantitatively analyze in a follow up work how to
869: take advantage of this observation in order to optimally search, during
870: these time periods, for sources that are off the galactic plane.
871: 
872: \begin{figure}
873: \includegraphics[width=6in]{Figure7.eps}
874: \caption{Three years of simulated TDI $X$ response to
875:   the WD-WD Galactic background signal. The time series of the LISA
876:   instrumental noise is displayed for comparison.}
877: \label{time_results}
878: \end{figure}
879: 
880: In Figure (\ref{Spectrum}) we plot, as functions of the Fourier
881: frequency, $f$, the windowed Fourier powers of both the signal and the
882: noise entering into the TDI $X$ combination.  Note that in the region
883: of the LISA band below $0.2$ millihertz the power of the WD-WD
884: background is smaller than that of the instrumental noise.
885: 
886: \begin{figure}
887: \includegraphics[width=6in]{Figure8.eps}
888: \caption{The amplitude of the Fourier transform of the WD-WD Galactic
889:   background gravitational wave signal and of the LISA instrumental
890:   noise entering into the TDI combination $X$.}
891: \label{Spectrum}
892: \end{figure}
893: 
894: \section{Cyclostationary processes}
895: \label{sec:cyclo}
896: 
897: The results of our simulation (Figure (\ref{time_results})) indicate
898: that the LISA $X^{\rm GW}$ response to the background can be regarded,
899: in a statistical sense, as a periodic function of time. This is
900: consequence of the deterministic (and periodic) motion of the LISA
901: array around the Sun. Since its autocorrelation function will also be
902: a periodic function of period one year, it follows that any LISA
903: response to the WD-WD background should no longer be treated as a
904: stationary random process but rather as a periodically correlated
905: random process. These kind of processes have been studied for many
906: years, and are usually referred to as cyclostationary random processes
907: (see \cite{H89} for a comprehensive overview of the subject and for more
908: references). In what follows we will briefly summarize the properties
909: of cyclostationary processes that are relevant to our problem.
910: 
911: A continuous stochastic process ${\cal X} (t)$ having finite second order
912: moments is said to be {\em cyclostationary\/} with period $T$ if the
913: following expectation values
914: \begin{eqnarray}
915: E[{\cal X} (t)] &=& m(t) = m(t + T) , \\
916: E[{\cal X} (t') {\cal X} (t)] &=& C(t',t) = C(t' + T,t + T)
917: \end{eqnarray}
918: are periodic functions of period $T$, for every $(t',t) \in {\bf R}
919: \times {\bf R}$. For simplicity from now on we will assume $m(t) = 0$.
920: 
921: If ${\cal X} (t)$ is cyclostationary, then the function $B(t,\tau) \equiv C(t
922: + \tau,t)$ for a given $\tau \in {\bf R}$ is periodic with period $T$,
923: and it can be represented by the following Fourier series
924: \begin{equation}
925: B(t,\tau) = \sum_{r = -\infty}^{\infty}B_r(\tau) e^{i
926: 2\pi\frac{r t}{T}} \ ,
927: \end{equation}
928: where the functions $B_r(\tau)$ are given by
929: \begin{equation}
930: B_r(\tau) = \frac{1}{T}\int^T_0B(t,\tau) e^{- i 2\pi r\frac{
931:     t}{T}} \ dt \ .
932: \label{eq:FB}
933: \end{equation}
934: The Fourier transforms $g_r(f)$ of $B_r(\tau)$ are the so called
935: ``cyclic spectra'' of the cyclostationary process ${\cal X} (t)$  \cite{H89}
936: \begin{equation}
937: g_r(f) = \int_{-\infty}^{\infty}B_r(\tau)e^{-i 2\pi f \tau } \ d
938: \tau \ .
939: \end{equation}
940: If a cyclostationary process is real, the following relationships
941: between the cyclic spectra hold
942: \begin{eqnarray}
943: B_{-r}(\tau) & = & B^*_r(\tau) \ ,
944: \label{eq:ksym1}
945: \\
946: g_{-r}(-f) & = &  g^*_r(f) \ ,
947: \label{eq:ksym2}
948: \end{eqnarray}
949: where the symbol $^*$ means complex conjugation.  This implies that,
950: for a real cyclostationary process, the cyclic spectra with $r \geq 0$
951: contain all the information needed to characterize the process itself.
952: 
953: The function $\sigma^2(\tau) \equiv B(0,\tau)$ is the variance of the
954: cyclostationary process ${\cal X} (t)$, and it can be written as a Fourier
955: decomposition as a consequence of Eq. (\ref{eq:FB})
956: \begin{equation}
957: \sigma^2(\tau) = \sum_{r=-\infty}^{\infty}H_r e^{ i
958: 2\pi\frac{r \tau}{T}},
959: \end{equation}
960: where $H_r \equiv B_r(0)$ are harmonics of the variance $\sigma^2$.
961: From Eq. (\ref{eq:ksym1}) it follows that $H_{-r} =
962: {H}^*_r$.
963: 
964: For a discrete, finite, real time series ${\cal X}_t$, $t =
965: 1,\ldots, {\cal N}$ we can estimate the cyclic spectra by generalizing
966: standard methods of spectrum estimation used with stationary
967: processes. Assuming again the mean value of the time series ${\cal
968:   X}_t$ to be zero, the cyclic autocorrelation sequences are defined
969: as
970: \begin{equation}
971: s_l^r = \frac{1}{\cal N}\sum_{t=1}^{{\cal N}-|l|}{\cal X}_t {\cal X}_{t+|l|}
972: e^{-\frac{i 2\pi r (t-1)}{T}} \ . 
973: \label{eq:cycorr}
974: \end{equation}
975: It has been shown \cite{H89} that the cyclic autocorrelations are
976: asymptotically (i.e.\ for $N \rightarrow \infty$) unbiased estimators
977: of the functions $B_r(\tau)$. The Fourier transforms of the cyclic
978: autocorrelation sequences $s_l^r$ are estimators of the cyclic spectra
979: $g_r(f)$. These estimators are asymptotically unbiased, and are called
980: ``inconsistent estimators'' of the cyclic spectra, i.e. their
981: variances do not tend to zero asymptotically.  In the case of Gaussian
982: processes \cite{H89} consistent estimators can be obtained by first
983: applying a lag window to the cyclic autocorrelation and then perform
984: a Fourier transform.  This procedure represents a generalization of
985: the well-known technique for estimating the spectra of stationary
986: random processes \cite{PW93}.
987: 
988: An alternative procedure for identifying consistent estimators of the
989: cyclic spectra is to first take the Fourier transform,
990: $\tilde{{\cal X}}(f)$, of the time series ${\cal X} (t)$
991: \begin{equation}
992: \tilde{{\cal X}}(f) = \sum_{t = 1}^{\cal N}  {\cal X}_t e^{-i 2\pi f (t - 1)}
993: \end{equation}
994: and then estimate the cyclic periodograms $g_r(f)$
995: \begin{equation}
996: g_r(f) = \frac{\tilde{{\cal X}}(f)\tilde{{\cal X}}^*(f - \frac{2\pi
997:     r}{T})}{\cal N} \ .
998: \end{equation}
999: By finally smoothing the cyclic periodograms, consistent estimators of
1000: the spectra $g_r(f)$ are then obtained.  The estimators of the
1001: harmonics $H_r$ of the variance $\sigma^2$ of a cyclostationary
1002: random process can be obtained by first forming a sample variance of
1003: the time series ${\cal X}_t$. The sample variance is obtained by
1004: dividing the time series ${\cal X}_t$ into contiguous segments of
1005: length $\tau_0$ such that $\tau_0$ is much smaller than the period $T$ of
1006: the cyclostationary process, and by calculating the variance
1007: $\sigma^2_I$ over each segment.  Estimators of the harmonics are
1008: obtained either by Fourier analyzing the series $\sigma^2_I$ or by
1009: making a least square fit to $\sigma^2_I$ with the appropriate number
1010: of harmonics. Note that the definitions of (i) zero order ($r = 0$)
1011: cyclic autocorrelation, (ii) periodogram, and (iii) zero order
1012: harmonic of the variance, coincide with those usually adopted for
1013: stationary random processes.  Thus, even though a cyclostationary time
1014: series is not stationary, the ordinary spectral analysis can be used
1015: for obtaining the zero order spectra.  Note, however, that
1016: cyclostationary random processes provide more spectral information
1017: about the time series they are associated with due to the existence of
1018: cyclic spectra with $r > 0$.
1019: 
1020: As an important and practical application, let us consider a time
1021: series $y_t$ consisting of the sum of a stationary random process,
1022: $n_t$, and a cyclostationary one ${\cal X}_t$ (i.e. $y_t = n_t + {\cal
1023:   X}_t$).  Let the variance of the stationary time series $n_t$ be
1024: $\nu^2$ and its spectral density be $\mathcal{E}(f)$.  It is
1025: easy to see that the resulting process is also cyclostationary. If the
1026: two processes are uncorrelated, then the zero order harmonic
1027: $\Sigma^2_0$ of the variance of the combined processes is equal to
1028: \begin{equation}
1029: \Sigma^2_0 = \nu^2 + \sigma^2_0 \ ,
1030: \end{equation}
1031: and the zero order spectrum, $G_0(f)$, of $y_t$ is
1032: \begin{equation}
1033: G_0(f) =  \mathcal{E}(f)  +  g_0(f) \ .
1034: \end{equation}
1035: The harmonics of the variance as well as the cyclic spectra of $y_t$
1036: with $r > 0$ coincide instead with those of ${\cal X}_t$.  In other words,
1037: the harmonics of the variance and the cyclic spectra of the process
1038: $y_t$ with $r > 0$ contain information only about the cyclostationary
1039: process ${\cal X}_t$, and are not ``contaminated'' by the stationary process
1040: $n_t$. 
1041: 
1042: \section{Analytic study of the background signal}
1043: \label{sec:cycloa}
1044: 
1045: In the case of the ensemble of $N$ WD-WD binaries, the total signal
1046: $s(t)$ is given by the following sum
1047: \begin{equation}
1048: s(t) = \sum^N_{i=1} X^{\rm GW} (t; {\bf \Lambda}_i) \ , 
1049: \end{equation}
1050: where ${\bf \Lambda}$ represents the set
1051: ($\phi_o,\iota,\psi,D,\beta,\lambda,{\cal M}_c, \omega_s$) of $8$
1052: parameters characterizing a GW signal. Since $N$ is large, we can
1053: expect the parameters of the signals to be randomly distributed and
1054: regard the signal $s(t)$ itself as a random process. Its mean, $m(t)$,
1055: and its autocorrelation function, $C(t',t)$ can then be calculated by
1056: assuming the probability distribution of the vector $\bf \Lambda$,
1057: $P({\bf \Lambda})$, to be the product of the five probability
1058: distributions, $P_1 (\phi_o)$, $ P_2 (\iota)$, $ P_3 (\psi)$, $ P_4
1059: (D, \beta, \lambda)$, and $ P_5({\cal M}_c, \omega_s)$ (as we did in
1060: our numerical simulation of the WD-WD background in Section
1061: \ref{WD_POP}).  By assuming the angles $\phi_o$ and $\psi$ to be
1062: uniformly distributed in the interval $[0, 2\pi)$, and $\cos\iota$ to
1063: be uniformly distributed in the interval $[-1,1]$, we can then perform
1064: the integrals over the angles $\phi_o$, $\psi$, and $\iota$
1065: analytically and obtain the following expressions
1066: \begin{eqnarray}
1067: m(t) & = & N \int_V  X^\mathrm{GW}(t) P({\bf \Lambda}) d {\bf \Lambda}
1068: \nonumber
1069: \\
1070: & = & \frac{N}{8 \pi^2}\int_0^{2\pi}d\phi_o \int_0^{2\pi}d\psi 
1071: \int_{-1}^1d\cos\iota
1072: \int_{V_4}\int_{V_5} X^\mathrm{GW}(t) P_4 P_5 dV_4 dV_5 = 0 \ , 
1073: \end{eqnarray}
1074: \begin{eqnarray}
1075: C(t',t) & = & N \int_V  X^\mathrm{GW}(t') X^\mathrm{GW}(t) 
1076:  P({\bf \Lambda}) d {\bf \Lambda}
1077: \nonumber
1078: \\
1079: & = & \frac{N}{16 \pi^2}\int_0^{2\pi}d\phi_o 
1080: \int_0^{2\pi}d\psi \int_{-1}^1d\cos\iota
1081: \int_{V_4}\int_{V_5} \Re [A(x, t') A^* (x, t) 
1082: e^{i[\phi(t) - \phi(t')]}] P_4 P_5 dV_4
1083: dV_5 \ .
1084: \end{eqnarray}
1085: Note that the mean value $m (t)$ is equal to $0$ as a consequence of
1086: averaging the antenna response over the polarization angle $\psi$.
1087: 
1088: In order to gain an analytic insight about the statistical properties
1089: of the autocorrelation function $C(t', t)$, in what follows we will
1090: adopt the zero-order long-wavelength approximation of the LISA
1091: response $X^{\rm GW} (t)$ obtained by fixing $n=0$ in (\ref{Xn}) and
1092: using the expression for the complex amplitude $A^{(0)}$ given in
1093: equation (\ref{An}).  After some long but straightforward algebra, the
1094: autocorrelation function, $C(t', t)$, can be written in the following
1095: form
1096: \begin{eqnarray}
1097: C(t',t) & = & \frac{16}{5} N \int_{V_4}\int_{V_5} x^4  h_o^2  
1098: [u_2 (t') u_2 (t) + v_2 (t') v_2 (t) + u_3 (t') u_3 (t) + v_3 (t') v_3
1099: (t) 
1100: \nonumber
1101: \\
1102: & - &
1103: u_2 (t') u_3 (t) - v_2 (t') v_3 (t) -
1104: u_3 (t') u_2 (t) - v_3 (t') v_2 (t)]
1105: \cos[\phi(t') - \phi(t)] P_4 P_5 \ dV_4 \ dV_5 \ ,
1106: \end{eqnarray}
1107: where $x = \omega_s L$, and $h_o = \frac{4 {{\cal M}_c}^{5/3}}{D}
1108: \left[\frac{\omega_s}{2}\right]^{2/3}$ (units in which the
1109: gravitational constant, $G$, and the speed of light, $c$, are equal to
1110: $1$).  For frequencies less than $1$ mHz the Doppler modulation in the
1111: phase $\phi(t)$ can be neglected making $\phi(t) \simeq \omega_s \,
1112: t$. If we now introduce a new time variable $\tau = t' - t$ and define
1113: $B(t,\tau) \equiv C(t + \tau,t)$, we have
1114: \begin{equation}
1115: B(t,\tau) = 
1116: \int_{0}^{\infty}
1117: {\cal P}(\omega_s) \cos(\omega_s \tau) \,d\omega_s \sum_{r = -8}^8 B_r(\tau)
1118: \ e^{i r\Omega t} \ ,
1119: \label{eq:cycor}
1120: \end{equation}
1121: where
1122: \begin{eqnarray}
1123: {\cal P}(\omega_s) &=& \frac{48 N}{5} \ {(\omega_s L)}^4 \omega_s^{4/3}
1124: \int_{\mathcal{M}_c} ({\sqrt{2} \mathcal{M}_c})^{10/3} P_5(\mathcal{M}_c,
1125: \omega_s) \ d\mathcal{M}_c \ ,
1126: \end{eqnarray}
1127: and
1128: \begin{equation}
1129: B_r(\tau) = \int_{V_4} b_r(\Omega\tau)
1130: \frac{P_4 (D,\beta,\lambda)}{D^2} \ dV_4 \ .
1131: \label{eq:autocyc} 
1132: \end{equation}
1133: The functions $b_r$ entering into equation (\ref{eq:autocyc})  are
1134: equal to
1135: \begin{eqnarray}
1136: b_0 &=& V_0^2 + U_0^2 +
1137: (V_1^2 + U_1^2)\cos(  \Omega \tau)
1138: + U_2^2 \cos(2 \Omega \tau)
1139: + (V_3^2 + U_3^2)\cos(3 \Omega \tau)
1140: \nonumber
1141: \\ 
1142: &&   + (V_4^2 + U_4^2)\cos(4 \Omega \tau)  +
1143: (V_0^2 - U_0^2)\cos(4 \gamma_0) \ ,
1144: \nonumber
1145: \\
1146: b_1 &=& e^{i (\Omega \tau/2 - \delta_0)}
1147: [(U_0 U_1 + V_0 V_1)\cos(\Omega \tau/2) +
1148:  U_1 U_2 \cos(3 \Omega \tau/2) +
1149: \nonumber
1150: \\ 
1151: && U_2 U_3 \cos(5 \Omega \tau/2)  +
1152: (U_3 U_4 + V_3 V_4)\cos(7 \Omega \tau/2)  +
1153: (-U_0 U_1 + V_0 V_1)\cos(\Omega \tau/2) e^{i 4 \gamma_0}]             
1154: \ ,
1155: \nonumber
1156: \\
1157: b_2 &=& e^{i (\Omega\tau - 2 \delta_0)}
1158: [U_0 U_2 \cos(\Omega \tau) +
1159: U_2 U_4 \cos(3 \Omega \tau) +
1160: (V_1 V_3 + U_1 U_3)\cos(2 \Omega \tau) +
1161: \nonumber
1162: \\ 
1163: && (- U_1^2/2 + V_1^2/2 - U_0 U_2 \cos(\Omega \tau)) e^{i 4 \gamma_0}]    
1164: \ ,
1165: \nonumber
1166: \\
1167: b_3 &=& e^{i 3 (\Omega \tau/2 - \delta_0)}
1168: [( U_0 U_3 + V_0 V_3)\cos(3 \Omega \tau/2) +
1169: ( U_1 U_4 + V_1 V_4)\cos(5\Omega \tau/2) + 
1170: \nonumber
1171: \\ 
1172: && ((-U_0 U_3 + V_0 V_3)\cos(3\Omega \tau/2)
1173: - U_1 U_2 \cos(    \Omega \tau/2)) e^{i 4 \gamma_0}]
1174: \ ,
1175: \nonumber
1176: \\
1177: b_4 &=& e^{i 2 (\Omega \tau - 2 \delta_0)}
1178: [(U_0 U_4 + V_0 V_4) \cos(2 \Omega \tau) +
1179: ((V_1 V_3 - U_1 U_3) \cos(\Omega\tau)  +                                   
1180: \nonumber
1181: \\ 
1182: && (V_0 V_4 - U_0 U_4) \cos(2 \Omega \tau)
1183: - U_2^2/2) e^{i 4 \gamma_0}] 
1184: \nonumber
1185: \ ,
1186: \\
1187: b_5 &=& e^{i 5 (\Omega\tau/2 - \delta_0) +  i 4 \gamma_0}
1188: [(V_1 V_4 - U_1 U_4)\cos(3 \Omega \tau/2) - U_2 U_3\cos(\Omega
1189: \tau/2)]    
1190: \ ,
1191: \nonumber
1192: \\
1193: b_6 &=& e^{i (3 \Omega\tau - 6 \delta_0 + 4 \gamma_0)}
1194: [-U_2 U_4 \cos(\Omega \tau) - U_3^2/2 + V_3^2/2]                           
1195: \ ,
1196: \nonumber
1197: \\
1198: b_7 &=& e^{i 7 (\Omega\tau/2 - \delta_0) + i 4 \gamma_0} \cos(\Omega \tau/2)
1199: [- U_3 U_4 + V_3 V_4]                                   
1200: \ ,
1201: \nonumber
1202: \\
1203: b_8 &=& e^{i 4 (\Omega\tau - 2 \delta_0 + \gamma_0)}
1204: [-U_4^2/2 + V_4^2/2] \ ,
1205: \label{eqs:cofbt}
1206: \end{eqnarray}
1207: where $\delta_0 = \lambda - \eta_0$, $\gamma_0 = \lambda - \eta_0 -
1208: \xi_0$, and the functions $U_i$, $V_i$ are give in equations (31-39)
1209: of \cite{KTV04}.  It is easy to see that the autocorrelation
1210: $B(t,\tau)$ is periodic in $t$ with period one year for a fixed
1211: $\tau$, making it a cyclostationary random process.  Note that, if the
1212: ecliptic longitude $\lambda$ is uniformly distributed, all the
1213: coefficients $b_r$ given in Eq.  (\ref{eqs:cofbt}) vanish for $r > 0$,
1214: and the random process $s(t)$ becomes stationary as the
1215: autocorrelation $C(t',t)$ now depends on the time difference $t'-t$.
1216: 
1217: The non-stationarity of the WD-WD background was first pointed out by
1218: Giampieri and Polnarev \cite{GP97} under the assumption of sources
1219: distributed anisotropically, and they also obtained the Fourier
1220: expansion of the sample variance and calculated the Fourier
1221: coefficient for simplified WD-WD binary distributions in the
1222: Galactic disc. What was however not realized in their work is that
1223: this non-stationary random process is actually cyclostationary, i.e.
1224: there exists cyclic spectra that can in principle allow us to infer
1225: more information about the WD-WD background than one could obtain
1226: by just estimating the zero-order spectrum.
1227: 
1228: If we now set $\tau = 0$ in Eq. \ref{eq:autocyc} we obtain the Fourier
1229: expansion of the variance $\sigma^2(t)$ of the cyclostationary
1230: process
1231: \begin{equation}
1232: \sigma^2(t) = B(t,0) =  \sum_{k = -8}^8 B_{k0} e^{i k\Omega t} \ ,
1233: \end{equation}
1234: where
1235: \begin{equation}
1236: \label{eq:sigk}
1237:  B_{k0} = {\cal P}_o \int_{V_5} b_{k0}
1238: \frac{P_4(D,\beta,\lambda)}{D^2} dV_4 \ ,
1239: \end{equation}
1240: with ${\cal P}_o = \frac{1}{2\pi}\int_0^{\infty}{\cal P}(\omega_s) \ d\omega_s$, and
1241: \begin{eqnarray}
1242: b_{00} &=&   U_0^2 + U_1^2 + U_2^2 + U_3^2 + U_4^2 + V_0^2 + V_1^2 +
1243: V_3^2 + V_4^2 + 
1244: \nonumber
1245: \\ 
1246: && (V_0^2 - U_0^2) \cos(4\gamma_0)  ,\\
1247: b_{10} &=&   e^{-i \delta_0} (U_0 U_1 + U_1 U_2 + U_2 U_3 + U_3 U_4 +
1248: V_0 V_1 + V_3 V_4 
1249: + 
1250: \nonumber
1251: \\ 
1252: && (-U_0 U_1 + V_0 V_1) e^{i 4 \gamma_0}) 
1253: \ ,
1254: \\
1255: b_{20} &=&   e^{-i 2 \delta_0} (U_0 U_2 + U_1 U_3 + U_2 U_4 + V_1 V_3
1256: +
1257:  \nonumber
1258: \\
1259: && (- U_1^2/2 + V_1^2/2 - U_0 U_2) e^{i 4 \gamma_0}) 
1260: \\
1261: b_{30} &=&   e^{-i 3 \delta_0} (U_0 U_3 + U_1 U_4 + V_0 V_3 + V_1 V_4 + \\ \nonumber
1262: && (-U_0 U_3 - U_1 U_2 + V_0 V_3) e^{i 4 \gamma_0}) 
1263: \, 
1264: \\
1265: b_{40} &=&   e^{-i 4 \delta_0} (U_0 U_4 + V_0 V_4 + \\ \nonumber
1266: && (- U_0 U_4 - U_1 U_3 + V_0 V_4 + V_1 V_3 - U_2^2/2) e^{i 4
1267:   \gamma_0})  
1268: \ ,
1269: \\
1270: b_{50} &=&  e^{i (4 \gamma_0 - 5 \delta_0)} (- U_1 U_4 - U_2 U_3 + V_1
1271: V_4)   
1272: \ ,
1273: \\
1274: b_{60} &=&  e^{i (4 \gamma_0 - 6 \delta_0)} (- U_2 U_4 - U_3^2/2 +
1275: V_3^2/2)  
1276: \ , 
1277: \\
1278: b_{70} &=&  e^{i (4 \gamma_0 - 7 \delta_0)} (- U_3 U_4 + V_3 V_4)  
1279: \ , 
1280: \\
1281: b_{80} &=&  e^{i (4 \gamma_0 - 8 \delta_0)} (- U_4^2/2 + V_4^2/2) \ .
1282: \end{eqnarray}
1283: 
1284: If we assume the function ${\cal P}(\omega_s)$ to change very little over a
1285: frequency bin, or equivalently choose $\tau$ to be such that
1286: $\Omega\tau \ll 1$, we can then approximate the functions $b_r$ with
1287: the functions $b_{r0}$.  Under this approximation the cyclic spectra
1288: of the process $s(t)$ can be shown to reduce to the following
1289: expression
1290: \begin{equation}
1291: g_r(\omega_s) = \frac{1}{2}{\cal P}(\omega_s)B_{0r} \ . 
1292: \label{eq:cycspeca}
1293: \end{equation}
1294: Thus under the above approximations the cyclic spectra are determined
1295: by one function of the Fourier frequency, and by the coefficients of
1296: the Fourier decomposition of the cyclic variance.  Note that this
1297: simplified representation of the cyclic spectra will not be valid if
1298: there are additional correlations between the parameters of the binary
1299: population. For example, if the chirp masses or the frequencies of the
1300: radiation emitted by the binaries are correlated with the positions of
1301: the binaries themselves in the Galactic disc, then the cyclic spectra
1302: will display a different frequency dependence from that implied by
1303: equation (\ref{eq:cycspeca}). In general we can expect the direct
1304: measurements of the cyclic spectra from the LISA data to allow us to
1305: infer properties of the distribution of the parameters characterizing
1306: the WD-WD population.  In other words, by analyzing the $17$ real and
1307: independent cyclic spectra we should be able to derive more
1308: information about the WD-WD binary population than we would have by
1309: simply looking at the ordinary spectrum.
1310: 
1311: \section{Data analysis of the background signal}
1312: \label{DATA_A}
1313: We have numerically implemented the methods outlined in Section
1314: \ref{sec:cyclo} and applied them to our simulated WD-WD background
1315: signal.  A comparison of the results of our simulation of the detached
1316: WD-WD background with the calculation of the background by Hils and
1317: Bender \cite{HB90,Lweb} is shown in Figure (\ref{fig:nel_ben_comp}).
1318: We find that the amplitude of the background from our simulation is a
1319: factor of more than $2$ smaller than that of Hils and Bender.
1320: The level of the WD-WD background is determined by the number of such
1321: systems in the Galaxy. We estimate that our number WD-WD binaries should be
1322: correct within a factor $5$ and thus the amplitude of the background should
1323: be right within a factor of $\sqrt{5}$.
1324: In Figure (\ref{fig:nel_ben_comp}) we have plotted the two backgrounds
1325: against the LISA spectral density and we have also included the LISA
1326: sensitivity curve. The latter is obtained by dividing the instrumental
1327: noise spectral density by the detector GW transfer function averaged
1328: over isotropically distributed and randomly polarized signals.  In the
1329: zero-order long wavelength approximation this averaged transfer
1330: function is equal to $\sqrt{3/20}$.
1331: 
1332: \begin{figure}
1333: \includegraphics[width=10cm]{Figure9.eps}
1334: \caption{Comparison of detached WD-WD background obtained from
1335:   binary population synthesis simulation ( \cite{NYP01,NYP04}) with
1336:   the WD-WD background calculated by Hils and Bender \cite{HB90}.  The
1337:   amplitude spectral density of the LISA instrumental noise and the
1338:   LISA sensitivity curve are drawn for comparison.  All spectral
1339:   densities are one-sided.}
1340: \label{fig:nel_ben_comp}
1341: \end{figure}
1342: 
1343: Our analysis was applied to $3$ years of LISA $X$ data consisting of a
1344: coherent superposition of signals emitted by detached WD-WD binaries,
1345: by semi-detached binaries (AM CVn systems), and of simulated
1346: instrumental noise. The noise was numerically generated by using the
1347: spectral density of the TDI $X$ observable given in \cite{ETA00}.  In
1348: addition a $1$ mHz low-pass filter was applied to our data set in
1349: order to focus our analyses to the frequency region in which the WD-WD
1350: stochastic background is expected to be dominant.
1351: 
1352: The results of the Fourier analysis of the sample variance of the
1353: background signal are shown in Figures (\ref{fig:wdwd_nel_var}) and
1354: (\ref{fig:wdwd_nel_har}). The top panel of Figure (\ref{fig:wdwd_nel_var})
1355: shows the sample variance of the simulated data for which the
1356: variances were estimated over a period of $1$ week; periodicity is
1357: clearly visible.  The bottom panel instead shows the Fourier analysis
1358: of the sample variance for which we have removed the mean from the
1359: sample variance time series. The vertical lines correspond to
1360: multiples of $1$ year; two harmonics can clearly be distinguished from
1361: noise.  The other peaks of the spectrum that fall roughly half way
1362: between the multiples of $1/{\rm year}$ frequency, are from the
1363: rectangular window inherent to the finite time series.
1364: \begin{figure}
1365: \includegraphics[width=10cm]{Figure10.eps}
1366: \caption{Top panel: The sample variance of the simulated 
1367:   WD-WD background observed by LISA. The data includes two populations
1368:   of WD-WD binaries, detached and semi-detached, which are added to
1369:   the LISA instrumental noise. The data is passed through a low-pass
1370:   filter with a cut-off frequency of $1$ mHz.  Bottom panel: Fourier
1371:   analysis of the sample variance.  Two harmonics are clearly resolved.
1372: \label{fig:wdwd_nel_var}}
1373: \end{figure}
1374: In Figure (\ref{fig:wdwd_nel_har}) we present the least square fit of
1375: $8$ harmonics to our $3$ years of simulated $X$ data. The number $8$
1376: comes from our theoretical predictions of the number of harmonics
1377: obtained in Section \ref{sec:cycloa} (see Eq. (\ref{eq:cycor})). We
1378: have calculated the magnitude of the harmonics and obtained the
1379: residuals. The results from the least square fit agree very well with
1380: those obtained via Fourier analysis (see also Figure
1381: (\ref{fig:har_cyclo_comp})). The magnitudes of the first and second
1382: harmonics resolved by Fourier analysis, for instance, agree with the
1383: corresponding least square fit estimates within a few hundredth of a
1384: percent.
1385: 
1386: It is useful to compare the results of our numerical analysis against
1387: the analytic calculations of Giampieri and Polnarev \cite{GP97}. Their
1388: analytic expressions for the harmonics of the variance of a background
1389: due to binary systems distributed in the galactic disc are given in
1390: Eq. (42) and shown in Figure (4) of \cite{GP97}.  Our estimation
1391: roughly matches theirs in that the 0th order harmonics is dominant and
1392: the first two harmonics have more power than the remaining ones. Our
1393: estimate of the power in the second harmonic, however, is larger than
1394: that in the first one, whereas they find the opposite.  We attribute
1395: this difference to their use of a Gaussian distribution of sources in
1396: the Galactic disc rather than the exponential that we adopted from
1397: \cite{NYP01}.  Comparison between these two results suggests that it
1398: should be possible to infer the distribution of WD-WD binaries in our
1399: Galaxy by properly analyzing the harmonics of the variance of the
1400: galactic background measured by LISA.  How this can be done will be
1401: the subject of a future work.
1402: 
1403: \begin{figure}
1404: \includegraphics[width=10cm]{Figure11.eps}
1405: \caption{Top panel: The sample variance of the WD-WD background data
1406:   and the least square fit to it using $8$ harmonics (small circles).
1407:   Middle panel: magnitude of the harmonics obtained from the least
1408:   square fit. Bottom panel: residual error between the fit and the
1409:   data.
1410: \label{fig:wdwd_nel_har}}
1411: \end{figure}
1412: In order to validate our simulation and data analysis method we have
1413: compared the results of our estimation of the power in the harmonics
1414: of the variance against the explicit analytic calculation. To estimate
1415: the powers we have used Eq. (\ref{eq:sigk}) and we have evaluated the
1416: integrals by numerical and Monte Carlo methods. In the numerical
1417: calculation of the harmonics we have limited our analysis to the
1418: population of detached WD-WD binaries. Thus in order to make the
1419: comparison meaningful we have performed Fourier analysis and least
1420: square fit of the time series consisting only of simulated detached
1421: WD-WD binaries (without semi-detached ones and LISA instrumental
1422: noise). The results of the comparison are given in Figure
1423: (\ref{fig:har_cyclo_comp}). We see that for the 0th order harmonic and
1424: the first two harmonics the agreement is very good.  For higher order
1425: harmonics there are large discrepancies between the numerical
1426: calculation and estimation by the least square fit, while by using the
1427: Fourier transform method, we cannot even resolve higher harmonics in
1428: our $3$-year data set. We conclude that only the two first harmonics
1429: can be extracted reliably from the data. We also observe a very good
1430: agreement between the Fourier and the least square method.
1431: \begin{figure}
1432: \includegraphics[width=10cm]{Figure12.eps}
1433: \caption{Comparison between the estimated power in the harmonics
1434:   obtained via (i) Fourier analysis, (ii) least square fit, and (iii)
1435:   numerical calculation based on Eq. \ref{eq:sigk}. The blue line is
1436:   the power spectrum of the variance of the data, red vertical lines
1437:   are obtained from least square fit and the black vertical lines are
1438:   from the numerical calculation.}
1439: \label{fig:har_cyclo_comp}
1440: \end{figure}
1441: As a next step in our analysis, we have estimated the cyclic spectra
1442: of the simulated WD-WD background signal. In Figure
1443: (\ref{fig:spec_cyclo}) we have shown the cyclic spectra estimated from
1444: the data.  We have also plotted the spectrum of the LISA instrumental
1445: noise and the main spectrum ($k = 0$) estimated from the simulation.
1446: We find that the main spectrum and two cyclic spectra for $k = 1$ and
1447: $k = 2$ have the largest magnitude and, over some frequency range,
1448: they lie above the LISA instrumental noise. The remaining spectra are
1449: an order of magnitude smaller and are very noisy. We also see that all
1450: the cyclic spectra have roughly the same slope. This is predicted by
1451: our analytic calculations in Section \ref{sec:cycloa} and it follows
1452: from the assumed independence between the location of the binaries in
1453: the Galaxy ($D, \lambda, \beta$) and their frequencies and chirp
1454: masses ($\omega_s, {\cal M}_c$). We also find the magnitude of the 2nd
1455: cyclic spectrum to be higher than the first, similarly to what we had
1456: for the harmonics of the variance. Note that we estimated the spectra from 
1457: the time series consisting of the WD-WD background added to 
1458: the LISA instrumental noise.
1459: \begin{figure}
1460: \includegraphics[width=10cm]{Figure13.eps}
1461: \caption{The main (k = 0) spectrum of the 
1462:   simulated WD-WD background signal (red), and the $8$ cyclic spectra
1463:   (magenta) estimated from the simulated data are shown. The spectral
1464:   density of the LISA instrumental noise (black) is shown for
1465:   reference.
1466: \label{fig:spec_cyclo}}
1467: \end{figure}
1468: Like the analysis we did for the variance, we have also compared the
1469: estimates of the cyclic spectra from our simulation against those
1470: obtained via numerical calculation of the equations derived in Section
1471: \ref{sec:cycloa}. The corresponding results are presented in Figures
1472: (\ref{fig:spec_comp}) and (\ref{fig:spec_cyclo_comp}), where it is shown
1473: that the agreement between the two is quite good.
1474: 
1475: \begin{figure}
1476: \includegraphics[width=10cm]{Figure14.eps}
1477: \caption{Estimated main ($k = 0$) spectrum of the WD-WD background
1478:   (red) against the calculated spectrum (black). The LISA spectral
1479:   density curve (blue) is shown for comparison. The $0$th order
1480:   spectrum contains the LISA instrumental and hence it differs from the
1481:   spectrum given in Figure (\ref{fig:nel_ben_comp}).
1482: \label{fig:spec_comp}}
1483: \end{figure}
1484: \begin{figure}
1485: \includegraphics[width=10cm]{Figure15.eps}
1486: \caption{Estimated cyclic spectra (black) against the calculated spectra (red).
1487:   The LISA spectral density curve (blue) is shown for comparison.
1488: \label{fig:spec_cyclo_comp}}
1489: \end{figure}
1490: 
1491: Our analysis has shown that the LISA data will allow us to compute
1492: $17$ independent cyclic spectra (the $8$ complex cyclic spectra
1493: $g_r(f) , r=1, 2, ...8$ and the real spectrum $g_0 (f)$) of the WD-WD
1494: galactic background, $5$ of which can be expected to be measured
1495: reliably.  We have also shown that by performing generalized spectral
1496: analysis of the LISA data we will be able to derive more information
1497: about the WD-WD binary population (properties of the distribution of
1498: its parameters) than we would have by only looking at the ordinary
1499: $g_0 (f)$ spectrum.
1500: 
1501: \section*{Acknowledgments}
1502: The supercomputers used in this investigation were provided by funding
1503: from the Jet Propulsion Laboratory Institutional Computing and
1504: Information Services, and the NASA Directorates of Aeronautics
1505: Research, Science, Exploration Systems, and Space Operations.  A.K.\ 
1506: acknowledges support from the National Research Council under the
1507: Resident Research Associateship program at the Jet Propulsion
1508: Laboratory. This work was supported in part by the Polish Science
1509: Committee Grant No. KBN 1 P03B 029 27.  This research was performed at
1510: the Jet Propulsion Laboratory, California Institute of Technology,
1511: under contract with the National Aeronautics and Space Administration.
1512: 
1513: \begin{thebibliography}{10}
1514: %
1515: \bibitem{PPA98} P. L. Bender, K. Danzmann, and the LISA Study Team,
1516: \emph{Laser Interferometer Space Antenna for the Detection of
1517:     Gravitational Waves, Pre-Phase A Report}, doc. MPQ 233
1518:   (Max-Planck-Instit\"ut f\"ur Quantenoptik, Garching, 1998).
1519: %
1520: \bibitem{NYP01}
1521: G.\ Nelemans, L.\ R.\ Yungelson, and S.\ F.\ Portegies-Zwart, {\it
1522:   A\&A}, {\bf 375}, 890 (2001).
1523: %
1524: \bibitem{HBW90} 
1525: D.\ Hils, P.\ L.\ Bender, and R.\ F.\ Webbink, {\it Ap. J.}, {\bf 360}, 75 (1900).
1526: %
1527: \bibitem{EIS87}
1528: Ch.\ Evans, Ickp Iben, and L. Smarr, {\it Ap. J.}, {\bf 323}, 129 (1987).
1529: 
1530: \bibitem{CL03} N.J. Cornish, and S.L. Larson, {\it Phys. Rev. D}, {\bf 67},
1531: 103001 (2003).
1532: %
1533: \bibitem{KT03} A. Krolak, and M. Tinto, In: {\it Proceedings of the
1534:   tenth Marcel Grossmann Meeting on General Relativity}, eds. M
1535:   Novello, S. Perez-Bergliaffa, and R. Ruffini, World Scientific,
1536:   Singapore (2005).
1537: %
1538: \bibitem{Seto04} N. Seto, {\it Phys. Rev. D}, {\bf 69}, 123005 (2004).
1539: %
1540: \bibitem{KTV04} A. Kr\'olak, M. Tinto, and M. Vallisneri, {\it Phys. Rev. D}, {\bf 70},
1541: 022003 (2004).
1542: %
1543: \bibitem{TEA02} M. Tinto, F.B. Estabrook, and J.W. Armstrong, {\it
1544:     Phys. Rev. D\/} {\bf 65}, 082003 (2002).
1545: %
1546: \bibitem{S04} D.A. Shaddock, {\it Phys. Rev. D}, {\bf 69 }, 022001 (2004).
1547: %
1548: \bibitem{TEA04} M. Tinto, F.B. Estabrook, and J.W. Armstrong, 
1549: {\it Phys. Rev. D}, {\bf 69}, 082001 (2004).
1550: %
1551: \bibitem{ETA00} F.B. Estabrook, M. Tinto, and J.W. Armstrong, 
1552: {\it Phys. Rev. D}, {\bf 62}, 042002 (2000).
1553: %
1554: \bibitem{AET99} J. W. Armstrong, F. B. Estabrook, and M. Tinto, {\it
1555:     Astrophys. J.}, {\bf 527}, 814 (1999).
1556: %
1557: \bibitem{CH03} N.J. Cornish, and R.W. Hellings, {\it Class. Quantum Grav.}, {\bf 20},
1558: s163 (2004).
1559: %
1560: \bibitem{STEA03} D.A. Shaddock, M. Tinto, F.B. Estabrook, and J.W.
1561:   Armstrong, {\it Phys. Rev. D}, {\bf 68}, 061303(R) (2003).
1562: %
1563: \bibitem{RCP04} L.J. Rubbo, N.J. Cornish, and O. Poujade, 
1564: {\it Phys. Rev. D}, {\bf 69}, 082003 (2004).
1565: %
1566: \bibitem{TL04} M. Tinto, and S.L. Larson, {\it Phys. Rev. D}, {\bf 70}, 062002 (2004).
1567: %
1568: \bibitem{NYP04} G.\ Nelemans, L.\ R.\ Yungelson, and S.\ F.\ 
1569:   Portegies-Zwart, {\it Mon. Not. Roy. Astron. Soc.}, {\bf 349}, 181
1570:   (2004).
1571: %
1572: \bibitem{HB90} D. Hils and P. L. Bender, {\it Ap. J.}, {\bf
1573:     360}, 75--94 (1990).
1574: %
1575: \bibitem{HB00} D. Hils and P. L. Bender, {\it Ap. J.}, {\bf
1576:     537}, 334--341 (2000).
1577: %
1578: \bibitem{BH97} P. L. Bender and D. Hils, {\it Class. Quantum Grav.},
1579:   {\bf 14}, 1439--1444 (1997).
1580: %
1581: \bibitem{MNS04} T.\ R.\ Marsh,  G.\ Nelemans, and D.\ Steeghs, 
1582: {\it Mon. Not. Roy. Astron. Soc.}, {\bf 350}, 113 (2004).
1583: %
1584: \bibitem{AS72} Abramowitz, M. and Stegun, I. A. (Eds.), 
1585: {\it Handbook of Mathematical Functions with Formulas, Graphs, and
1586:   Mathematical Tables}, 9th Edition, Dover, New York, pp. 355-389, (1972).
1587: %
1588: \bibitem{GP97} G. Giampieri and A. G. Polnarev, {\it Mon. Not. R.
1589:     Astron. Soc.}, {\bf 291}, 149--161 (1997).
1590: %
1591: \bibitem{NBH81} A.H. Nuttall, {\it IEEE Transactions on Acoustics,
1592:     Speech, and Signal Processing}, {\bf ASSP-29}, 84, (1981).
1593: %
1594: \bibitem{ValPriv04} Michele Vallisneri (private communication), (2004).
1595: %
1596: \bibitem{H89} H. L. Hurd, {\it IEEE Transactions on Information
1597:     Theory}, {\bf 35}, 350 (1989).
1598: %
1599: \bibitem{PW93} D. B. Percival and A. T. Walden, {\it Spectral Analysis
1600:     for Physical Applications}, Cambridge University Press (1993),
1601:   Chapter 6.
1602: %
1603: \bibitem{Lweb}
1604: Plots and data for Lisa sensitivity curve and WD-WD background due
1605: to Hils and Bender can be found at the following URL:
1606: ``http://www.srl.caltech.edu/~shane/sensitivity/MakeCurve.html''.
1607: \end{thebibliography}
1608: \end{document}
1609: 
1610: 
1611: