gr-qc0303013/ms.tex
1: \documentclass[preprint,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
2: \usepackage{graphicx}% Include figure files
3: \usepackage{dcolumn}% Align table columns on decimal point
4: \usepackage{bm}% bold math
5: 
6: \begin{document}
7: 
8: \title{Implementation of Time-Delay Interferometry for LISA}
9: 
10: \author{Massimo Tinto} 
11: \email{Massimo.Tinto@jpl.nasa.gov}
12: \altaffiliation [Also at: ]{Space Radiation Laboratory, California
13:   Institute of Technology, Pasadena, CA 91125} 
14: \affiliation{Jet Propulsion Laboratory, California 
15: Institute of Technology, Pasadena, CA 91109}
16: 
17: \author{Daniel A. Shaddock}
18: \email{Daniel.A.Shaddock@jpl.nasa.gov}
19: \affiliation{Jet Propulsion Laboratory, California 
20: Institute of Technology, Pasadena, CA 91109}
21: 
22: \author{Julien Sylvestre} 
23: \email{jsylvest@ligo.caltech.edu}
24: \altaffiliation [Also at: ]{LIGO Laboratory, California
25: Institute of Technology, Pasadena, CA 91125}
26: \affiliation{Jet Propulsion Laboratory, California 
27: Institute of Technology, Pasadena, CA 91109}
28: 
29: \author{J.W. Armstrong}
30: \email{John.W.Armstrong@jpl.nasa.gov}
31: \affiliation{Jet Propulsion Laboratory, California Institute of Technology,
32:  Pasadena, CA 91109}
33: 
34: \date{\today}
35: \vskip24pt
36: 
37: \begin{abstract}
38:   We discuss the baseline optical configuration for the Laser
39:   Interferometer Space Antenna (LISA) mission, in which the lasers are
40:   not free-running, but rather one of them is used as the main
41:   frequency reference generator (the {\it master}) and the remaining
42:   five as {\it slaves}, these being phase-locked to the master (the
43:   {\it master-slave configuration}).  Under the condition that the
44:   frequency fluctuations due to the optical transponders can be made
45:   negligible with respect to the secondary LISA noise sources (mainly
46:   proof-mass and shot noises), we show that the entire space of
47:   interferometric combinations LISA can generate when operated with
48:   six independent lasers (the {\it one-way method}) can also be
49:   constructed with the {\it master-slave} system design.  The
50:   corresponding hardware trade-off analysis for these two optical
51:   designs is presented, which indicates that the two sets of systems
52:   needed for implementing the {\it one-way method}, and the {\it
53:     master-slave configuration}, are essentially identical.  Either
54:   operational mode could therefore be implemented without major
55:   implications on the hardware configuration.
56:   
57:   We then derive the required accuracies of armlength knowledge, time
58:   synchronization of the onboard clocks, sampling times and
59:   time-shifts needed for effectively implementing Time-Delay
60:   Interferometry for LISA.  We find that an armlength accuracy of
61:   about $16$ meters, a synchronization accuracy of about $50$ ns, and
62:   the time jitter due to a presently existing space qualified clock
63:   will allow the suppression of the frequency fluctuations of the
64:   lasers below to the level identified by the secondary noise sources.
65:   A new procedure for sampling the data in such a way to avoid the
66:   problem of having time shifts that are not integer multiples of the
67:   sampling time is also introduced, addressing one of the concerns
68:   about the implementation of Time-Delay Interferometry.
69: \end{abstract}
70: 
71: \pacs{04.80.N, 95.55.Y, 07.60.L}
72: 
73: \maketitle
74: 
75: \section{Introduction}
76: 
77: LISA, the Laser Interferometer Space Antenna, is a three-spacecraft
78: deep space mission, jointly proposed to the National Aeronautics and
79: Space Administration (NASA) and the European Space Agency (ESA). The
80: LISA scientific objective is to detect and study low-frequency cosmic
81: gravitational radiation by observing phase differences of laser beams
82: interchanged between drag-free spacecraft \cite{1}.
83: 
84: Modeling each spacecraft with two optical benches, carrying
85: independent lasers, frequency generators (called Ultra Stable
86: Oscillators), beam splitters and photo receivers, the measured
87: eighteen time series of frequency shifts (six obtained from the six
88: one-way laser beams between spacecraft pairs, six from the beams
89: between the two optical benches on each of the three spacecraft, and
90: six more from modulation of the laser beams with USO data) were
91: previously analyzed by Tinto {\it et al.} \cite{2}.  There it was
92: shown that there exist several combinations of these eighteen
93: observables which exactly cancel the otherwise overwhelming phase noise
94: of the lasers, the phase fluctuations due to the non-inertial motions
95: of the six optical benches, and the phase fluctuations introduced by
96: the three Ultra Stable Oscillators into the heterodyned measurements,
97: while leaving effects due to passing gravitational waves.
98: 
99: The analysis presented in \cite{2} relied on the assumptions that (i)
100: the frequency offsets of any pair of independent lasers (assumed there
101: to be $ \approx \ 300 \ {\rm MHz}$) could be observed within the
102: detection bandwidths of the photo receivers where the one-way Doppler
103: measurements are performed, and (ii) the telemetry data rate needed by
104: two of the three spacecraft to transmit their measured one-way Doppler
105: data to the third spacecraft (where the interferometric combinations
106: are synthesized) is adequate.  Although the technology LISA will be
107: able to use should make possible the implementation of Time-Delay
108: Interferometry (TDI) as discussed in \cite{2}, the possibility of
109: optimizing the design of the optical layout, while at the same time
110: minimizing the number of Doppler data needed for constructing the
111: entire space of interferometric observables, was not analyzed there.
112: Here we extend those results to a different optical configuration, in
113: which one of the six lasers is the provider of the frequency reference
114: (albeit time-delayed) for the other five via phase-locking.  This {\it
115:   master-slave} optical design could provide potential advantages,
116: such as smaller frequency offsets between beams from pairs of
117: different lasers, hardware redundancy, reliability, and can result in
118: a smaller number of measured data. An outline of the paper is given
119: below.
120: 
121: In Section II we summarize TDI, the data processing technique needed
122: for removing the frequency fluctuations of the six lasers used by
123: LISA, and other noises.  In order to show that the entire space of
124: interferometric observables LISA can generate can also be
125: reconstructed by using a master-slave optical configuration, we
126: consider the simple case of spacecraft that are stationary with
127: respect to each other. After showing that the entire space of
128: interferometric observables can be obtained by properly combining four
129: generators, ($\alpha, \beta, \gamma, \zeta$), we then derive the
130: expressions for the four generators corresponding to the master-slave
131: optical configuration.  By imposing some of the one-way measurements
132: entering into ($\alpha, \beta, \gamma, \zeta$) to be zero (the so
133: called {\it locking conditions}), we show that the expressions for
134: these generators can be written in terms of the one-way and two-way
135: Doppler measurements corresponding to the locking configuration we
136: analyzed. Section III and Appendix A provide a theoretical derivation
137: and estimation of the magnitude of the phase noise expected to be
138: generated by an optical transponder. In Section IV we analyze and
139: compare the hardware requirements needed for implementing both optical
140: designs, while in Section V we turn to the estimation of the armlength
141: and time synchronization accuracies, as well as time-shift and
142: sampling time precisions needed for successfully implementing TDI with
143: LISA.  Our comments and conclusions are finally presented in Section
144: VI.
145: 
146: \section{Time-Delay Interferometry}
147: 
148: The description of TDI for LISA is greatly simplified if we adopt the
149: notation shown in Figure 1, where the overall geometry of the LISA
150: detector is defined. The spacecraft are labeled $1$, $2$, $3$ and
151: distances between pairs of spacecraft are $L_1$, $L_2$, $L_3$, with
152: $L_i$ being opposite spacecraft $i$.  Unit vectors between spacecraft
153: are $\hat n_i$, oriented as indicated in figure 1.  We similarly index
154: the phase difference data to be analyzed: $s_{31}$ is the phase
155: difference time series measured at reception at spacecraft 1 with
156: transmission from spacecraft 2 (along $L_3$).
157: \begin{figure}
158: \centering
159: \includegraphics[width=2.5in, angle=-90]{Fig1.ps}
160: \caption{Schematic LISA configuration.  Each spacecraft is equidistant
161: from the point O, in the plane of the spacecraft.  Unit vectors
162: $\hat n_i$ point between spacecraft pairs with the indicated
163: orientation.  At each vertex spacecraft there are two optical
164: benches (denoted 1, $1^*$, etc.), as indicated.}
165: \end{figure}
166: Similarly, $s_{21}$ is the phase difference series derived from
167: reception at spacecraft $1$ with transmission from spacecraft $3$. The
168: other four one-way phase difference time series from signals exchanged
169: between the spacecraft are obtained by cyclic permutation of the
170: indices: $1$ $\to$ $2$ $\to$ $3$ $\to$ $1$.  We also adopt a useful
171: notation for delayed data streams: $s_{31,2} = s_{31}(t - L_2)$,
172: $s_{31,23} = s_{31}(t - L_2 - L_3) = s_{31,32}$, etc.  (we take the
173: speed of light $c = 1$ for the analysis).  Six more phase difference
174: series result from laser beams exchanged between adjacent optical
175: benches within each spacecraft; these are similarly indexed as
176: $\tau_{ij}$ ($i,j = 1, 2, 3 \ ; \ i \ne j$).
177: 
178: The proof-mass-plus-optical-bench assemblies for LISA spacecraft
179: number $1$ are shown schematically in figure 2.  We take the left-hand
180: optical bench to be bench number $1$, while the right-hand bench is
181: $1^*$.  The photo receivers that generate the data $s_{21}$, $s_{31}$,
182: $\tau_{21}$, and $\tau_{31}$ at spacecraft $1$ are shown.  The phase
183: fluctuations of the laser on optical bench $1$ is $p_1(t)$; on optical
184: bench $1^*$ it is $p^*_1(t)$ and these are independent (the lasers are
185: for the moment not "locked" to each other, and both are referenced to
186: their own independent frequency stabilizing device).  We extend the cyclic
187: terminology so that at vertex $i$ ($i = 1, 2, 3$) the random displacement
188: vectors of the two proof masses are respectively denoted $\vec
189: \delta_i(t)$ and $\vec \delta^*_i(t)$, and the random displacements
190: (perhaps several orders of magnitude greater) of their optical benches
191: are correspondingly denoted $\vec \Delta_i(t)$ and $\vec
192: \Delta^*_i(t)$.  As pointed out in \cite{3}, the analysis does
193: \underline {not} assume that pairs of optical benches are rigidly
194: connected, i.e. $\vec \Delta_i \neq \vec \Delta^*_i$, in general.  The
195: present LISA design shows optical fibers transmitting signals both
196: ways between adjacent benches.  We ignore time-delay effects for these
197: signals and will simply denote by $\mu_i(t)$ the phase fluctuations
198: upon transmission through the fibers of the laser beams with
199: frequencies $\nu_{i}$, and $\nu^*_{i}$.  The $\mu_i (t)$ phase shifts
200: within a given spacecraft might not be the same for large frequency
201: differences $\nu_{i} - \nu^*_{i}$. For the envisioned frequency
202: differences (a few hundred megahertz), however, the remaining
203: fluctuations due to the optical fiber can be neglected \cite{4}.  It
204: is also assumed that the phase noise added by the fibers is
205: independent of the direction of light propagation through them.
206: 
207: Figure 2 endeavors to make the detailed light paths for these
208: observations clear.  An outgoing light beam transmitted to a distant
209: spacecraft is routed from the laser on the local optical bench using
210: mirrors and beam splitters; this beam does not interact with the local
211: proof mass.  Conversely, an {\it {incoming}} light beam from a distant
212: spacecraft is bounced off the local proof mass before being reflected
213: onto the photo receiver where it is mixed with light from the laser on
214: that same optical bench. The inter-spacecraft phase data are denoted
215: $s_{31}$ and $s_{21}$ in figure 2.
216: \begin{figure}
217: \centering
218: \includegraphics[width=6.0 in, angle=0]{Fig2.ps}
219: \caption{Schematic diagram of proof-masses-plus-optical-benches
220:   for a LISA spacecraft.  The left-hand bench, $1$, reads out the
221:   phase signals $s_{31}$ and $\tau_{31}$.  The right hand bench,
222:   $1^*$, analogously reads out $s_{21}$ and $\tau_{21}$.  The random
223:   displacements of the two proof masses and two optical benches are
224:   indicated (lower case $\vec \delta_i$ for the proof masses, upper
225:   case $\vec \Delta_i$ for the optical benches.)}
226: \end{figure}
227: Beams between adjacent optical benches within a single spacecraft are
228: bounced off proof masses in the opposite way.  Light to be {\it
229:   {transmitted}} from the laser on an optical bench is {\it {first}}
230: bounced off the proof mass it encloses and then directed to the other
231: optical bench.  Upon reception it does {\it not} interact with the
232: proof mass there, but is directly mixed with local laser light, and
233: again down converted. These data are denoted $\tau_{31}$ and
234: $\tau_{21}$ in figure 2.
235: 
236: The terms in the following equations for the $s_{ij}$ and $\tau_{ij}$
237: phase measurements can now be developed from figures 1 and 2, and they
238: are for the particular LISA configuration in which all the lasers have
239: the same nominal frequency $\nu_0$, and the spacecraft are stationary
240: with respect to each other. The analysis covering the configuration
241: with lasers of different frequencies and spacecraft moving relative to
242: each other was done in \cite{2}, and we refer the reader to that
243: paper.
244: 
245: Consider the $s_{31} (t)$ process (equation (\ref{eq:3})) below.  The
246: photo receiver on the left bench of spacecraft $1$, which (in the
247: spacecraft frame) experiences a time-varying displacement $\vec
248: \Delta_1$, measures the phase difference $s_{31}$ by first mixing the
249: beam from the distant optical bench $2^*$ in direction $\hat n_3$, and
250: laser phase noise $p^*_2$ and optical bench motion $\vec \Delta^*_2$
251: that have been delayed by propagation along $L_3$, after one bounce
252: off the proof mass ($\vec \delta_1$), with the local laser light (with
253: phase noise $p_1$). Since for this simplified configuration no
254: frequency offsets are present, there is of course no need for any
255: heterodyne conversion. 
256: 
257: In equation (\ref{eq:4}) the $\tau_{31}$ measurement results from
258: light originating at the right-bench laser ($p^*_1$, $\vec
259: \Delta^*_1$), bounced once off the right proof mass ($\vec
260: \delta^*_1$), and directed through the fiber (incurring phase shift
261: $\mu_1(t)$), to the left bench, where it is mixed with laser light
262: ($p_1$).  Similarly the right bench records the phase differences
263: $s_{21}$ and $\tau_{21}$.  The laser noises, the gravitational wave
264: signals, the optical path noises, and proof-mass and bench noises,
265: enter into the four data streams recorded at vertex
266: $1$ according to the following expressions \cite{2}
267: \begin{eqnarray}
268: s_{21} & = &  s^{\rm gw}_{21} + s^{\rm opt. \ path}_{21} +
269: p_{3,2} - p^{*}_1 + \nu_{0} \ \left[ 2 {\hat n_2} \cdot {\vec \delta^{*}_1}  - 
270: {\hat n_2} \cdot {\vec \Delta^{*}_1} - {\hat n_2} \cdot {\vec
271:   \Delta_{3,2}} \right] \ ,
272: \label{eq:1}
273: \\
274: \tau_{21} & = &  p_{1} - p^{*}_1 + 2 \ \nu_{0} \ {\hat n_3} \cdot 
275: ({\vec \delta_1} - {\vec \Delta_1}) + \mu_1 \ ,
276: \label{eq:2}
277: \\
278: s_{31} & = & s^{\rm gw}_{31} + s^{\rm opt. \ path}_{31} + 
279: p^*_{2,3} - p_1 + \nu_{0} \ \left[ - \ 2 {\hat n_3} \cdot {\vec \delta_1} +
280: {\hat n_3} \cdot {\vec \Delta_1} + {\hat n_3} \cdot {\vec
281:   \Delta^*_{2,3}} \right] \ ,
282: \label{eq:3}
283: \\
284: \tau_{31} & = & p^{*}_{1} - p_1 
285: - \ 2 \ \nu_{0} \ {\hat n_2} \cdot ({\vec \delta^*_1} - {\vec
286:   \Delta^*_1}) + \mu_1 \ .
287: \label{eq:4}
288: \end{eqnarray}
289: \noindent
290: Eight other relations, for the readouts at vertices 2 and 3, are given
291: by cyclic permutation of the indices in equations
292: (\ref{eq:1})-(\ref{eq:4}).
293: 
294: The gravitational wave phase signal components, $s^{\rm gw}_{ij} \ , \
295: i, j = 1, 2, 3$, in equations (\ref{eq:1}) and (\ref{eq:3}) are
296: given by integrating with respect to time the equations (1), (2) of
297: reference \cite{5} that relate metric perturbations to frequency
298: shifts. The optical path phase noise contributions, $s^{\rm opt. \ 
299:   path}_{ij}$, which includes shot noise from the low signal-to-noise
300: ratio (SNR) in the links between the distant spacecraft, can be
301: derived from the corresponding term given in \cite{3}. The $\tau_{ij}$
302: measurements will be made with high SNR so that for them the shot
303: noise is negligible.
304: 
305: The laser-noise-free combinations of phase data can readily be
306: obtained from those given in \cite{3} for frequency data.  We use the
307: same notations: $X$, $Y$, $Z$, $\alpha$, $\beta$, $\gamma$, $\zeta$,
308: etc., but the reader should keep in mind that here these are phase
309: measurements.
310: 
311: The phase fluctuations, $s_{ij} \ , \ \tau_{ij} \ , \ i,j=1, 2, 3$,
312: are the fundamental measurements needed to synthesize all the
313: interferometric observables unaffected by laser and optical bench
314: noises.  If we assume for the moment these phase measurements to be
315: continuous functions of time, the three armlengths to be perfectly
316: known and constant, and the three clocks onboard the spacecraft to be
317: perfectly synchronized, then it is possible to cancel out exactly the
318: phase fluctuations due to the six lasers and six optical benches by
319: properly time-shifting and linearly combining the twelve measurements
320: $s_{ij} \ , \ \tau_{ij} \ , \ i,j=1, 2, 3$.  The simplest such
321: combination, the totally symmetrized Sagnac response $\zeta$, uses all
322: the data of Figure 2 symmetrically
323: \begin{eqnarray}
324: \zeta & = & s_{32,2} - s_{23,3} + s_{13,3} - s_{31,1} + s_{21,1} - s_{12,2}
325: \nonumber \\
326: & & + {1 \over 2}
327: [ ({\tau_{23}} - {\tau_{13}})_{,12}
328: + ({\tau_{31}} - {\tau_{21}})_{,23}
329: + ({\tau_{12}} - {\tau_{32}})_{,13}]
330: \nonumber \\
331: & & +
332: {1 \over 2} 
333: [ ({\tau_{23}} - {\tau_{13}})_{,3}
334: + ({\tau_{31}} - {\tau_{21}})_{,1}
335: + ({\tau_{12}} - {\tau_{32}})_{,2}] ,
336: \label{eq:5}
337: \end{eqnarray}
338: and its transfer functions to instrumental noises and gravitational
339: waves are given in \cite{3} and \cite{5} respectively.  In
340: particular, $\zeta$ has a ``six-pulse response'' to gravitational
341: radiation, i.e.  a $\delta$-function gravitational wave signal
342: produces six distinct pulses in $\zeta$ \cite{5}, which are located
343: with relative times depending on the arrival direction of the wave and
344: the detector configuration.
345: 
346: Together with $\zeta$, three more interferometric combinations,
347: ($\alpha, \beta, \gamma$), jointly generate the entire space of
348: interferometric combinations \cite{3}, \cite{5}, \cite{6}. Their
349: expressions in terms of the measurements $s_{ij}$, $\tau_{ij}$ are as
350: follows
351: \begin{eqnarray}
352: \alpha & = & s_{21} - s_{31} + s_{13,2} - s_{12,3} + s_{32,12} -
353: s_{23,13} + {1 \over 2} \ [({\tau_{23}} - {\tau_{13}})_{,2}
354: + ({\tau_{23}} - {\tau_{13}})_{,13} + ({\tau_{31}} - {\tau_{21}})
355: \nonumber \\
356: & &
357: \ + \ ({\tau_{31}} - {\tau_{21}})_{,123} + ({\tau_{12}} - {\tau_{32}})_{,3}
358: + ({\tau_{12}} - {\tau_{32}})_{,12}] \ ,
359: \label{eq:6}
360: \end{eqnarray}
361: \noindent
362: with $\beta$, and $\gamma$ derived by permuting the spacecraft indices
363: in $\alpha$. Like in the case of $\zeta$, a $\delta$-function
364: gravitational wave produces six pulses in $\alpha$, $\beta$, and
365: $\gamma$. In equations (\ref{eq:5}, \ref{eq:6}) it is important to
366: notice that the $\tau_{ij}$ measurements from each spacecraft always
367: enter into the interferometric measurements as differences taken at
368: the same time. This property naturally suggests a locking
369: configuration that makes these differences equal to zero, as we will
370: show in the next section.
371: 
372: We remind the reader that the four interferometric responses ($\alpha,
373: \beta, \gamma, \zeta$) satisfy the following relationship
374: \begin{equation}
375: \zeta - \zeta_{,123} =  \alpha_{,1} - \alpha_{,23} + \beta_{,2} -
376: \beta_{,13} + \gamma_{,3} - \gamma_{,12} \ .
377: \label{eq:7}
378: \end{equation}
379: Jointly they also give the expressions of the interferometric
380: combinations derived in \cite{3}, \cite{5}: the Unequal-arm
381: Michelson (${\rm X}, {\rm Y}, {\rm Z}$), the Beacon (${\rm P}, {\rm
382:   Q}, {\rm R}$), the Monitor (${\rm E}, {\rm F}, {\rm G}$), and the
383: Relay (${\rm U}, {\rm V}, {\rm W}$) responses
384: \begin{eqnarray}
385: {\rm X}_{,1} & = & \alpha_{,23} - \beta_{,2} - \gamma_{,3} + \zeta \ ,
386: \label{eq:8} \\
387: {\rm P} & = & \zeta - \alpha_{,1} \ ,
388: \label{eq:9} \\
389: {\rm E} & = & \alpha - \zeta_{,1} \ ,
390: \label{eq:10} \\
391: {\rm U} & = & \gamma_{,1} - \beta \ ,
392: \label{eq:11}
393: \end{eqnarray}
394: with the remaining expressions obtained from equations (\ref{eq:8},
395: \ref{eq:9}, \ref{eq:10}, \ref{eq:11}) by permutation of the spacecraft
396: indices. All these interferometric combinations have been shown to add
397: robustness to the mission with respect to failures of subsystems, and
398: potential design, implementation, or cost advantages \cite{3},
399: \cite{5}.
400:  
401: \subsection{Locking Conditions}
402: 
403: The space of all possible interferometric combinations can be
404: generated by properly time shifting and linearly combining the four
405: combinations ($\alpha, \beta, \gamma, \zeta$), as given above.
406: Although they have been derived by applying TDI to the twelve one-way
407: Doppler data, in what follows we will show that they can also be
408: written in terms of properly selected and time shifted two-way and
409: one-way Doppler measurements. These can be generated by phase locking
410: five of the six lasers to one of them, as it is described below.
411: 
412: Assume, without loss of generality, the laser on bench $1^*$ to be the
413: master. Although there are several other possible locking schemes, the one
414: chosen minimizes the number of locking conditions between the master
415: and any given slave. Furthermore, locking schemes relying on more than
416: one master could be implemented, but we will not address those in this
417: paper. We will also assume the spacecraft to be stationary relative to
418: each other. This assumption simplifies the analysis, and does not
419: affect the validity of the general result \cite{8}.
420: \begin{figure}
421: \centering
422: \includegraphics[width=5in, angle=0]{Fig3.ps}
423: \caption{Simplified optical layout of the LISA interferometer,
424: showing all the optical benches, proof masses and lasers.}
425: \end{figure}
426: Under this assumption, the frequency provided by the master laser
427: $1^*$ can be used as input reference for the slaved lasers. In other
428: words the slaves will then have the same center frequency as the master, and
429: their phase fluctuations will be related to the fluctuations of the
430: master laser as well as any other fluctuations introduced into the
431: received light beam prior to reception and locking.
432: 
433: In order to understand the topology of the beams as the various slaves
434: are locked to the master, let us follow the light paths from the
435: master laser, $1^*$, to the slaves, as shown in Figure 3.  Let us
436: start first with the light beam that is bounced off the back of the
437: proof-mass $1^*$. This beam is then directed to the other bench, $1$,
438: where it is used as the input frequency reference for the laser there,
439: and the measurement $\tau_{31}$ is made. Light is then re-transmitted
440: back to bench $1^*$ where the measurement $\tau_{21}$ is performed.
441: Since the phase of the laser $1$ is locked to that of the master, the
442: relative phase fluctuations $\tau_{31}$ can be adjusted as follows
443: \begin{equation}
444: \tau_{31} = \tau_{21} \ ,
445: \label{eq:12}
446: \end{equation}
447: where we have assumed the noise introduced by phase-locking to be
448: negligible (on this point see the theoretical derivations in Section
449: III and Appendix A). Similarly, light beams from lasers $1$ and $1^*$
450: are transmitted to the lasers on the benches $2^*$ and $3$
451: respectively, where the lasers are locked to the incoming beams.  From
452: benches $2^*$ and $3$ beams are transmitted to the lasers on benches
453: $2$ and $3^*$ respectively, where again locking is performed similarly
454: to what is done onboard spacecraft $1$.  Finally, along arm $1$, only
455: two one-way relative phase measurements can be performed as it is easy
456: to see.  Since for the moment we have assumed a configuration with
457: stationary spacecraft, the optical configuration described above can
458: be translated into the following {\it locking conditions} on some
459: relative phase fluctuation measurements
460: \begin{equation}
461: \tau_{31}  = \tau_{21} \ \ \ \ , \ \ \ \ 
462: \tau_{13}  =  \tau_{23} \ \ \ \ , \ \ \ \ 
463: \tau_{32}  =  \tau_{12} \ \ \ \ , \ \ \ \ s_{23} = s_{32} = 0 \ .
464: \label{eq:13b}
465: \end{equation}
466: The locking conditions define specific relationships among the phase
467: fluctuations from various noise sources. As an example, the conditions
468: \begin{equation}
469: \tau_{31} = \tau_{21} \ \ \ \ , \ \ \ \ s_{23} = 0 \ ,
470: \end{equation}
471: imply the following relationships among the laser phase fluctuations,
472: the proof-mass noises, the gravitational wave signal, and the two
473: bench noises onboard spacecraft $1$
474: \begin{eqnarray}
475: p^*_1 & = & p_1 + \nu_{0} \ {\hat n_3} \cdot 
476: ({\vec \delta_1} - {\vec \Delta_1}) + \nu_{0} \ {\hat n_2} \cdot 
477: ({\vec \delta^*_1} - {\vec \Delta^*_1}) \ , 
478: \label{eq:13c}
479: \\
480: 0 & = & s^{\rm gw}_{23} + s^{\rm opt. \ path}_{23} + 
481: p^*_{1,2} - p_3 + \nu_{0} \ \left[ - \ 2 {\hat n_2} \cdot {\vec \delta_3} +
482: {\hat n_2} \cdot {\vec \Delta_3} + {\hat n_2} \cdot {\vec
483:     \Delta^*_{1,2}} \right] \ ,
484: \label{eq:13d}
485: \end{eqnarray}
486: with similar expressions following from the other locking conditions
487: given in equation (\ref{eq:13b}).
488: 
489: If we now substitute the locking conditions into the expressions for
490: ($\alpha, \beta, \gamma, \zeta$), we obtain their expressions in terms
491: of the remaining measurements
492: \begin{eqnarray}
493: \zeta_{lo.} & = &  s_{13,3} - s_{31,1} + s_{21,1} - s_{12,2} \ ,
494: \label{eq:14}
495: \\
496: \alpha_{lo.} & = & s_{21} - s_{31} + s_{13,2} - s_{12,3} \ ,
497: \label{eq:15}
498: \\
499: \beta_{lo.} & = & - s_{12} + s_{21,3} + s_{13,23} - s_{31,12} \ ,
500: \label{eq:16}
501: \\
502: \gamma_{lo.} & = & s_{13} - s_{31,2} + s_{21,13} - s_{12,23} \ ,
503: \label{eq:17}
504: \end{eqnarray}
505: where the data ($s_{12}$, $s_{13}$) are one-way, and ($s_{21}$,
506: $s_{31}$) are effectively two-way Doppler measurements
507: due to locking.
508: 
509: The verification that the combinations ($\alpha_{lo.}, \beta_{lo.},
510: \gamma_{lo.}, \zeta_{lo.}$) exactly cancel the laser phase
511: fluctuations as well as the fluctuations due to the mechanical
512: vibrations of the optical benches, can be performed by substituting
513: the locked phase processes (such as \ref{eq:13c}, \ref{eq:13d}) into the
514: expressions for ($s_{12}, s_{13}, s_{21}, s_{31}$) (which are given by
515: equations (\ref{eq:1}, \ref{eq:3}) and their permutations), and by
516: further replacing them into equations (\ref{eq:14} - \ref{eq:17}).
517: 
518: The main result of implementing locking is quantitatively shown by
519: equations (\ref{eq:14} - \ref{eq:17}) in that the number of
520: measurements needed for constructing the entire space of
521: interferometric combinations LISA will be able to generate is smaller
522: by a factor of three than the number of measurements needed when only
523: one-way data are used.
524: 
525: Once ($\alpha_{lo.}, \beta_{lo.}, \gamma_{lo.}, \zeta_{lo.}$) are
526: constructed according to the expressions given in equations
527: (\ref{eq:14} - \ref{eq:17}), all the other interferometric
528: combinations can be derived by applying the identities given in
529: equations (\ref{eq:8} - \ref{eq:11}). As an example, it is
530: straightforward to show that equation (\ref{eq:8}) implies the
531: following expression for the unequal-arm Michelson combination $X_{lo.}$
532: \begin{equation}
533: X_{lo.} = [s_{21} - s_{31}] - [s_{21,33} - s_{31,22}] \ , 
534: \label{eq:18}
535: \end{equation}
536: which coincides with the expression for $X$, derived for the first time
537: in \cite{7}, in terms of the two-way Doppler measurements from the two
538: LISA arms.
539: 
540: As a final comment, we have analyzed also several locking
541: configurations needed when the spacecraft are moving relative to each
542: other. We have found that there exist techniques, when locking is
543: implemented, which are similar to the one analyzed in \cite{2} for
544: removing the noise of the onboard Ultra Stable Oscillators from the
545: phase measurements. The conclusions derived above for the case of
546: stationary spacecraft are therefore general, and an analysis
547: covering locking configurations with moving spacecraft
548: is available in \cite{8}.
549: 
550: \section{Phase Locking Performance}
551: 
552: %% This entire section has been altered starting from here, Daniel 1/23/03
553: 
554: The locking conditions given in equation (\ref{eq:13b}) reflect the
555: assumption that the noise due to the optical transponders is
556: negligible. In this section we analyze the noise added by the process
557: of locking the phase of the local laser to the phase of the received
558: light. This noise will be in addition to the optical path and USO
559: noises, which we consider separately.
560: 
561: A block diagram of the phase locking control system is shown in figure
562: \ref{Servo}(a). The system consists of a photoreceiver, phasemeter,
563: controller, actuator and laser (see section IV for a description of
564: these subsystems). Each of these subsystems can be characterized by
565: its transfer function and, when it applies, by a noise contribution
566: \cite{Abramovici}. The main inputs to the system are the phase noise
567: of the local laser, $p_L(s)$, and the phase fluctuations of the signal
568: beam from the distant spacecraft, $p_S(s)$ (with $s=\sigma + i \omega$
569: being the Laplace variable).  The closed loop output is the phase
570: noise of the retransmitted laser beam, $p_{CL}(s)$.
571: \begin{figure}[httb]
572: \centering 
573: \includegraphics[width=7in, angle=0]{Fig4.ps}
574: \caption{a) Block diagram of phase locking control system showing transfer 
575:   functions and noise contributions of various system components.  (b)
576:   Simplified control system block diagram.}
577: \label{Servo} 
578: \end{figure}
579: 
580: From the block diagram shown in figure 4a, it is easy to see that the
581: closed loop output phase $p_{CL}(s)$ can be written in terms of the
582: free-running laser phase noise, $p_L(s)$, the input signal phase
583: fluctuations, $p_S(s)$, and the various feedback components' transfer
584: functions and noises shown in Figure 4, as follows
585: \begin{equation}
586: p_{CL} = p_L+L  \left\{ N_A + A \ G \left[ N_M+M
587: \left(N_R+R\left\{N_D+p_S - p_{CL}\right\}\right) \right] \right\} \ .
588: \end{equation} 
589: This equation can be solved for $p_{CL}$
590: \begin{equation}
591: p_{CL} = 
592: {
593: {
594: p_L+L  \left\{ N_A + A \ G \left[ N_M+M
595: \left(N_R+R\left\{N_D+p_S\right\}\right) \right] \right\}}
596: \over 
597: {1+L \ A \ G \ M \ R}
598: } \ ,
599: \label{eq:18bis}
600: \end{equation}
601: where we have denoted with $N_R (s)$, $N_M (s)$, and $N_A (s)$ the
602: noises due to the photoreceiver, phasemeter, and actuator
603: respectively.  The detection noise, $N_D(s)$ is the error in the
604: measurement of the relative phase of the two beams. This noise is the
605: fundamental limit to the phase measurement process for the LISA
606: detection system (see Appendix A). Since the link from the phasemeter
607: to the controller will be digital, the controller noise will be due
608: only to the finite precision of the digital phase information. We thus
609: assume this noise to be negligible and do not include it in this
610: analysis.
611: 
612: Since (i) the product $M (s) \ R (s) = 1$ (the phase at the phasemeter output
613: ought to be equal to the phase at the input point of the
614: photoreceiver), (ii) the contribution of the actuator noise to the
615: output, $L(s) N_A(s)$, is much smaller than the noise from the
616: free-running laser (the free running laser noise is measured with the
617: actuators attached and thus intrinsically contains this noise source),
618: and (iii) the laser transfer function, $L(s)$, is a passive low pass
619: filter with a pole at several GHz (and so can be ignored in this
620: discussion), we conclude that the block diagram in \ref{Servo} (a) now
621: simplifies to that shown in figure \ref{Servo}(b), with the closed
622: loop output now given by
623: \begin{eqnarray}
624: p_{CL}&=&p_L+AG\left[N_T+p_S-p_{CL}\right] \ \ \ 
625: \longrightarrow p_{CL}=\frac{p_L+AG\left[N_T+p_S\right]}{1+AG} \\
626: p_{CL}-p_S&\simeq&N_T+\frac{p_L}{AG}\,\,\,\,\, (\rm{for}\,\,AG\gg1) \ .
627: \label{closedloop}
628: \end{eqnarray}
629: The quantity $N_T(s)$ is the total noise at the input to the
630: controller and is given by
631: \begin{eqnarray}
632: N_T&=&N_M+MN_R+N_D) \ .
633: \label{NT}
634: \end{eqnarray}
635: Equation (\ref{closedloop}) shows that phase locking will drive the
636: phase of the local laser, $p_{CL}(s)$, to that of the received laser,
637: $p_S(s)$, with an error introduced by two terms. The first term is the
638: total measurement noise in the phase measurement, $N_T(s)$, which is
639: in turn determined by the detection noise, photoreceiver noise and
640: phasemeter noise (see equation \ref{NT}). Recall that the detection
641: noise, $N_D(s)$, is the fundamental error in the measurement of the
642: relative phase of the two beams for the LISA detection system. A
643: rigorous calculation of this noise source, included in appendix
644: \ref{QMC}, shows that its root power spectral density is equal to
645: $\sim1~\mu{\rm cycle}/\sqrt{\rm Hz}$ at the output of the phase meter.
646: $N_R(s)$ is the electronic noise of the photoreceiver at the beat note
647: frequency.  Referenced to the output of the phasemeter this will be
648: well below the $1~\mu{\rm cycle}/\sqrt{\rm Hz}$ level. A phasemeter
649: noise floor of $\sim1~\mu{\rm cycle}/\sqrt{\rm Hz}$ has been set as a
650: requirement for the phasemeter noise. This level of performance has
651: already been demonstrated (\cite{8bis}, \cite{8SHA}) over the
652: frequency range of interest (1~mHz to 1 Hz) albeit with heterodyne
653: frequencies of a few kilohertz . Given these estimates of the
654: individual noise sources a total measurement error, $N_T (s)$ of less
655: than $2 \ \mu {\rm cycles}/\sqrt{\rm Hz}$ is expected.
656: 
657: The second source of error in Eq. (\ref{closedloop}) represents the
658: finite suppression of the free running laser noise.  This term is
659: inversely proportional to the loop gain and so can be reduced by
660: increasing the gain. Very high gains should be possible with LISA as
661: the frequencies of interest are very low \cite{Abramovici}, $\approx
662: 10$ ~Hz and lower.  A free-running laser frequency noise of
663: 1~MHz/$\sqrt{\rm Hz}$ at 1~mHz corresponds to a phase noise of
664: $10^9$~cycles/$\sqrt{\rm Hz}$.  A total loop gain of $10^{15}$ is
665: therefore required to suppress the contribution of laser frequency
666: noise down to 1~$\mu{\rm cycle}/\sqrt{\rm Hz}$. At low frequencies the
667: laser phase is altered by changing the temperature of the laser
668: crystal. This actuator has a typical (voltage to frequency) gain of
669: 5~GHz/Volt or $5 \times10^{12}~{\rm cycles}/\sqrt{\rm Hz}$ at 1~mHz.
670: Thus a controller gain of 200~V/cycle is needed. This gain requirement
671: could be eased by ``pre-stabilizing'' the laser frequency, for example
672: by locking to a low finesse cavity or other frequency reference.
673: Initial results from bench top experiments
674: (\cite{McNamaraPL},\cite{Ye}) indicate that loop gains of the order of
675: $10^{15}$ should be achievable, and hence that pre-stabilization may
676: not be necessary.
677: 
678: The master-slave configuration phase locking requirement should be
679: that the noise introduced by the locking process is insignificant
680: compared to the $20~\mu{\rm cycle}/\sqrt{\rm Hz}$ of optical path
681: noise allocated in the LISA noise budget \cite{1}. As the analysis
682: above has shown, phase locking is limited only by the measurement
683: noise (assuming adequate gain). This measurement noise is common to
684: both the one-way and the master-slave schemes and so, given adequate
685: gain of the phase locking loop, there will be no difference in the
686: performance of the two systems.
687: 
688: \section{Hardware Requirements}
689: 
690: In this section we compare the hardware needed by the two
691: implementations of TDI discussed in the previous sections. The
692: discussion will focus on the minimum hardware requirements and will
693: not fully consider redundancy or fallback options.  We will consider
694: LISA as composed of several basic subsystems, and compare the type and
695: quantity of the components required by each scheme.
696: 
697: \subsection{Laser Frequency Stabilization System}
698: 
699: Both schemes rely on a sufficiently high laser frequency stability.
700: This is because the cancellation of laser frequency noise via TDI is
701: not exact due to finite accuracy of the arm length knowledge, finite
702: timing accuracy, imperfect clock synchronization, and sampling time
703: jitters. The frequency stabilization system could be composed of
704: either an optical cavity, a gas cell, or a combination of both. The
705: output is an error signal proportional to the difference between the
706: laser frequency and the resonance frequency of the reference. Assuming
707: that an optical cavity is used as the frequency reference with a
708: Pound-Drever-Hall locking \cite{PDH} readout, the frequency
709: stabilization system will consist of an electro-optic phase modulator,
710: an optical cavity, a photoreceiver, a double-balanced mixer, and a low
711: pass filter \cite{8tris}.
712: 
713: \subsection{Phasemeter}
714: 
715: A phasemeter is a device capable of measuring the phase of a
716: photoreceiver output relative to the local USO. One could distinguish
717: between two types of phasemeter: single-quadrature phasemeters,
718: denoted PM(S), and full-range phasemeters, denoted PM(F). An example
719: of a single-quadrature phasemeter is a mixer. It will have only a
720: limited linear range as the output is generally sinusoidal. A
721: single-quadrature phasemeter has several potential advantages over a
722: full-range phasemeter including lower noise, higher speed operation,
723: increased reliability and lower power consumption. However, it is
724: restricted in usefulness to closed loop operation only and is
725: therefore not suitable for use in the one-way method. A
726: single-quadrature phasemeter could potentially be used in the
727: master-slaves configuration where one laser is phase locked to another
728: with a fixed phase shift (see section IV. E).
729: 
730: A full-range phasemeter has an output that is linearly proportional to
731: phase over the entire range $-\pi$ to $\pi$. An example of a
732: full-range phasemeter is a zero-crossing time interval analyzer. As we
733: show in section V. E, phasemeters used in an open-loop configuration
734: must have a dynamic range of at least $\sim10^{10}$ at $0.1$ mHz.
735: 
736: \subsection{Controller}
737: 
738: A controller takes the signal from either a frequency stabilization
739: system or phasemeter, amplifies and filters it appropriately, and
740: feeds it back to the frequency/phase actuators of a laser. A
741: controller is needed to frequency lock a laser to a frequency
742: reference or to phase lock one laser to another. In practice the
743: frequency locking and phase locking controllers will differ by a pole
744: in the controller transfer function and by a gain factor. Depending on
745: the sophistication of its design, a controller could potentially be
746: reconfigured in-flight to perform either function.
747: 
748: \subsection{Photoreceivers}
749: 
750: Each scheme requires twelve photoreceivers to measure the interference
751: from the front and back of the proof masses. The photoreceiver unit
752: will consist of a photodiode and low-noise electronic amplifiers.
753: Although the photoreceivers for LISA will contain quadrant photodiodes
754: for alignment sensing, this is irrelevant for the following discussion
755: which assumes that single element photodiodes are used.
756: 
757: \begin{table}
758:  {\centering \begin{tabular}{|c|c|c|}
759:  \hline
760: Spacecraft/Bench & Master-Slave Configuration &
761: One-Way Method\\
762:  \hline
763:  \hline
764: SC $1^*$& 1 FS, 2 PR, 2 PM(F), 1 Controller&1 FS, 2 PR, 2 PM(F), 1 Controller\\
765:  \hline
766:  SC 1& 2 PR, 2 PM(F), 1 Controller&1 FS, 2 PR, 2 PM(F), 1 Controller\\
767: \hline
768: SC $2^*$&2 PR, 1 PM(F), 1 PM(S), 1 Controller&1 FS, 2 PR, 2 PM(F), 1 Controller\\
769: \hline
770: SC 2&2 PR, 2 PM(F), 1 Controller&1 FS, 2 PR, 2 PM(F), 1 Controller\\
771: \hline
772: SC $3^*$&2 PR, 2 PM(F), 1 Controller&1 FS, 2 PR, 2 PM(F), 1 Controller\\
773: \hline
774: SC 3&2 PR, 1 PM(F), 1 PM(S), 1 Controller&1 FS, 2 PR, 2 PM(F), 1 Controller\\
775: \hline
776: 
777: \end{tabular}\par}
778: 
779: \caption{\label{tbl:comp1} Comparison of minimum system components
780: required for each scheme. FS: frequency stabilization system, 
781: PM(S): single-quadrature phasemeter, PM(F) full-range phasemeter, 
782: PR: photoreceiver.}
783: \end{table}
784: 
785: \subsection{Requirements for the Master-Slave Configuration}
786: 
787: Table I summarizes the system components needed on each optical bench.
788: The quantities shown represent the minimum requirements with no
789: redundancy included. In the master-slave configuration scheme one
790: master laser is frequency stabilized to its own frequency reference.
791: All other lasers are phase locked in a chain to this master laser, as
792: described in section IIA. For this reason the minimum system
793: requirement is only one frequency stabilization system. However, the
794: capability of stabilizing other lasers to a local stabilization system
795: should be included for redundancy against failure of the master's
796: stabilizing device.  Providing each laser with a frequency
797: stabilization system would provide a high level of redundancy and
798: maintain compatibility with the one-way mode of operation.
799: 
800: The master-slave configuration will require at least four full-range
801: phasemeters for the main signal read out photoreceivers, $s_{21}$,
802: $s_{31}$, $s_{12}$ and $s_{13}$. Full-range phasemeters will also be
803: required for measuring the signals derived from the back side of the
804: proof masses, $\tau_{ij}$, as the lasers on adjacent benches are
805: locked by suppressing the difference of the phasemeter outputs,
806: $\tau_{ij}-\tau_{kj}=0$. If just one of these phasemeter outputs were
807: used for phase locking then a single-quadrature phasemeter could be
808: utilized. However, the noise in the optical fiber linking the benches,
809: $\mu_{i}$, would then be imposed on the phase of the slave laser.
810: Although this noise would be removed by including the second detector
811: in the time-delay interferometry processing, this would impose
812: unnecessary requirements on the stability of the fiber link to prevent
813: increasing the slave laser's frequency noise. Furthermore, by
814: suppressing the difference of these phasemeter outputs, we do not need
815: to record this information for processing, as only the difference of
816: the phasemeter outputs appears in the TDI equations (see for example,
817: equations \ref{eq:5} and \ref{eq:6}). Single-quadrature phasemeters
818: could potentially be used on the remaining two photoreceivers where
819: phase locking of the beams returning to spacecraft $1$ ($s_{23} =
820: s_{32} = 0$) is performed. This is a relatively minor simplification,
821: as suitable full-range phasemeters must be developed for the remaining
822: ten photoreceivers. If the phase locking hierarchy must be reordered,
823: for example due to a frequency stabilization system failure, then
824: full-range phasemeters will also be required at these positions.
825: Finally, implementing full-range phasemeters at all photoreceiver
826: outputs will maintain compatibility with the one-way method. For these
827: reasons we will drop the (F) or (S) suffix for the phasemeters and
828: assume that only full-range phasemeters are used.
829: 
830: Table II summarizes the total number of components required and the
831: number of data streams to be recorded by each scheme. Using the
832: master-slave configuration only four data streams remain to be
833: measured ($s_{21}$, $s_{31}$, $s_{12}$ and $s_{13}$) as all other
834: variables have been suppressed to effectively zero and therefore do
835: not need to be recorded (although they should be monitored to ensure
836: proper operation of the phase locking systems). One potential concern
837: with the master-slave configuration is in the non-local nature of the
838: control system, that is to say the main phase input to the control
839: loop comes from the light from the distant spacecraft. The amplitude
840: and phase of this beam could be adversely affected by many factors
841: such as spacecraft alignment. Although this will also affect the
842: quality of the one-way measurements, it could be more detrimental to
843: the master-slave configuration. For example, the signal intensity could
844: become so low to cause loss of phase lock entirely. Furthermore,
845: because all the slaves are linked to the master by the phase-locking
846: chain, if one phase-locking link is disrupted then all downstream
847: links may also be lost. The severity of this non-local control problem
848: depends on how often lock will be disrupted, and the difficulty of
849: lock reacquisition.
850: 
851: \begin{table}
852:  {\centering \begin{tabular}{|c|c|c|c|}
853:  \hline
854: Subsystem&Master-Slave Configuration &One-Way Method\\
855:  \hline
856:  \hline
857: FS&1&6\\
858:  \hline
859:  PR&12&12\\
860: \hline
861:  PM&12&12\\
862: \hline
863: Controllers&6&6\\
864: \hline
865: Observables&4&9\\
866: \hline
867: \end{tabular}\par}
868: \caption{\label{tbl:comp2}Summary of minimum system components
869: required for each scheme. FS: frequency stabilization system, 
870: PM: phasemeter, PR: photoreceiver, Observables: number of data 
871: streams required for processing.}
872: \end{table}
873: 
874: \subsection{Requirements for the One-way Method}
875: 
876: The one-way method employs a very symmetric configuration consisting
877: of three identical spacecraft each containing two identical optical
878: benches. The components needed on each optical bench are shown in the
879: right hand column of Table I. The six lasers are frequency stabilized
880: to their six respective frequency references. The phases of the beat
881: notes of each local laser with the lasers from the adjacent spacecraft
882: and bench are measured by full-range phasemeters to provide twelve
883: data streams.  The data from the phasemeters at the back of the proof
884: masses on adjacent benches can be combined before being recorded
885: without loss of generality. This reduces the total number of data
886: streams to be recorded and processed to nine, as shown in Table II.
887: However, as we will show in section V C the number of data streams to
888: be exchanged between spacecraft will be the same for the one-way and
889: master-slaves configuration.
890: 
891: From Tables I and II it is clear that the one-way and master-slaves
892: configurations are almost identical in terms of the quantity of
893: components required. However, there are several more subtle
894: differences in the hardware requirements of the two schemes. A
895: disadvantage of the one-way method is that the laser frequencies may
896: differ by as much as $\pm300$~MHz if high finesse cavities are used as
897: the frequency references. This maximum frequency offset is determined
898: by the free spectral range of the reference cavity, where a cavity
899: with a round trip optical path of 0.5~m has been assumed. This large
900: frequency offset will place greater demands on the photoreceivers'
901: bandwidth than the master-slave configuration where frequency offsets
902: can be kept to the minimum dictated by the Doppler shifts (less than
903: $\pm 10$~MHz). Not only does the high heterodyne frequency place strict
904: requirements on the photodetector bandwidth, but also on the bandwidth
905: stability. For example, assume that photoreceivers with a bandwidth of
906: $f_{bw}\approx1$~GHz are used for the main signal readouts. Although a
907: heterodyne frequency, $f_{h}$, of 300~MHz is within the 1~GHz
908: photoreceiver bandwidth there will be a 0.05~cycle phase delay at this
909: frequency (the phase shift is equal to -arctan$(f_{h}/f_{bw})$ radians
910: for a simple single-pole type frequency response). If the bandwidth of
911: a photodetector changes, then this phase shift will also change in a
912: way that is indistinguishable from the effects of a gravitational
913: wave. A simple calculation shows that a bandwidth change of a mere
914: $0.023\%$ (or 230kHz) would introduce a phase signal of
915: 10~$\mu$cycles/$\sqrt{\rm Hz}$ for information at 300~MHz. If the
916: heterodyne frequency is kept to 10~MHz or less, then a photoreceiver
917: bandwidth change of more than $0.6\%$ (or 6~MHz) would be needed to
918: produce a 10~$\mu$cycles/$\sqrt{\rm Hz}$ phase shift.
919: 
920: In section III it was shown that a total loop gain of $10^{15}$ at 1
921: mHz is required to ensure that the phase locking loop performs
922: correctly. Using the one-way method the phase locking loops are
923: replaced by frequency locking to the reference cavity. The frequency
924: stabilization system is expected to be limited by fluctuations in
925: length of the cavity at a level of the order of 10 Hz/$\sqrt{\rm Hz}$
926: at 1 mHz.  Under this condition no advantage is gained by suppressing
927: the measured laser noise below this level and so a loop gain of
928: $\approx10^{6}$ should suffice. The one-way method therefore requires
929: a controller gain of only $2\times10^{-7}$~Volts/cycle at 1~mHz
930: compared to the 200~Volts/cycle needed for the master-slave
931: configuration.
932: 
933: \section{Accuracy and precision requirements}
934: 
935: The limitations on the effectiveness of the TDI technique, either when
936: the one-way or the master-slave configuration is implemented, come
937: not only from all the secondary noise sources affecting the
938: measurements $s_{ij} \ , \ \tau_{ij} \ , \ i,j=1, 2, 3$ (such as
939: proof-mass and optical path noises) but most importantly from the
940: finite accuracy and precision of the quantities needed to synthesize
941: the laser-noise free observables themselves. In order to synthesize
942: the four generators of the space of all interferometric combinations,
943: we need:
944: 
945: \noindent
946: (i) to know the distances between the three pairs of spacecraft;
947: 
948: \noindent
949: (ii) to synchronize the clocks onboard the three spacecraft, which
950: are used in the data acquisition and digitization process;
951: 
952: \noindent
953: (iii) to be able to apply time-delays that are not integer multiples of
954: the sampling time of the digitized phase measurements;
955: 
956: \noindent
957: (iv) to minimize the effects of the jitter of the
958: sampling times themselves; and 
959: 
960: \noindent
961: (v) to have sufficiently high dynamic range in the digitized data in
962: order to be able to recover the gravitational wave signal after
963: removing the laser noise.
964: 
965: In the following subsections we will assume the secondary random
966: processes to be due to the proof masses and the optical path noises
967: \cite{3}. We will estimate the minimum values of the accuracies and
968: precisions of the physical quantities listed above that allow the
969: suppression of the laser frequency fluctuations below the level
970: identified by the secondary noise sources. For each physical quantity
971: the estimate of the accuracy and/or precision needed will be performed
972: by assuming all the remaining errors to be equal to zero.  Our
973: estimates, therefore, will provide only an order of magnitude estimate
974: of the accuracies and/or precisions needed for successfully
975: implementing TDI.
976: 
977: \subsection{Armlength accuracy}
978: 
979: The TDI combinations described in the previous sections rely on the
980: assumption of knowing the armlengths sufficiently accurately to
981: suppress laser noise well below other noises. Since the three
982: armlengths will be known only within the accuracies $\delta L_i \ , \ 
983: i=1, 2, 3$ respectively, the cancellation of the laser frequency
984: fluctuations from the combinations ($\alpha, \beta, \gamma, \zeta$)
985: will no longer be exact.  In order to estimate the magnitude of the
986: laser fluctuations remaining in these data sets, let us define $\hat
987: L_i \ , \ i=1, 2, 3$ to be the estimated armlengths of LISA.  They are
988: related to the {\it true} armlengths $L_i \ , \ i=1, 2, 3$, and the
989: accuracies $\delta L_i \ , \ i=1, 2, 3$ through the following
990: expressions
991: \begin{equation}
992: \hat L_i = L_i + \delta L_i \ \ \ \ \ , \ \ \ \ \ i=1, 2, 3 \ .
993: \label{eq:19}
994: \end{equation}
995: In what follows we will treat the three armlengths $L_i \ , \ i=1, 2,
996: 3$ as constants equal to $16.7$ light seconds. We will derive later on
997: the time scale during which such an assumption is valid. We will also
998: assume to know with infinite accuracies and precisions all the
999: remaining physical quantities needed to successfully synthesize the
1000: TDI generators.
1001: 
1002: If we now substitute equation (\ref{eq:19}) into equations
1003: (\ref{eq:5}), and expand it to first order in $\delta L_i$, it is easy
1004: to derive the following approximate expression for ${\hat \zeta} (t)$,
1005: which now will show a non-zero contribution from the laser noises
1006: \begin{equation}
1007: {\hat \zeta} (t) \simeq \zeta (t) + 
1008: [{\dot p}_{2,13} - {\dot p}^*_{3,12}] \ \delta L_1
1009: + [{\dot p}_{3,12} - {\dot p}^*_{1,23}] \ \delta L_2
1010: + [{\dot p}_{1,23} - {\dot p}^*_{2,13}] \ \delta L_3 \ ,
1011: \label{eq:20}
1012: \end{equation}
1013: where the ``${\dot \ }$'' denotes time derivative. Time-Delay Interferometry
1014: can be considered effective if the magnitude of the remaining
1015: fluctuations from the lasers are smaller than the fluctuations due to
1016: the other noise sources entering in $\zeta (t)$, namely proof mass and
1017: optical path noises. This requirement implies a limit in the
1018: accuracies of the measured armlengths.
1019: 
1020: Let us assume the six laser phase fluctuations to be uncorrelated to
1021: each other, their one-sided power spectral densities to be equal, the
1022: three armlengths to differ by a few percent, and the three armlength
1023: accuracies also to be equal. By requiring the magnitude of the
1024: remaining laser noises to be smaller than the secondary noise sources,
1025: it is straightforward to derive, from Eq. (\ref{eq:20}) and the
1026: expressions for the proof mass and optical path noises entering into
1027: $\zeta (t)$ given in \cite{3}, the following constraint on the common
1028: armlength accuracy $| \delta L_{\zeta}|$
1029: \begin{equation}
1030: |\delta L_{\zeta}| \le \ {1 \over {2 \pi f}} \ 
1031: \sqrt{ 
1032: {4 \ \sin^2(\pi f L) \ S^{proof \ mass}_p (f) + S^{optical \
1033:     path}_p (f)} \over {S_p (f)}
1034: } \ .
1035: \label{eq:21}
1036: \end{equation}
1037: Here $S_p$, $S^{proof \ mass}_p$, $S^{optical \ path}_p$ are the
1038: one-sided power spectral densities of the phase fluctuations of a
1039: stabilized laser, a single proof mass, and a single-link optical path
1040: respectively \cite{3}. If we take them to be equal to the following
1041: functions of the Fourier frequency $f$ \cite{2,3}
1042: \begin{eqnarray}
1043: S_p (f) & = & 2.3 \times 10^{-1}  \ f^{- 8/3} + 1.4 \times  10^{-9} \ f^{-27/5} \ {\rm
1044:   cycles}^2 \ {\rm Hz}^{-1} \ ,
1045: \label{eq:22}
1046: \\
1047: S^{proof \ mass}_p (f) & = & 5.8 \times 10^{-21} \ f^{-4} \ {\rm
1048:   cycles}^2 \ {\rm Hz}^{-1} \ ,
1049: \label{eq:23}
1050: \\
1051: S^{optical \ path}_p (f) & = & 4.1 \times 10^{-10} \ {\rm
1052:   cycles}^2 \ {\rm Hz}^{-1} \ ,
1053: \label{eq:24}
1054: \end{eqnarray}
1055: (where $f$ is in Hz), we find that the right-hand-side of the
1056: inequality given by equation (\ref{eq:21}) reaches its minimum of
1057: about $16$ meters at the Fourier frequency $f_{min} = 1.0 \ \times
1058: 10^{-4} \ {\rm Hz}$, over the assumed ($10^{-4}, 1$) Hz LISA band.
1059: This implies that, if the armlength knowledge $| \delta L_{\zeta} |$
1060: can be made smaller than $16$ meters, the magnitude of the residual
1061: laser noise affecting the $\zeta$ combination will be below that
1062: identified by the secondary noises.  This reflects the fact that the
1063: armlength accuracy is a decreasing function of the frequency.  For
1064: instance, at $10^{-3}$ Hz the armlength accuracy goes up by almost an
1065: order of magnitude to about $155$ meters.
1066: 
1067: A perturbative analysis similar to the one described above can be
1068: performed for the remaining generators ($\alpha, \beta, \gamma$). We
1069: find that the corresponding inequality for the armlength accuracy
1070: required for the $\alpha$ combination, $| \delta L_{\alpha} |$, is
1071: equal to \cite{3,7}
1072: \begin{equation}
1073: |\delta L_{\alpha}| \le \ {1 \over {2 \pi f}} \ 
1074: {
1075: \sqrt{ 
1076: {[8 \ \sin^2(3 \pi f L) + 16 \ \sin^2(\pi f L)]
1077: \ S^{proof \ mass}_\phi (f) + 6 \ S^{optical \
1078:     path}_\phi (f)} \over {6 \ S_p (f)}
1079: }} \ ,
1080: \label{eq:25}
1081: \end{equation}
1082: with similar inequalities also holding for $\beta$ and $\gamma$.
1083: Equation (\ref{eq:25}) implies a minimum of the function on the
1084: right-hand-side equal to about $31$ meters at the Fourier frequency
1085: $f_{min} = 1.0 \ \times 10^{-4} \ {\rm Hz}$, while at $10^{-3}$ Hz
1086: the armlength accuracy goes up to $180$ meters.
1087: 
1088: Armlength accuracies significantly smaller than the level derived
1089: above can be achieved by implementing laser ranging measurements along
1090: the three LISA arms \cite{1}, and we do not expect this to be a
1091: limitation for TDI.
1092: 
1093: In relation to the accuracies derived above, it is interesting to
1094: calculate the time scales during which the armlengths will change by
1095: an amount equal to the accuracies themselves.  This identifies the
1096: minimum time required before updating the armlength values
1097: in the TDI combinations.
1098: 
1099: It has been calculated by Folkner {\it et al.} \cite{9} that the
1100: relative longitudinal speeds between the three pairs of spacecraft,
1101: during approximately the first year of the LISA mission, can
1102: be written in the following approximate form
1103: \begin{equation}
1104: V_{i,j} (t) = V^{(0)}_{i,j} \ \sin\left({{2 \pi t} \over
1105: T_{i,j}}\right) \qquad \qquad (i,j) = (1,2) \ ; \ (1,3) \ ; \ (2,3) \ ,
1106: \label{eq:26}
1107: \end{equation}
1108: where we have denoted with $(1,2), (1,3), (2,3)$ the three possible
1109: spacecraft pairs, $V^{(0)}_{i,j}$ is a constant velocity, and
1110: $T_{i,j}$ is the period for the pair $(i, j)$.  In reference \cite{9}
1111: it has also been shown that the LISA trajectory can be selected in
1112: such a way that two of the three arms' rates of change are essentially
1113: equal during the first year of the mission.  Following reference
1114: \cite{9}, we will assume $V^{(0)}_{1,2} = V^{(0)}_{1,3} \ne
1115: V^{(0)}_{2,3}$, with $V^{(0)}_{1,2} = 1 $ m/s, $V^{(0)}_{2,3} = 13 $
1116: m/s, $T_{1,2} = T_{1,3} \approx 4 $ months, and $T_{2,3} \approx 1$
1117: year.  From equation (\ref{eq:26}) it is easy to derive the variation
1118: of each armlength, for example $\Delta L_3 (t)$, as a function of the
1119: time $t$ and the time scale $\delta t$ during which it takes place
1120: \begin{equation}
1121: \Delta L_3 (t) = V^{(0)}_{1,2} \ \sin\left({{2 \pi t} \over
1122: T_{1,2}}\right) \delta t \ .
1123: \label{eq:27}
1124: \end{equation}
1125: Equation (\ref{eq:27}) implies that a variation in armlength $\Delta
1126: L_3 \approx 10 \ {\rm m}$ can take place during different time scales,
1127: depending on when during the mission this change takes place.  For
1128: instance, if $t \ll T_{1,2}$ we find that the armlength $L_3$ changes
1129: by more than its accuracy ($\approx 10$ meters) after a time $\delta t = 2.3
1130: \times 10^3$ seconds.  If however $t \simeq T_{1,2}/4$, the armlength
1131: will change by the same amount after only $ \delta t \simeq 10$ seconds
1132: instead.
1133: 
1134: \subsection{Clock synchronization accuracy}
1135: 
1136: The effectiveness of the TDI data combinations requires the clocks
1137: onboard the three spacecraft to be synchronized.  Since the clocks
1138: will be synchronized with a finite accuracy, the laser noises will no
1139: longer cancel out exactly and the fraction of the laser frequency
1140: fluctuations that will remain into the TDI combinations will be
1141: proportional to the magnitude of the synchronization accuracy.  In
1142: order to identify the minimum level of off-synchronization among the
1143: clocks that can be tolerated, we will proceed by treating one of the
1144: three clocks (say the clock onboard spacecraft $1$) as the master
1145: clock defining the time for LISA, and the other two to be synchronized
1146: to it.  The relativistic (Sagnac) time-delay effect due to the fact
1147: that the LISA trajectory is a combination of two rotations, each with
1148: a period of one year, will have to be accounted for in the
1149: synchronization procedure.  This is a procedure well known in the field of
1150: time-transfer, and we refer the reader to the appropriate literature
1151: for discussions on this point \cite{10}. Here we will disregard this
1152: relativistic effect, and assume it can be compensated for with an
1153: accuracy better than the actual synchronization accuracy we
1154: derive below.
1155: 
1156: Let us denote by $ \delta t_2$, $ \delta t_3$, the time accuracies
1157: (time-offsets) for the clocks onboard spacecraft $2$ and $3$
1158: respectively. If $t$ is the time onboard spacecraft $1$, then what is
1159: believed to be time $t$ onboard spacecraft $2$ and $3$ is actually
1160: equal to the following times
1161: \begin{eqnarray}
1162: {\hat t_2} = t + \delta t_2 \ ,
1163: \label{eq:28}
1164: \\
1165: {\hat t_3} = t + \delta t_3 \ .
1166: \label{eq:29}
1167: \end{eqnarray}
1168: If we now substitute equations (\ref{eq:28}, \ref{eq:29}) into the
1169: equation (\ref{eq:5}) for $\zeta$, for instance, and expand it to
1170: first order in $\delta t_i \ , \ i=2, 3$, it is easy to derive the
1171: following approximate expression for ${\hat \zeta} (t)$, which shows
1172: the following non-zero contribution from the laser noises
1173: \begin{equation}
1174: {\hat \zeta} (t) \simeq  \zeta (t) + 
1175: [{\dot p}_{1,23} - {\dot p}^*_{3,12} + {\dot p}^*_{2,13} - {\dot
1176:   p}_{2,13}] \ \delta t_2
1177: + [{\dot p}_{2,13} - {\dot p}^*_{1,23} + {\dot p}^*_{3,12} - {\dot
1178:   p}_{3,12}] \ \delta t_3 \ .
1179: \label{eq:30}
1180: \end{equation}
1181: By requiring again the magnitude of the remaining fluctuations from
1182: the lasers to be smaller than the fluctuations due to the other
1183: (secondary) noise sources affecting $\zeta (t)$, it is possible to
1184: derive an upper limit for the accuracies of the synchronization of the
1185: clocks.  If we assume again the six laser phase fluctuations to be
1186: uncorrelated to each other, their one-sided power spectral densities
1187: to be equal, the three armlengths to differ by a few percent, and the
1188: two time-offsets' magnitudes to be equal, by requiring the magnitude
1189: of the remaining laser noises to be smaller than the secondary noise
1190: sources it is easy to derive the following constraint on the time
1191: synchronization accuracy $| \delta t_{\zeta}|$
1192: \begin{equation}
1193: |\delta t_{\zeta}| \le \ {1 \over {2 \pi f}} \ 
1194: \sqrt{ 
1195: {12 \ \sin^2(\pi f L) \ S^{proof \ mass}_p (f) + 3 \ S^{optical \
1196:     path}_p (f)} \over {4 \ S_p (f)}
1197: } \ ,
1198: \label{eq:31}
1199: \end{equation}
1200: with $S_p$, $S^{proof \ mass}_p$, $S^{optical \ path}_p$ again as
1201: given in equations (\ref{eq:22}-\ref{eq:24}).
1202: 
1203: We find that the right-hand-side of the inequality given by equation
1204: (\ref{eq:31}) reaches its minimum of about $47$ nanoseconds at the
1205: Fourier frequency $f_{min} = 1.0 \ \times 10^{-4} \ {\rm Hz}$. In other
1206: words, clocks synchronized at a level of accuracy better than $47$
1207: nanoseconds will imply a residual laser noise that is smaller than 
1208: the secondary noise sources entering into the $\zeta$ combination.
1209: 
1210: An analysis similar to the one described above can be performed for the
1211: remaining generators ($\alpha, \beta, \gamma$). For them we find that the
1212: corresponding inequality for the accuracy in the synchronization of
1213: the clocks is now equal to
1214: \begin{equation}
1215: |\delta t_{\alpha}| \le \ {1 \over {2 \pi f}} \ 
1216: \sqrt{ 
1217: {[4 \ \sin^2(3 \pi f L) + 8 \ \sin^2(\pi f L)]
1218: \ S^{proof \ mass}_p (f) + 3 \ S^{optical \
1219:     path}_p (f)} \over {4 \ S_p (f)}
1220: } \ ,
1221: \label{eq:32}
1222: \end{equation}
1223: with equal expressions holding also for $\beta$ and $\gamma$.  The
1224: function on the right-hand-side of equation (\ref{eq:32}) has a
1225: minimum equal to $88$ nanoseconds at the Fourier frequency $f_{min} =
1226: 1.0 \ \times 10^{-4} \ {\rm Hz}$. As for the armlength accuracies,
1227: also the timing accuracy requirements become less stringent at higher
1228: frequencies. At $10^{-3}$ Hz, for instance, the timing accuracy for
1229: $\zeta$ and $\alpha, \beta, \gamma$ go up to $446$ and $500$ ns
1230: respectively.
1231: 
1232: A $50$ ns accuracy translates into a $15$ meter armlength accuracy,
1233: which we argued earlier to be easily achievable by the use of laser
1234: ranging. We therefore expect the synchronization of the three
1235: clocks to be achievable at the level derived above.
1236: 
1237: \subsection{Telemetered signals and their sampling}
1238: 
1239: To reduce the LISA-to-Earth telemetry requirements, it is expected
1240: that the normal operational mode will not telemeter the phase time
1241: series to the ground directly. Rather, we expect the TDI observables
1242: to be computed at the LISA array, and then only the (relatively low
1243: data rate) laser-noise-free combinations transmitted to Earth. Thus,
1244: both implementations of TDI discussed in this paper (the one-way
1245: method and the master-slave configuration) require phase measurements
1246: data to be exchanged among the spacecraft in order to synthesize the
1247: four generators of the space of all interferometric combinations.
1248: Although it is clear that the master-slave configuration implies a
1249: smaller number of measurements than that required by the one-way
1250: method, the actual number of data that will need to be exchanged among
1251: the spacecraft can be made to be exactly the same for both, making
1252: their inter-spacecraft telemetry requirements identical. This can
1253: easily be understood by rewriting the four generators ($\alpha, \beta,
1254: \gamma, \zeta$) in the following forms
1255: \begin{eqnarray}
1256: \zeta & = & 
1257: [s_{21,1}- s_{31,1}] + {1 \over 2} [
1258: ({\tau_{31}} - {\tau_{21}})_{,1}
1259: + ({\tau_{31}} - {\tau_{21}})_{,23}
1260: ]
1261: \nonumber \\
1262: & & + [s_{32,2}- s_{12,2}] + {1 \over 2} [
1263: ({\tau_{12}} - {\tau_{32}})_{,2}
1264: + ({\tau_{12}} - {\tau_{32}})_{,13}
1265: ]
1266: \nonumber \\
1267: & & + [s_{13,3}- s_{23,3}]
1268: + {1 \over 2} [
1269: ({\tau_{23}} - {\tau_{13}})_{,3}
1270: + ({\tau_{23}} - {\tau_{13}})_{,12}
1271: ] \ ,
1272: \label{eq:33}
1273: \end{eqnarray}
1274: \begin{eqnarray}
1275: \alpha & = & [s_{21} - s_{31}] + {1 \over 2} \ [({\tau_{31}} - {\tau_{21}})
1276: + ({\tau_{31}} - {\tau_{21}})_{,123}]
1277: \nonumber \\
1278: & & + [s_{32,12} - s_{12,3}] + {1 \over 2} \ [({\tau_{12}} - {\tau_{32}})_{,3}
1279: + ({\tau_{12}} - {\tau_{32}})_{,12}]
1280: \nonumber \\
1281: & & + [s_{13,2} - s_{23,13}] + {1 \over 2} \ [({\tau_{23}} - {\tau_{13}})_{,2}
1282: + ({\tau_{23}} - {\tau_{13}})_{,13}] \ .
1283: \label{eq:34}
1284: \end{eqnarray}
1285: Equations (\ref{eq:33}, \ref{eq:34}) show that each generator can be
1286: formed by summing three different linear combinations of the data,
1287: each involving phase measurements performed onboard only a specific
1288: spacecraft. As an example, let us assume without loss of generality
1289: that $\zeta$ will be synthesized onboard spacecraft $1$. This means
1290: that spacecraft $2$ and $3$ will simply need to telemeter to spacecraft
1291: $1$ the particular combinations of the measurements they have made,
1292: which enter into the $\zeta$ combination. Since the space of all the
1293: interferometric combinations can be constructed by using four
1294: generators ($\alpha, \beta, \gamma, \zeta$) we conclude that
1295: spacecraft $2$ and $3$ will each have to telemeter to spacecraft $1$
1296: four uniquely defined combinations of the measurements they have
1297: performed.
1298: 
1299: The time-delay interferometric combinations require use of phase
1300: measurements that are time-shifted with enough accuracy to bring the
1301: laser phase noise below the secondary noise sources.  The required
1302: time resolution in the time-shifts should be equal to about $\sim 50$
1303: ns for shifts tens of seconds in size.  This is because the correct
1304: sample of the shifted data should be as accurate as the armlength
1305: accuracy itself.  It can be shown that performing the time-shifting,
1306: on data sampled at $\sim 10$ Hz, by using digital interpolation
1307: filters, does not provide the required accuracy to effectively cancel
1308: the laser phase noises (see Appendix B for a detailed calculation).
1309: 
1310: An alternative approach \cite{11} for achieving a timing accuracy of
1311: at least $50$ ns would be to sample each measurement at $\sim 20$ MHz
1312: or higher and store $\sim 2.0 \times 10^9$ samples in a ring buffer
1313: for obtaining the data points of this measurement at the needed times.
1314: The phasemeter would then average these measurements over a fixed time
1315: period (perhaps a tenth of a second) centered around the sampled times
1316: at which the phase measurements are needed.  The data is then
1317: exchanged among the spacecraft and the TDI combinations are formed.
1318: 
1319: This method can however be further refined by actually sampling every
1320: phase difference a few times, each time at $\sim 10$ Hz, but with a
1321: delay between the start time of every sampled version of the same
1322: phase difference. That is, we envision triggering the phasemeter such
1323: that the time series are sampled at the times required to form the TDI
1324: combinations. In this case the limitation of the finite sampling
1325: time in the determination of the delayed phase measurement is replaced
1326: by the timing precision of the phase measurements, which can be many
1327: orders of magnitude smaller than the smallest sampling time of the
1328: phasemeter, as we will show below. 
1329: 
1330: As a concrete example, the ($\alpha_{lo.}, \beta_{lo.}, \gamma_{lo.},
1331: \zeta_{lo.}$) basis in the master-slave configuration requires
1332: measurements $\{s_{21}, s_{21,1}, s_{21,3}, s_{21,13}, s_{31},
1333: s_{31,1}, s_{31,2}, s_{31,12}\}$ from spacecraft 1, measurements $\{
1334: s_{12}, s_{12,2}, s_{12,3}, s_{12,23} \}$ from spacecraft 2, and
1335: measurements $\{s_{13}, s_{13,2}, s_{13,3}, s_{13,23}\}$ from
1336: spacecraft 3. By sampling the data $s_{21}$ at the times $n / f_s, n /
1337: f_s - L_1, n / f_s - L_3, n / f_s - L_1 - L_3$, where $f_s \sim 10$ Hz
1338: is the sampling frequency, and $n = 0, 1, 2, ...$ (and similarly for
1339: $s_{12}, s_{13}$ and $s_{31}$) we can obtain the entire data at the
1340: required times, these being limited only by the timing precision of
1341: the phasemeters, the time synchronization accuracies of the clocks,
1342: and the armlength accuracies. This scheme requires sampling each
1343: signal four times at a sampling frequency ($10$ Hz) much smaller than
1344: what would be needed if sampling the data to the granularity required
1345: by the TDI combinations. Of course to correctly sample at $10$ Hz we
1346: must first ensure that the signal frequency bandwidth is less than $5$
1347: Hz to avoid aliasing problems.
1348: 
1349: In practice, the times at which every sample from a given signal are
1350: taken could be adjusted every $1/f_s$, in order to protect the quality
1351: of the laser noise cancellation against drifting armlengths.  This
1352: requires an adequate model of the spacecraft orbits, which could be
1353: updated as needed from spacecraft ranging data.
1354: 
1355: \subsection{Sampling time jitter}
1356: 
1357: The sampling times of all the measurements needed for synthesizing the
1358: TDI combinations will not be constant, due to the intrinsic timing
1359: jitters of the digitizing systems (USOs and phasemeters). Within the
1360: digitizing system, the USO is expected to be the dominant source of
1361: time jittering in the sampled data. Presently existing, space
1362: qualified, USO can achieve an Allan standard deviation of about
1363: $10^{-13}$ for integration times from $1$ to $10000$ seconds.  This
1364: timing stability translates into a time jitter of about $10^{-13}$
1365: seconds over a period of $1$ second. A perturbative analysis including
1366: the three sampling time jitters due to the three clocks shows that any
1367: laser phase fluctuations remaining in the four TDI generators will
1368: also be proportional to the sampling time jitters. Since the latter
1369: are approximately four orders of magnitude smaller than the armlength
1370: and clocks synchronization accuracies derived earlier, we conclude
1371: that the magnitude of laser noise residual into the TDI combinations
1372: due to the sampling time jitters can be made negligible.
1373: 
1374: \subsection{Data digitization and bit-accuracy requirement}
1375: 
1376: As shown in figure 5, the maximum of the ratio of the laser noise and
1377: of the secondary noises phase fluctuation amplitudes occurs at the
1378: lower end of the LISA bandwidth, and is $\sim 10^{10}$ at 0.1 mHz.
1379: This corresponds to the minimum dynamic range for the phasemeters to
1380: correctly measure the laser fluctuations and the weaker signals
1381: simultaneously. An additional safety factor of $\sim 10$ should be
1382: sufficient to avoid saturation if the noises are well described by
1383: Gaussian statistics.
1384: 
1385: \begin{figure}[httb]
1386: \centering 
1387: \includegraphics[width=6in, angle=0]{Fig5.ps}
1388: \caption{Phase fluctuations spectra are plotted versus Fourier
1389:   frequency for: (upper curve) raw laser noise having spectral density
1390:   $2.3 \times 10^{-1} \ f^{- 8/3} + 1.4 \times 10^{-9} \ f^{-27/5} \ 
1391:   {\rm cycles}^2 \ {\rm Hz}^{-1}$, and (lower curves) residual noises
1392:   entering into the various TDI combinations.  The armlength has been
1393:   assumed to be equal to $ L = 16.67 \ {\rm sec}$.}
1394: \label{DR} 
1395: \end{figure}
1396: 
1397: In terms of requirements on the digital signal processing subsystem,
1398: this dynamic range implies that approximately $36$ bits are needed
1399: when combining the signals in TDI, only to bridge the gap between
1400: laser frequency noise and the other noises and gravitational wave
1401: signals. More bits might be necessary to provide enough information to
1402: efficiently filter the data when extracting weak gravitational wave
1403: signals embedded into noise.
1404: 
1405: \section{Summary and Conclusions}
1406: 
1407: A comparative analysis of different schemes for implementing
1408: Time-Delay Interferometry with LISA has been presented. In particular,
1409: we have shown that the master-slave configuration is capable of
1410: generating the entire space of interferometric combinations identical
1411: to that derived by using the one-way scheme. This was done under the
1412: assumption that the noise from the optical transponders was
1413: negligible. Our analysis of the phase-locking control systems forming
1414: the optical transponders shows that this is a valid assumption,
1415: indicating that the noise introduced can be expected to be $\approx 1
1416: \ \mu {\rm cycle}/\sqrt{\rm Hz}$.
1417: 
1418: A comparison of the hardware required for each scheme shows that the
1419: subsystems needed are almost identical, with the only difference being
1420: the number of frequency stabilization systems. This difference is
1421: perhaps not significant when redundancy options are considered.  The
1422: main disadvantage of the one-way method is that the laser frequencies
1423: might be offset by several hundred megahertz, given the currently
1424: envisioned optical-cavity-based frequency stabilization systems. This
1425: places challenging constraints on the photoreceiver bandwidth and
1426: bandwidth stability. On the other hand, the master-slave configuration
1427: has no such problem, allowing the beat-note on the photoreceiver to be
1428: the minimum determined by the Doppler shift.  However, there may be
1429: concerns with the non-local nature of the phase locking system, since
1430: its performance could be influenced, for example, by pointing
1431: stability (which also has implications on lock acquisition). Further
1432: studies on these issues should be performed.
1433: 
1434: Given the similarities between the two schemes, in principle either
1435: operational mode could be implemented without major implications on
1436: the hardware configuration. Ultimately detailed engineering studies
1437: will identify the preferred approach.
1438: 
1439: A derivation of the armlengths and clocks synchronization accuracies,
1440: as well as a determination of the precision requirement on the
1441: sampling time jitter, have also been derived. We found that an
1442: armlength accuracy of about 16 meters, a synchronization accuracy of
1443: about $50$ ns, and the time jitter due to a presently existing Ultra
1444: Stable Oscillator will allow the suppression of the frequency
1445: fluctuations of the lasers below the level identified by the secondary
1446: noise sources. A new procedure for sampling the data in such a way to
1447: avoid the problem of having time shifts that are not integer multiples
1448: of the sampling time was also presented, addressing one of the concerns
1449: about the implementation of Time-Delay Interferometry.
1450: 
1451: \section*{Acknowledgments}
1452: 
1453: We would like to thank Dr. Alex Abramovici for useful discussions on
1454: phase locking control systems, and Dr. Frank B. Estabrook for several
1455: stimulating conversations. This research was performed at the Jet
1456: Propulsion Laboratory, California Institute of Technology, under
1457: contract with the National Aeronautics and Space Administration.
1458: 
1459: \appendix
1460: \section{Fundamental limit of the LISA phase transponder}
1461: \label{QMC}
1462: 
1463: The following is a derivation of the noise added by the optical phase
1464: measurement. This calculation only considers noise added by the
1465: detection system, it does not include the 20~$\mu$cycles$/\sqrt{\rm
1466:   Hz}$ due to the optical-path noise, and it is purely quantum
1467: mechanical \cite{Bachor}. The reason for performing the calculation in
1468: quantum mechanical terms is to highlight that the measurement process
1469: itself does not add shot noise and that, in principle, a perfect phase
1470: measurement of the field could be made.
1471: 
1472: Let us consider the optical configuration shown in figure
1473: \ref{diag1}(b), which is equivalent to the optical arrangement for the
1474: LISA interferometer (figure \ref{diag1}(a)). Let the annihilation
1475: operators for the local and distant lasers be $\hat{a}$ and $\hat{b}$
1476: respectively. We can represent these operators as the sum of an
1477: average (complex number) component and an operator component
1478: representing the field fluctuations of zero mean values:
1479: \begin{eqnarray}
1480: \hat{a}&=&\alpha+\delta\hat{a} \ ,
1481: \label{defa1}
1482: \\
1483: \hat{b}&=&\beta+\delta\hat{b} \ .
1484: \label{defb1}
1485: \end{eqnarray}
1486: In Eqs. (\ref{defa1}, \ref{defb1}) we define $\alpha$ and $\beta$ to
1487: be real numbers. We will assume throughout these calculations that
1488: $\delta\hat{a},\delta\hat{b}\ll\alpha,\beta$ and so terms that are of
1489: second order in these quantities will be ignored. This approximation
1490: holds even for the low intensities found in the LISA interferometer.
1491: \begin{figure}[htb]
1492: \vspace{.2in}
1493: \centerline {
1494: \includegraphics[width=4in]{Fig6.ps}
1495: }
1496: \caption{(a) Simplified optical arrangement for phase locking in
1497:   LISA. (b) Equivalent optical layout for modeling purposes.
1498:   $\epsilon$ is the intensity reflectivity of the beam splitter.}
1499: \vspace{.2in}
1500: \label{diag1}
1501: \end{figure}
1502: 
1503: In an offset phase locking system the field $\hat{b}$ is offset from
1504: field $\hat{a}$ by a radial frequency $\omega_{b}$. In this case
1505: equations \ref{defa1} and \ref{defb1} become
1506: \begin{eqnarray}
1507: \hat{a}&=&\alpha+\delta\hat{a} \label{defa2}\\
1508: \hat{b}&=&\beta{\rm e}^{i\omega_{b} t}+\delta\hat{b}{\rm
1509:   e}^{i\omega_{b} t} \label{defb2} \ .
1510: \end{eqnarray}
1511: The annihilation operator for the field at the photoreceiver,
1512: $\hat{c}$, will contain some fraction of $\hat{a}$ and $\hat{b}$,
1513: \begin{eqnarray}
1514: \hat{c}&=&\sqrt{1-\epsilon}\,\hat{a}+\sqrt{\epsilon}\,\hat{b} \ .
1515: \end{eqnarray}
1516: 
1517: The photoreceiver measures a quantity proportional to the photon
1518: number of this field, $n_c=\hat{c}^\dagger\hat{c}$, 
1519: \begin{eqnarray}
1520: \hat{c}^\dagger\hat{c}
1521: &=&(1-\epsilon) \left[\alpha^2+\alpha(\delta\hat{a}+\delta\hat{a}^\dagger)\right] 
1522: +\epsilon\left[\beta^2+\beta(\delta\hat{b}+\delta\hat{b}^\dagger)\right]   \\
1523: &&
1524: +\sqrt{\epsilon}\sqrt{1-\epsilon}\left[\alpha\beta({\rm e}^{i\omega_{b} t}+{\rm e}^{-i\omega_{b} t}) 
1525: +\alpha(\delta\hat{b}{\rm e}^{i\omega_{b} t}+\delta\hat{b}^\dagger{\rm e}^{-i\omega_{b} t})
1526: +\beta(\delta\hat{a}{\rm e}^{-i\omega_{b} t}+\delta\hat{a}^\dagger{\rm e}^{i\omega_{b} t})\right] \nonumber\\
1527: &=&
1528: (1-\epsilon) \left[\alpha^2+\alpha(\delta\hat{a}+\delta\hat{a}^\dagger)\right] 
1529: +\epsilon\left[\beta^2+\beta(\delta\hat{b}+\delta\hat{b}^\dagger)\right] 
1530: \nonumber  \\
1531: &&
1532: +2\sqrt{\epsilon}\sqrt{1-\epsilon}\Big[ (\alpha\beta +
1533: \alpha(\delta\hat{b}+\delta\hat{b}^\dagger)+\beta(\delta\hat{a}+\delta\hat{a}^\dagger))\cos(\omega_{b} t) \nonumber \\
1534: &&+
1535: (\alpha(i\delta\hat{b}-i\delta\hat{b}^\dagger)-\beta(i\delta\hat{a}-i\delta\hat{a}^\dagger))\sin(\omega_{b}
1536: t)\Big] \ . 
1537: \label{complex1}
1538: \end{eqnarray}
1539: To simplify the notation, we define the quadrature operators which
1540: represent the fluctuations in the amplitude ($\delta\hat{X}^{+}$) and
1541: phase ($\delta\hat{X}^{-}$) quadratures of the operators $\hat{a}$ and
1542: $\hat{b}$,
1543: \begin{eqnarray}
1544: \delta\hat{X}^{+}_{a}&=&(\delta\hat{a}+\delta\hat{a}^\dagger) \ , \\
1545: \delta\hat{X}^{+}_{b}&=&(\delta\hat{b}+\delta\hat{b}^\dagger) \ ,  \\
1546: \delta\hat{X}^{-}_{a}&=&i(\delta\hat{a}-\delta\hat{a}^\dagger) \ ,  \\
1547: \delta\hat{X}^{-}_{b}&=&i(\delta\hat{b}-\delta\hat{b}^\dagger) \ .
1548: \end{eqnarray}
1549: Each of these quantities is an observable of unit variance for a
1550: coherent state (idealized laser). Substituting these expressions into
1551: equation \ref{complex1} we obtain,
1552: \begin{eqnarray}
1553: \hat{c}^\dagger\hat{c}&=&
1554: (1-\epsilon)\left[\alpha^2+\alpha\delta\hat{X}^{+}_{a}\right]+\epsilon\left[\beta^2+\beta\delta\hat{X}^{+}_{b}\right] 
1555: \nonumber \\
1556: &&+ 2\sqrt{\epsilon}\sqrt{1-\epsilon}\left[(\alpha\beta 
1557: +\alpha\delta{X}^{+}_{b}+\beta\delta{X}^{+}_{a})\cos(\omega_{b}
1558: t)+(\alpha\delta{X}^{-}_{b}-\beta\delta{X}^{-}_{a})\sin(\omega_{b}
1559: t)\right] \ . 
1560: \end{eqnarray}
1561: This equation contains three terms which can now be identified. The
1562: first two terms arise from the intensities of fields $\hat{a}$ and
1563: $\hat{b}$ respectively. These terms are non-interferometric in nature
1564: and contain the intensity fluctuations of the individual input beams
1565: scaled by the efficiency of coupling to the photoreceiver (beam
1566: splitter ratio). The third term represents the interference between
1567: the two fields and provides a beat note at frequency $\omega_{b}$, the
1568: difference frequency of the two fields. This beat note itself has two
1569: parts, an intensity noise part oscillating as $\cos(\omega_b t)$, and
1570: a phase difference part oscillating as $\sin(\omega_b t)$.
1571: 
1572: The phase difference can be obtained, for example, by using a mixer to
1573: demodulate the beat note down to zero. Mathematically, this is a
1574: multiplication by $\sin(\omega_{b} t)$. Terms with a cosine multiplier
1575: will only exhibit higher harmonics whereas terms with a sine
1576: multiplier will mixed down to base band frequencies.
1577: 
1578: After low-pass filtering the mixer output, we obtain the following
1579: expression for the error signal
1580: \begin{eqnarray}
1581: V_{e}& \equiv & \frac{(1-\epsilon)}{2}   \alpha\delta\hat{X}^{+}_{a,\omega_{b}} 
1582: + \frac{\epsilon}{2} \beta\delta\hat{X}^{+}_{b,\omega_{b}}
1583: +
1584: \sqrt{\epsilon}\sqrt{1-\epsilon}(\alpha\delta{X}^{-}_{b}-\beta\delta{X}^{-}_{a})
1585: \ ,
1586: \end{eqnarray}
1587: where we have ignored the noise of the oscillator used in the mixing
1588: process. The quantities $\hat{X}^{+}_{a,\omega_{b}}$ and
1589: $\hat{X}^{+}_{b,\omega_{b}}$ are the amplitude quadrature operators
1590: evaluated at the offset frequency. The factor of 1/2 in the
1591: coefficients of these terms arises as we have only taken the sine
1592: component of the intensity noise at the heterodyne frequency. For
1593: heterodyne frequencies of 10~MHz and greater the intensity noise is
1594: shot noise limited thus
1595: $\langle(\hat{X}^{+}_{a,\omega_{b}})^2\rangle\approx
1596: \langle(\hat{X}^{+}_{b,\omega_{b}})^2\rangle\approx 1$. The important
1597: point is that the error signal is proportional to the intensity noise
1598: of each beam and the relative phase noise of the two lasers.
1599: 
1600: Assuming perfect phase locking, the error signal $V_e$ is driven to
1601: zero by actuating on the phase of $\hat{a}$. Setting $V_e=0$ we find
1602: that the controller will attempt to force the phase quadrature
1603: fluctuations to be,
1604: \begin{eqnarray}
1605: \delta{X}^{-}_{a}&=&\frac{\alpha}{\beta}\delta{X}^{-}_{b}+\frac{\sqrt{1-\epsilon}  
1606: \alpha\delta\hat{X}^{+}_{a,\omega_{b}}}{2 \sqrt{\epsilon}\beta}  
1607: + \frac{\sqrt{\epsilon}\delta\hat{X}^{+}_{b,\omega_{b}}}{2
1608:   \sqrt{1-\epsilon}} \label{perfectPL} \ .
1609: \end{eqnarray}
1610: 
1611: Ultimately, what is of interest is the difference between the phases
1612: of the fields $\hat{d}$ and $\hat{a}$. The annihilation operator for
1613: the outgoing beam, $\hat{d}$, is also made up of a linear combination
1614: of $\hat{a}$ and $\hat{b}$
1615: \begin{eqnarray}
1616: \hat{d}&=&\sqrt{\epsilon}\,\hat{a}-\sqrt{1-\epsilon}\,\hat{b} \ ,
1617: \end{eqnarray}
1618: which implies the following phase quadrature fluctuations of $\hat{d}$, $\delta
1619: \hat{X}^-_d$
1620: \begin{eqnarray}
1621: \delta X^-_d&=&\sqrt{\epsilon}\delta \hat{X}^-_a
1622: -\sqrt{1-\epsilon}\delta \hat{X}^-_b \ .
1623: \end{eqnarray}
1624: Under the operational configuration of perfect phase locking, we can
1625: substitute into the equation above the expression for $\delta{X}^{-}_{a}$ given in equation
1626: \ref{perfectPL}
1627: \begin{eqnarray}
1628: \delta
1629: X^-_d&=&(\sqrt{\epsilon}\frac{\alpha}{\beta}-\sqrt{1-\epsilon})\delta\hat{X}^{-}_{b}+\frac{\sqrt{1-\epsilon}
1630:   \alpha\delta\hat{X}^{+}_{a,\omega_{b}}}{2 \beta}
1631: +\frac{\epsilon\delta\hat{X}^{+}_{b,\omega_{b}}}{2\sqrt{1-\epsilon}} \ .
1632: \label{complex}
1633: \end{eqnarray}
1634: Since the power of the laser A (proportional to $\alpha^2$) will be a
1635: factor of $\approx10^{8}$ larger than the power of laser B (
1636: proportional to $\beta^2$), and also that $\langle(\delta
1637: \hat{X}^+_b)^2\rangle\sim 1$, i.e.  the signal laser intensity is
1638: approximately shot noise limited at the heterodyne frequency, we find
1639: that Eq. (\ref{complex}) can be approximated as follows
1640: \begin{eqnarray}
1641: \delta X^-_d&\simeq&\frac{\alpha}{\beta}\left[\sqrt{\epsilon}\delta
1642:   \hat{X}^-_b+\frac{\sqrt{1-\epsilon}\delta
1643:     \hat{X}^+_{a,\omega_{b}}}{2}\right] \ .
1644: \end{eqnarray}
1645: 
1646: The quantity of interest is the phase fluctuations in radians. To
1647: compare the phase fluctuations between the two fields we need to
1648: normalize the phase quadrature operator by the square root of the
1649: average photon number. For $\hat{d}$, this means dividing by
1650: $\sqrt{\epsilon}\alpha$
1651: \begin{eqnarray}
1652: \delta\phi_b&=&\frac{\delta\hat{X}^-_b}{\beta} \ , \\
1653:  \delta\phi_d&=&\frac{\delta\hat{X}^-_b}{\beta}
1654:  +\frac{\sqrt{1-\epsilon}}{2\sqrt{\epsilon}}\frac{\delta
1655:  \hat{X}^+_{a,\omega_{b}}}{\beta} \ .
1656: \end{eqnarray}
1657: Thus the phase of the incoming beam will differ from the phase of the
1658: outgoing beam by the following amount
1659: \begin{eqnarray}
1660:  \delta\phi_d-\delta\phi_b&=&\frac{\sqrt{1-\epsilon}}
1661: {2\sqrt{\epsilon}}\frac{\delta \hat{X}^+_{a,\omega_{b}}}
1662: {\beta} \ ,
1663: \end{eqnarray}
1664: while the root-mean-squared value of the phase error can be written
1665: as
1666: \begin{eqnarray}
1667: \sigma_\phi=\sqrt{\frac{1-\epsilon}{4\epsilon}\frac{\langle|\delta\hat{X}^+_{a,\omega_b}|^2\rangle}{\bar{n}_b}}
1668: \ .
1669: \end{eqnarray}
1670: Here $\bar{n}_b = \beta^2$ is the average number of photons in the
1671: weak signal laser, and $\langle|\delta\hat{X}^+_{a,\omega_b}|^2
1672: \rangle$ is the variance of the local oscillator intensity
1673: fluctuations relative to the variance of quantum noise. Thus the error
1674: depends on two parameters, the beam splitter ratio and the intensity
1675: noise of the local oscillator. Assuming the local laser intensity is
1676: shot noise limited at the modulation frequency and a beam splitter
1677: ratio of approximately 100:1 ($\epsilon=0.99$) gives a phase error
1678: with a standard deviation, $\sigma_\phi$, of approximately $5\%$ of
1679: the shot noise limit or less than 1~$\mu$cycle$/\sqrt{\rm Hz}$ error.
1680: This error is therefore much less than the 20~$\mu$cycles$/\sqrt{\rm
1681:   Hz}$ optical path noise.
1682: 
1683: If necessary, this source of error could be removed by altering the
1684: detection system to add a second detector allowing subtraction and
1685: consequent cancellation of the intensity noise of the local oscillator
1686: laser. Cancellation factors of 100 are readily achievable, albeit with
1687: a slight increase in system complexity, effectively removing this
1688: error contribution entirely.
1689: 
1690: \section{Interpolation error for generating shifted data points}
1691: 
1692: Let $g(t)$ be the true signal.  It is sampled at intervals $\Delta t =
1693: 1/f_s$, where $f_s$ is the sampling frequency, to produce the discrete
1694: data $g[n] = g(t + n / f_s)$, $n=...,-1,0,1,...$.  Consider the
1695: problem of estimating $g(t)$ at some time which falls in between two
1696: sampling times, i.e. at time $t$ where $0 < |t - n_0/f_s| < \Delta t$,
1697: for $n_0$ the value of $t/\Delta t$ rounded to the nearest integer.
1698: It is well known that for an infinitely long dataset, this estimation
1699: can be done without error using the Shannon formula \cite{12},
1700: assuming that the signal has zero power above the Nyquist frequency
1701: ($f_s/2$).  The error in the estimation with a finite digital filter
1702: can be approximated using a truncated version of the Shannon formula.
1703: This digital filter might not be the best one for all signals, but it
1704: should be sufficiently close to the optimal filter.  The estimated
1705: function is given by
1706: \begin{equation}
1707: g_N(t) = \sum_{n=-N}^N g[n + n_0] \;{\rm sinc}(f_s t - n - n_0) \ ,
1708: \end{equation}
1709: for a digital filter of length $2N+1$, and where $\;{\rm
1710:   sinc}(x)=\sin(\pi x)/\pi x$. The estimation error is $e_N(t) = g(t)
1711: - g_N(t)$.
1712: 
1713: Changing the sampling frequency will not improve the function
1714: estimator for a fixed digital filter size: a lower sampling frequency
1715: would be insufficient to represent the high frequency signal
1716: components, and a higher sampling frequency would reduce the size of
1717: the interval over which the function is sampled to build the
1718: estimator.
1719: 
1720: Assuming $g(t)$ to be a wide sense stationary stochastic process
1721: \cite{13} with autocorrelation function $R_g(|t_1-t_2|) = R_g(t_1,t_2)
1722: = E[g(t_1) g(t_2)]$ ($E[]$ denotes the expectation value), it follows
1723: that the autocorrelation function of the estimation error is equal to
1724: \begin{eqnarray}
1725: R_{e_N}(t_1,t_2) = R_g(t_1 - t_2) + \nonumber \\
1726: \sum_{m,n=-N}^N R_g\left(\frac{n+n_1-m-n_2}{f_s}\right)\;{\rm
1727:   sinc}(f_s t_1 - n - n_1)
1728: \;{\rm sinc}(f_s t_2 - m - n_2) \nonumber \\
1729: - \sum_{n=-N}^N R_g(t_1 - n/f_s - n_2/f_s)\;{\rm sinc}(f_s t_2 - n - n_2) \nonumber \\
1730: - \sum_{n=-N}^N R_g(t_2 - n/f_s - n_2/f_s)\;{\rm sinc}(f_s t_1 - n - n_1),
1731: \end{eqnarray}
1732: where $n_1$ ($n_2$) is the value of $t_1/\Delta t$ ($t_2/\Delta t$)
1733: rounded to the nearest integer.  This equation shows that $e_N(t)$ is
1734: not wide sense stationary; however, it is wide sense cyclo-stationary,
1735: since $R_{e_N}(t_1,t_2) = R_{e_N}(t_1 + m \Delta t, t_2 + m \Delta t)$
1736: for every integer $m$.  This is just a consequence of the error
1737: varying quasi-periodically with the interpolation time; it is zero
1738: when $t$ matches a sample time, maximum at the middle between two
1739: sample times, etc.
1740: 
1741: A fair estimate of the estimation error magnitude can be obtained by
1742: considering the stochastic process $\bar{e}_N(t) = e_N(t + \theta)$,
1743: where $\theta$ is a random variable uniformly distributed in $[0,
1744: \Delta t]$.  $\bar{e}_N(t)$ is wide sense stationary, and its
1745: autocorrelation function is \cite{13}
1746: \begin{equation}
1747: R_{\bar{e}_N}(\tau) = f_s \int_0^{\Delta t} R_{e_N}(t+\tau,t) dt \ .
1748: \end{equation}
1749: The Fourier transform of $R_{\bar{e}_N}(\tau)$ gives an estimate of
1750: the spectrum of the noise induced by the digital filters interpolation
1751: errors.  In particular, $R_{\bar{e}_N}(0)$ is the broadband standard
1752: deviation of the noise.
1753: 
1754: Taking $g(t)$ to be a laser phase noise with a power spectral density
1755: that scales like $1/f^2$, and restricting attention to the frequency
1756: range 0.1 mHz $< f <$ 1 Hz, one can calculate numerically that
1757: $R_{\bar{e}_N}(0) / R_g(0) = 9\times 10^{-7}$ for $N=10$. Therefore, a
1758: filter with $N=10$ is good enough only to produce a broadband error on
1759: the shifted time series that is $\sim 3$ orders of magnitude smaller
1760: than the laser phase noise amplitude. Changing $N$ in the numerical
1761: integrations shows that $R_{\bar{e}_N}(0)$ scales roughly like $1/N$.
1762: This implies that it would be impossible to use digital filters on a
1763: slowly sampled time series to achieve the levels of noise cancellation
1764: required by the TDI combinations.
1765: 
1766: \begin{thebibliography} {}
1767:   
1768: \bibitem{1} P. Bender, K. Danzmann, \& the LISA Study Team, {\it{Laser
1769:       \ Interferometer \ Space \ Antenna \ for \ the \ Detection \ of
1770:       \ Gravitational \ Waves, \ Pre-Phase \ A \ Report}}, $\bf{MPQ
1771:     233}$ (Max-Planck-Instit\"ut f\"ur Quantenoptik, Garching), July
1772:   1998.
1773:   
1774: \bibitem{2} M. Tinto, F.B. Estabrook, \& J.W. Armstrong, {\it Phys.
1775:     Rev. D}, {\bf 65}, 082003 (2002).
1776:   
1777: \bibitem{3} F.B. Estabrook, M. Tinto, \& J.W. Armstrong, {\it Phys.
1778:     Rev. D}, {\bf 62}, 042002 (2000).
1779:   
1780: \bibitem{4} J. Gowar, {\it Optical Communication Systems},
1781:   (Prentice/Hall International, 1984), pp. 136-140.
1782:   
1783: \bibitem{5} J.W. Armstrong, F.B. Estabrook, \& M. Tinto, {\it Ap. J.},
1784:   {\bf 527}, 814 (1999).
1785:   
1786: \bibitem{6} S.V. Dhurandhar, K.R. Nayak, \& J.-Y. Vinet {\it Phys.
1787:     Rev. D}, {\bf 65}, 102002 (2002).
1788:   
1789: \bibitem{7} M. Tinto, \& J.W. Armstrong {\it Phys. Rev. D}, {\bf 59},
1790:   102003 (1999).
1791:   
1792: \bibitem{8} D.A. Shaddock, {\it Carrier-Sideband USO noise
1793:     cancellation}, available at the following URL address:
1794:     $http://www.srl.caltech.edu/LISA/TDI\_WP/ShaddockUSO.pdf$, 
1795: unpublished.
1796:   
1797: \bibitem{Abramovici} A. Abramovici and J. Chapsky, {\it Feedback
1798:     Control Systems a Fast-Track Guide for Scientists and Engineers},
1799:   Kluwer Academic Publishers, Boston (2000).
1800:   
1801: \bibitem{8bis} O. Jennrich, R.T. Stebbins, P.L. Bender, and S. Pollack
1802:   {\it Clas. Quantum Grav.}, {\bf 18}, 4159, (2001).
1803:   
1804: \bibitem{8SHA} D.A. Shaddock, {\it Digital I-Q Phasemeter by use of
1805:     under-sampling}. In Preparation.
1806:   
1807: \bibitem{McNamaraPL} P.W. McNamara, H. Ward, and J. Hough, {\it Laser
1808:     Phase Locking for LISA: Experimental Status}. In: {\it Proceedings
1809:     of the Second International LISA Symposium on the Detection and
1810:     Observation of Gravitational Waves in Space}, ed. W.M. Folkner,
1811:     AIP Conference Proceedings 456, Woodbury, New York, (1998).
1812:   
1813: \bibitem{Ye} J. Ye and J.L. Hall, {\it Opt. Lett.}, {\bf 24}, 1838,
1814:   (1999)
1815:   
1816: \bibitem{PDH} R.W.P. Drever, J.L. Hall, F.V. Kowalski, J.Hough, 
1817: G.M. Ford, A.J. Munley, and H.Ward, {\it Appl. Phys. B}, {\bf 31},
1818: 97 (1983). 
1819:   
1820: \bibitem{8tris} P.W. McNamara, H. Ward, J. Hough, and D. Robertson,
1821:   {\it Clas. Quantum Grav.}, {\bf 14}, 1543, (1997).
1822:   
1823: \bibitem{9} W.M. Folkner, F. Hechler, T.H. Sweetser, M.A. Vincent, and
1824:   P.L. Bender, {\it Clas. Quantum Grav.}, {\bf 14}, 1405, (1997).
1825:    
1826: \bibitem{10} L. Schmidt, and M. Miranian, {\it Common-View Time
1827:     Transfer Using Multi-Channel GPS Receivers}. In: {\it Proceedings
1828:     of the 29th Annual Precise Time and Time Interval (PTTI) Meeting},
1829:   ed. L.A. Breakiron, United States Naval Observatory, (1997).
1830:   
1831: \bibitem{11} R.W. Hellings, {\it Phys. Rev. D}, {\bf 64}, 022002
1832:   (2001)
1833:   
1834: \bibitem{Bachor} H.A. Bachor and T.C. Ralph, {\it A guide to
1835:     experiments in Quantum Optics}, John Wiley \& Sons, New York (2003)
1836:   
1837: \bibitem{12} R.J. Marks, {\it Introduction to Shannon sampling and
1838:     interpolation theory}, Springer-Verlag, Berlin (1991).
1839:   
1840: \bibitem{13} A. Papoulis, {\it Probability, Random Variables, and
1841:     Stochastic Processes}, McGraw-Hill, New York, (1991).
1842: 
1843: \end{thebibliography}
1844: 
1845: \end{document}
1846: