0808.3400/ms.tex
1: % $Id: comprehensive.tex,v 1.19 2008/08/25 15:39:40 garyb Exp $	
2: \documentclass[11pt,preprint,flushrt]{aastex}
3: %\documentclass{emulateapj}
4: \renewcommand{\topfraction}{0.8}
5: 
6: \def\eqq#1{Equation~(\ref{#1})}
7: \newcommand\etal{{\it et al.\/}}
8: \newcommand\eg{{\it e.g.\ }}
9: \newcommand\ie{{\it i.e.\ }}
10: 
11: \begin{document}
12: 
13: \slugcomment{CVS \$Revision: 1.19 $ $ \$Date: 2008/08/25 15:39:40 $ $}
14: 
15: \title{Comprehensive Two-Point Analyses of Weak Gravitational Lensing Surveys} 
16: \author{Gary M. Bernstein}
17: \affil{Dept. of Physics and Astronomy, University of Pennsylvania,
18: Philadelphia, PA 19104}
19: \email{garyb@physics.upenn.edu}
20: 
21: \begin{abstract}
22: We present a framework for analyzing weak gravitational lensing survey
23: data, including lensing and source-density observables, plus spectroscopic
24: redshift calibration data.  All two-point observables are predicted in
25: terms of parameters of a perturbed Robertson-Walker metric, making
26: the framework independent of the models for gravity, dark energy, or
27: galaxy properties.  For Gaussian fluctuations the 2-point model determines the
28: survey likelihood function and allows Fisher-matrix forecasting.  The
29: framework 
30: includes nuisance terms for the major systematic errors: shear
31: measurement errors, magnification bias and redshift calibration
32: errors, intrinsic galaxy alignments, and inaccurate theoretical
33: predictions.   We propose flexible parameterizations of the many
34: nuisance parameters related to galaxy bias and intrinsic alignment.
35: For the first time we can integrate many different observables and
36: systematic errors into a single analysis.
37: As a first application of this framework, we demonstrate that:
38: uncertainties in power-spectrum theory cause very minor degradation to
39: cosmological information content; nearly all useful information
40: (excepting baryon oscillations) is
41: extracted with $\approx 3$ bins per decade of angular scale; and the rate
42: at which galaxy bias varies with redshift substantially influences the
43: strength of cosmological inference.  The framework will permit careful
44: study of the interplay between numerous observables, systematic
45: errors, and spectroscopic calibration data for large weak-lensing surveys.
46: 
47: \end{abstract}
48: 
49: \keywords{gravitational lensing; cosmological parameters; relativity}
50: 
51: \section{Introduction}
52: Weak gravitational lensing of background sources can produce
53: exceptionally strong constraints on cosmological parameters and tests
54: of General Relativity.  Initial analyses considered the 2-point
55: correlation function (or, equivalently, power spectrum) of the shear
56: pattern induced on a single population of background galaxies
57: \citep{Jordi, Kaiser92, Blandford91}.  A wealth of new statistics,
58: however, have been suggested as more powerful means to extract
59: information from weak lensing (WL): cross-power spectra of multiple
60: source populations with distinct redshift distributions
61: (a.k.a. ``tomography'') \citep{Hu}; the correlation of shear with
62: foreground galaxy clusters \citep{JainTaylor}, or more generally the
63: cross-correlation of lensing shear with the galaxy distribution
64: \citep{BJ04, Zhang}; joint analyses of density-density,
65: density-shear, and shear-shear correlations in an imaging survey
66: \citep{HJ04}; cross-correlation of magnification as well as shear
67: \citep{Jainmag}; use of the CMB \citep{HO02,HS03} or recombination-era 21~cm
68: signals \citep{MW07,Pen04,ZZ06} as source planes; cross-correlation of source
69: density or shear with a distinct spectroscopic galaxy survey
70: population \citep{Newman, SKZC}; and the use of 3-point statistics
71: \citep{TJ04} or statistics such as peak counts \citep{Hennawi,Wang,Laura} to
72: move beyond 2-point information.  Each of these potential innovations
73: has been individually analyzed and shown to improve cosmological
74: constraints.  The first goal of this paper is to consider the {\em
75:   simultaneous} use of all of these observable statistics: can we
76: forecast the cosmological information that they will yield
77: collectively in future surveys?  Can we start to develop a framework
78: in which all these signals could be analyzed simultaneously in a real
79: experiment? 
80: 
81: In parallel with the increasing variety of proposed WL signals, the
82: community has identified a series of potential astrophysical and
83: instrumental non-idealities in WL data which, if ignored, would lead
84: to substantial systematic errors in the inferred cosmology.  These
85: include: finite accuracy in our ability to predict the deflecting mass
86: power spectrum due to nonlinearities \citep{JS97} and baryonic physics
87: \citep{Zhan06, Jing}; intrinsic alignments (IA) between galaxy shapes
88: \citep{CroftMetzler} and between galaxy shapes and the local mass distribution
89: \citep{Hirata2} that are not induced by lensing; multiplicative
90: ``shear calibration'' errors in the derivation of lensing shear from
91: galaxy images \citep{Ishak04,HTBJ}; additive ``spurious shear'' due
92: to uncorrected PSF 
93: ellipticity or other imaging systematics \citep{HTBJ,AmaraRefregier}; 
94: and errors in
95: the assignment of redshifts to the source populations
96: \citep{MaHutererHu}.  The impact of these systematic-error sources on
97: cosmological inferences have been analyzed by different means, but a
98: second goal of this paper is to produce a comprehensive forecast that
99: considers the presence of them all simultaneously. 
100: 
101: Previous work has shown that these multiple sources of information and
102: systematic in WL surveys can interact in interesting ways.  For
103: example, in the presence of tomographic data, many systematics are
104: readily distinguishable from cosmological signals and can hence be
105: diagnosed and corrected internally to a survey; this approach is
106: called {\em self-calibration} \citep{HTBJ}.  It has also been shown
107: that combining galaxy density and lensing correlations can lead to
108: self-calibration of shear calibration errors \citep{BJ04} and the
109: uncertainties in galaxy biasing \citep{HJ04, Zhan06}.  Intrinsic
110: alignments of galaxies can be diagnosed and corrected if tomographic
111: information is available \citep{KingSchneider}, however this places
112: substantially greater demands on the precision and accuracy of
113: redshift assignment than would otherwise be needed \citep{BridleKing}.
114: These investigations raise important practical questions: will the
115: self-calibration techniques continue to succeed when we attempt to
116: simultaneously self-calibrate several different systematic errors?  Do
117: cross-correlation techniques reduce uncertainties in redshift
118: distributions to negligible levels, or is it necessary to make a
119: complete spectroscopic redshift survey of some size to measure
120: redshift distributions directly \citep{MaBernstein}?  This paper will
121: present a formalism through which all these questions can be answered,
122: but we defer to later papers the application of the framework to these
123: issues. 
124: 
125: A third goal of this work is to describe the constraints by WL in a
126: language that is not tied to a specific cosmological model.  Most
127: forecasts for WL survey constraints are done within the context of a
128: Universe that has homogeneous dark energy with equation of state
129: $w=w_0 + w_a(1-a)$.  Projecting the WL experiment onto this model
130: gives concrete predictions, but obscures what the WL is really
131: measuring.  So the analysis framework presented here will be {\em
132:   dark-energy agnostic}, meaning that no specific model is assumed.
133: We will be very explicit about the assumptions made in the analysis
134: and try to keep them to a minimum.  In fact a great strength of WL
135: experiments are their ability to test General Relativity itself, so we
136: seek an analysis method that is general enough to incorporate such
137: tests. Similar to the approach of \citet{KST}, our analysis results in
138: constraints on the distance and growth functions $D(z)$ and
139: $g_\phi(z)$, without reference to the particular dark-energy or
140: gravity modifications that might cause deviations from $\Lambda$CDM.
141: 
142: In the following section we describe a ``kitchen-sink'' formalism for
143: WL survey observables that allows the incorporation of all suggested
144: 2-point statistics and very general treatments of nearly all proposed
145: systematic errors.  In \S\ref{speclike} we give a likelihood function
146: and Fisher matrix for an unbiased spectroscopic redshift survey of
147: source galaxies.  Then we briefly describe a software implementation
148: of the lensing and spectroscopy likelihood calculations.
149: We describe our model for the evolution of the lensing-potential power
150: spectrum in \S\ref{powermodel}, and \S\ref{nuisance} we describe
151: generic models used for the nuisance functions required in the
152: lensing-survey analysis. In \S\ref{tuneup} we use the implementation
153: of these methods to investigate the proper choices for the bin sizes
154: and grid spacings needed to turn the lensing analysis into a tractable
155: finite-dimensional problem.  Further application of the framework to
156: survey forecasting will be done in future papers.
157: 
158: An earlier version of this WL analysis formalism was used to generate
159: forecasts for the Dark Energy Task Force \citep{DETF}, and is
160: described in an appendix to that report.  
161: 
162: \section{The Weak Lensing 2-Point Likelihood}
163: \subsection{Observables}
164: \label{observsec}
165: We make the assumption that {\em the
166:   Universe has only weak scalar
167:   perturbations to a homogeneous and isotropic 4-dimensional metric.}
168: In this case 
169: the metric can be 
170: written in the Newtonian gauge as a perturbed Robertson-Walker metric:
171: \begin{equation}
172: ds^2 = (1+2\Psi)dt^2 - a^2(t)(1+2\Phi)\left[d\chi^2 +
173:   \chi_0^2S_k^2(\chi/\chi_0)(d\theta^2+\sin^2\theta d\varphi^2)\right]
174: \end{equation}
175: 
176: We assign all mass and sources in the Universe to 
177: a series of narrow spherical shells centered at redshifts
178: $a_i=(1+z_i)^{-1}$ for $i\in \{1,\ldots,N_z\}$.    
179: There is a comoving angular
180: diameter distance $D_i$ to each shell, and the comoving radial extent of
181: each shell is $\Delta\chi_i$.  Note that the Robertson-Walker metric
182: formula for angular-diameter distance is $D=\chi_0 S_k(\chi/\chi_0)$,
183: where $\chi_0$ is the comoving radius of curvature of the Universe.   
184: For small values of the curvature $\omega_k\equiv -k/\chi^2_0$ 
185: we have 
186: \begin{equation}
187: \Delta\chi_i \approx \Delta D (1-\omega_k D_i^2/2)
188: = { D_{i+1} - D_{i-1} \over 2} (1-\omega_k D_i^2/2).
189: \end{equation}
190: The Robertson-Walker metric also requires $\Delta\chi_i = \Delta z_i /
191: h(z_i) = \Delta a_i / a_i^2 h(a_i)$.  In this paper the Hubble
192: parameter will be written as $H(z)=h(z)H_{100}$, $H_{100}=100\,{\rm
193:   km}\,{\rm s}^{-1}\,{\rm Mpc}^{-1}$, and all distances will be in
194: units of $c/H_{100}=2998$~Mpc. 
195: 
196: We assume that the photon sources in a survey will be
197: divided into a series of {\em sets} $\alpha\in\{1,2,\ldots, N_s\}$.
198: Note the use of latin
199: indices for redshift shells, greek for source sets.
200: We follow \citet{HJ04} by assigning
201: each source set up to two observables: first its sky-plane density
202: fluctuations $g_\alpha(\theta,\varphi)$, and second a lensing convergence
203: $\kappa_\alpha(\theta,\varphi)$.  The convergence $\kappa$ might be
204: inferred from the shear or flexion \citep{BaconFlex} of galaxies, by a
205: quadratic estimator on the 
206: CMB or 21-cm radiation fluctuations, or by any other observable except the
207: source density.  The sources can be assigned to sets by photometric or
208: spectroscopic redshift, or even cruder color criteria
209: \citep{ConnollyJain}, but there could be other criteria such as galaxy
210: type, or perhaps observation by different instruments.  We demand 
211: only that the criteria for division of the sources be spatially
212: homogeneous, and that the division be invariant under application of
213: gravitational lensing distortion.  For notational convenience we
214: assign each set a nominal redshift $z_\alpha$, but a set can span a
215: broad redshift range.  If the sources are discrete objects such as
216: galaxies, then 
217: the mean density on the sky of members of each set are denoted
218: $n_\alpha$.
219: 
220: A source in set $\alpha$ has a probability $p_{\alpha i}$ of lying on
221: redshift shell $i$.   The collection of galaxies in set $\alpha$ on
222: shell $i$ will be called the {\em subset} $\alpha i$.  The survey is
223: assumed to tell us only which {\em set} any individual galaxy belongs to,
224: but not which {\em subset.}  The $p_{\alpha i}$ are parameters which
225: must be constrained by the lensing survey data or by additional  
226: observations, {\it e.g.} a spectroscopic redshift survey.  
227: 
228: When the lensing sources are drawn from a spectroscopic survey (or
229: when the source is the CMB), then the redshift
230: probability is known {\it a priori}, and in particular the sets are
231: probably divided by redshift so that $p_{\alpha i}$ is essentially the
232: identity matrix.  The formalism can obviously accommodate the
233: simultaneous analysis of WL samples with varying modes of redshift
234: assignment. 
235: 
236: Both the source density fluctuation $g_\alpha$ and convergence $\kappa_\alpha$
237: have a component due to
238: intrinsic fluctuations plus a component due to gravitational lensing.
239: Both are also measured as weighted sums over their respective subsets.
240: We have 
241: \begin{eqnarray}
242: \label{density1}
243: 1+g_\alpha(\theta,\varphi) & = & \sum_i p_{\alpha i}\left[1+g^{\rm
244:     int}_{\alpha i}(\theta,\varphi)\right]
245:  \left[1 + q_{\alpha i} \kappa^{\rm lens}_i(\theta,\varphi)\right] \\
246: \kappa_\alpha(\theta,\varphi) & = & \sum_i p_{\alpha i}\left[\kappa^{\rm int}_{\alpha i}(\theta,\varphi) + 
247: (1+f_{\alpha i}) \kappa^{\rm lens}_i(\theta,\varphi)\right].
248: \end{eqnarray}
249: Here we have assigned each subset a {\em magnification bias factor}
250: $q_{\alpha i}$ and a {\em shear calibration factor} $f_{\alpha i}$.
251: In a simple flux-limited selection, the magnification bias factor will be
252: determined by the logarithmic slope of the counts vs flux, and is typically of
253: order unity.  The shear calibration factor allows for the possibility
254: that the inferred lensing convergence is mis-measured by some factor
255: $1+f_{\alpha i}$ due to multiplicative errors in the lensing
256: methodology, \eg as investigated by \citet{STEP1}.
257: 
258: In the limit $g^{\rm int}\ll 1$ and $\kappa^{\rm lens}\ll 1$,
259: we can drop the second-order term in Equation~\ref{density1} and write
260: \begin{eqnarray}
261: \label{density2}
262: g_\alpha & = & \sum_i p_{\alpha i}\left[g^{\rm int}_{\alpha i} +
263: q_{\alpha i} \kappa^{\rm lens}_i\right] \\
264: \label{kappa2}
265: \kappa_\alpha & = & \sum_i p_{\alpha i}\left[\kappa^{\rm int}_{\alpha i} + 
266: (1+f_{\alpha i}) \kappa^{\rm lens}_i\right].
267: \end{eqnarray}
268: In this case the equations are linear in all the angular functions $g$
269: and $\kappa$, so we can decompose them into spherical harmonic
270: coefficient $g_{\alpha \ell m}$, $\kappa^{\rm lens}_{i\ell m}$,
271: etc, and Equations~(\ref{density2}) and (\ref{kappa2}) hold
272: independently for every harmonic $\ell m$.  We will henceforth assume
273: that the spherical-harmonic decomposition has been executed for all
274: the angular functions $g$, $\kappa$, and suppress the $\ell m$
275: indices for brevity.
276: 
277: We note that while $\kappa^{\rm lens}\ll 1$ is a good approximation
278: over most of the sky, $g^{\rm int}\ll1$ is a poor approximation for
279: thin density slices on smaller angular scales.  We will forge ahead
280: nonetheless with the assumption that {\em lensing magnification simply
281:   adds to the intrinsic density fluctuations}, recognizing that a real
282: analysis of data 
283: with magnification bias may require inclusion of the nonlinear
284: coupling between spherical harmonics that is induced by magnification
285: bias on highly structured density fields. 
286: 
287: The lensing convergence is determined entirely by the metric if we
288: make the assumption that light rays are following its null geodesics.
289: The paths of null geodesics are determined by the {\em lensing
290:   potential}
291: \begin{equation}
292: \phi \equiv {1 \over 2}(\Psi-\Phi).
293: \end{equation}
294: For each of our redshift shells we define a projected lensing
295: potential via
296: \begin{equation}
297: \psi_i \equiv 2 \nabla^2_\theta \int_{\Delta\chi} \phi a\, d\chi,
298: \label{lenspsi}
299: \end{equation}
300: where the derivatives are taken with respect to angles on the sky.  We
301: will generally assume that $\psi$, like the observables, has been
302: decomposed into spherical harmonics, and we will take the flat-sky
303: approximation.
304: 
305: With the definition (\ref{lenspsi}), and the adoption of the
306: weak-lensing limit and Born approximation, the lensing convergence is
307: \begin{eqnarray}
308: \label{lenslaw}
309: \kappa^{\rm lens}_i 
310: & = & \sum_j A_{ij} {\psi_j
311:   \over 2a_jD_j}, \\
312: \label{lenslaw2}
313: A_{ij} & \equiv & 
314: \left\{
315: \begin{array}{cl}
316: {D_{ij} \over D_i} \approx (1-D_j / D_i) (1-\omega_k
317: D_i D_j / 2) & z_i>z_j, \\
318: 0 & z_i\le z_j.
319: \end{array}
320: \right.
321: \end{eqnarray}
322: $D_{ij}$ is the comoving angular diameter distance to $z_i$ as viewed
323: from $z_j$.
324: In summary, the observables from the survey are, for each spherical harmonic:
325: \begin{eqnarray}
326: \label{observables}
327: g_\alpha & = & \sum_ip_{\alpha i}
328: \left[ q_{\alpha i} \sum_j A_{ij} {\psi_j \over 2a_j D_j}
329:  + g^{\rm int}_{\alpha i } \right] \\
330: \nonumber
331: \kappa_\alpha & = & \sum_ip_{\alpha i}
332: \left[ (1+f_{\alpha i}) \sum_j A_{ij} {\psi_j \over 2 a_j D_j}
333:  + \kappa^{\rm int}_{\alpha i } \right].
334: \end{eqnarray}
335: We reiterate that these equations depend only upon the assumption of
336: a Robertson-Walker metric with scalar perturbations,
337: plus the approximation that magnification bias and
338: intrinsic density fluctuations are additive.
339: 
340: The equations for the two observables are symmetric under the
341: interchange of $g\leftrightarrow \kappa$ and
342: $q\leftrightarrow(1+f)$.  Since $q\sim 1+f$, the lensing effects are
343: similar.  However the intrinsic density fluctuations $g^{\rm int}$ are
344: $\approx300\times$ stronger than $\kappa^{\rm int}$, breaking the
345: symmetry.  Density-field observations are dominated by the intrinsic
346: signal while convergence (shear) observations are dominated by lensing
347: effects.
348: 
349: \subsection{Degeneracies}
350: Equations (\ref{lenslaw})--(\ref{observables}) reveal a family of
351: degeneracies present in lensing observations, as described in
352: \citet{curvature}.  The transformations
353: \begin{eqnarray}
354: D_j & \rightarrow & D_j(1 + \alpha_0 + \alpha_1D_j + \alpha_2D_j^2) \nonumber
355: \\
356: \psi_i & \rightarrow & \psi_i(1 + \alpha_0 + 2\alpha_1D_j +
357: 2\alpha_2D_j^2) 
358: \label{degen} \\
359: \omega_k & \rightarrow & \omega_k + 2\alpha_2 \nonumber
360: \end{eqnarray}
361: leave the observables unchanged, to first order in $\{\alpha_0,
362: \alpha_1D, \alpha_2D^2, \omega_kD^2\}$. It will hence be impossible
363: for lensing$+$density surveys to constrain $\omega_k$ or any quadratic
364: (in $D$) deviations to $\ln D$, unless there are prior constraints on
365: these variables or on $\psi, g^{\rm int},$ or $\kappa^{\rm int}$.
366: Constraint on these three degeneracies is unlikely to arise from
367: models of intrinsic clustering or alignment, since it is unlikely that
368: {\it a priori} models of the redshift dependence of galaxy bias could
369: reach high precision.  We hence expect that these degeneracies are
370: going to be broken by theoretical models of the potential fluctuation
371: power spectrum, or by other distance indicators such as supernovae or BAO.
372: 
373: \subsection{Limber Approximation}
374: To forecast the constraints on the parameters of this model, we
375: require a likelihood expression for the observables.  The two
376: fundamental assumptions we make are:
377: \begin{enumerate}
378: \item The distributions of the lensing potential, intrinsic galaxy density
379:   fluctuations, and 
380:   intrinsic shape correlations $\psi_i, g^{\rm int}_{\alpha i},$ and
381:   $\kappa^{\rm int}_{\alpha i}$ are described by a multivariate
382:   Gaussian with zero mean.
383: \item The Limber approximation is valid and there is no correlation
384:   between these variables on distinct redshift shells or between
385:   different spherical harmonics:
386: \begin{eqnarray}
387: \nonumber
388: \langle X_{i\ell m} Y_{j\ell^\prime m^\prime} \rangle & = & \delta_{ij}
389: \delta_{\ell \ell^\prime}
390: \delta_{m m^\prime}
391: \left(D_i^2\Delta\chi_i\right)^{-1} P_i^{XY}(\ell/D_i) \\
392: \label{limber}
393: \langle X_{i\ell m} \psi_{j\ell^\prime m^\prime} \rangle & = & -2\delta_{ij}
394: \delta_{\ell \ell^\prime}
395: \delta_{m m^\prime}
396: a_i\left(\ell/D_i\right)^2 P_i^{X\phi}(\ell/D_i) \\
397: \nonumber
398: \langle \psi_{i\ell m} \psi_{j\ell^\prime m^\prime} \rangle & = & \delta_{ij}
399: \delta_{\ell \ell^\prime}
400: \delta_{m m^\prime}
401: 4a_i^2D_i^2\Delta\chi_i\left(\ell/D_i\right)^4 P_i^{\phi\phi}(\ell/D_i) 
402: \end{eqnarray}
403: where $X,Y \in \{g^{\rm int},\kappa^{\rm int}\}$, and $P_i^{XY}(k)$ is the 3-d
404: cross-spectrum of variables $X$ and $Y$ at epoch $a_i$.
405: \end{enumerate}
406: The first assumption insures that the likelihood of an observation is fully
407: specified by the expected covariance matrix of the observables.  The second
408: assumption implies that this covariance matrix can be expressed in
409: terms of the 3-d cross-power spectra of the
410: lensing potential, subset densities, and subset intrinsic correlations
411: $\{\phi_i,g^{\rm int}_{\alpha i},\kappa^{\rm int}_{\alpha i}\}$ at
412: each redshift shell. 
413: 
414: \subsection{Biases and Correlations}
415: A typical convention is to express the galaxy density power
416: $P^{gg}$ as a bias-scaled version of the mass spectrum $P^m$, plus a Poisson
417: shot-noise contribution, and then describing the mass-galaxy
418: covariance $P^{mg}$ with a correlation coefficient: 
419: \begin{eqnarray}
420: P^{gg} & = & (b^g)^2 P^m + {1 \over
421:   \rho} \\
422: P^{mg} & = & b^g r^g P^m.
423: \end{eqnarray}
424: The comoving volume number density $\rho$ of the sources determines
425: the shot noise for a Poisson process, but
426: there is no guarantee that the galaxies are distributed in the mass
427: distribution by a Poisson process.  Even when the galaxies do not have
428: Poissonian shot noise, we can still usually write the power in this
429: way for some bias parameter $b$; we just might keep in mind that
430: $r^g>1$ is formally allowed if the sources are not Poisson-distributed.
431: 
432: Most generally, both the
433: bias and correlation coefficients are different for each source {\em
434:   subset} as well as being functions of comoving wavenumber $k$.  Each
435: set $\alpha$ has a nominal redshift $z_\alpha$, and each subset has a
436: redshift deviation $\Delta z_{\alpha i}=z_i-z_\alpha$. Galaxies with bad
437: photo-z errors could easily have different bias from those with good
438: photo-z's; for example, highly-biased early types tend to have better
439: photo-z's.  So our analysis methods should allow for this
440: complication. 
441: 
442: We will adopt the bias/correlation notation for the intrinsic galaxy
443: density fluctuations {\em and} for the intrinsic convergence
444: $\kappa^{\rm int}$, except that we will parameterize the bias and
445: covariance with respect to the lensing potential rather than mass
446: distribution.  If $P^{gg}_{i\alpha\beta}$ is the 3d cross-power
447: between density fluctuations in subsets $\alpha i$ and $\beta i$ at
448: wavenumber $k$, and we
449: write $P^\phi_i$ for the lensing-potential 3d power spectrum, then we
450: express: 
451: \begin{eqnarray}
452: \nonumber
453: P^{gg}_{i\alpha\beta} & = &
454: b^g_{\alpha i}b^g_{\beta i}r^{gg}_{\alpha\beta i} 
455: \left({2 a_i \over 3 \omega_m}\right)^2 k^4P_i^\phi  +
456: {\delta_{\alpha\beta} \over \rho_{\alpha i}} \\
457: \label{pkk}
458: P^{\kappa\kappa}_{i\alpha\beta} & = &
459: b^\kappa_{\alpha i}b^\kappa_{\beta i}r^{\kappa\kappa}_{\alpha\beta i} 
460: \left({2 a_i \over 3 \omega_m}\right)^2 k^4P_i^\phi  +
461: {\delta_{\alpha\beta}\sigma^2_\gamma \over \rho_{\alpha i}} \\
462: \nonumber
463: P^{g\kappa}_{i\alpha\beta} & = &
464: b^g_{\alpha i}b^\kappa_{\beta i}r^{g\kappa}_{\alpha\beta i} 
465: \left({2 a_i \over 3 \omega_m}\right)^2 k^4P_i^\phi  
466: \end{eqnarray}
467: where $\rho_{\alpha i}$ is a comoving volume density of the galaxy subset
468: in the shell and $\sigma_\gamma$ is a measure of the shear noise per
469: galaxy.  These describe the normal ``shape noise'' term in the shear
470: power spectrum and the shot noise in the density field.  If flexions
471: or other observables are used to infer the convergence, then the
472: shape noise term may have a different form.
473: 
474: And if $P^{g\phi}_{i\alpha}$ is the cross-power between the density of
475: subset $i\alpha$ and lensing potential, we express
476: \begin{eqnarray}
477: \label{pgphi}
478: P^{g\phi}_{i\alpha} & = &
479: -b^g_{\alpha i}r^g_{\alpha i} 
480: {2 a_i \over 3 \omega_m} k^2P_i^\phi \\
481: \nonumber
482: P^{\kappa\phi}_{i\alpha} & = &
483: -b^\kappa_{\alpha i}r^\kappa_{\alpha i} 
484: {2 a_i \over 3 \omega_m} k^2P_i^\phi. 
485: \end{eqnarray}
486: Note that specifying the bias and correlation $b^\kappa$ and $r^\kappa$
487: of the intrinsic convergence with the lensing potential is equivalent
488: to giving the ``GI'' and
489: ``II'' intrinsic-alignment information, in the notation of \citet{Hirata2}.
490: 
491: The lensing power $P^\phi$ is a function of $z$ and $k$.  The biases
492: and correlation coefficients $b^\kappa_{\alpha i}, r^\kappa_{\alpha i},
493: b^g_{\alpha i},$ and $r^g_{\alpha i}$ are functions
494: of $k$, the nominal redshift $z_\alpha$ of the source set, and $\Delta
495: z_{\alpha i}$, the difference between the subset redshift and the
496: nominal set redshift.
497: 
498: Most complicated are the cross-correlation coefficients such as
499: $r^{gg}_{\alpha\beta i}$, which are, most generally, functions of $k$,
500: $z_i$, 
501: and {\em both} redshift errors $\Delta z_{\alpha i}$ and $\Delta
502: z_{\beta i}$. In order for the covariance matrix of all these fields
503: to be symmetric, we require
504: $r^{gg}_{\alpha\alpha i}=r^{\kappa\kappa}_{\alpha\alpha i}=1$ and
505: the symmetry $r^{XY}_{\alpha\beta i}=r^{YX}_{\beta\alpha i}.$  
506: Otherwise the correlation coefficients are free to vary, subject to
507: the constraint that the overall correlation matrix of the potential and
508: all fluctuations must remain non-negative.
509: 
510: This parameterization of the fluctuations of the potential and the
511: intrinsic fluctuations is completely general---we have not introduced
512: any further assumptions into the model as long as all the $b$'s and
513: $r$'s and $P^\phi$'s are free parameters (non-negative in the last
514: case).
515: 
516: \subsection{The Two-Point Statistics}
517: Combining the formula for observables (\ref{observables}), the Limber formulae
518: (\ref{limber}), and the bias notation (\ref{pkk}), (\ref{pgphi}),
519: the covariance matrix for the observables $\{g_\alpha,\kappa_\alpha\}$
520: at a given multipole can be broken into three submatrices: 
521: \begin{eqnarray}
522: \label{gg1}
523: C^{gg}_{\alpha\beta} \equiv \langle g_\alpha g_\beta \rangle 
524:  & = & \sum_{ij} 
525: p_{\alpha i}p_{\beta j} \left\{
526:  \rule[-3ex]{0ex}{6ex} %% This is to match bracket size
527: q_{\alpha i}q_{\beta j}\sum_n
528: A_{in}A_{jn}\Delta\chi_n k_n^4 P^\phi_n(k_n)
529: \right. \\
530: \nonumber
531:  & & \mbox{} + q_{\alpha i} A_{ij} { 2a_j  \over 3 \omega_m} D_j^{-1}
532: b^g_{\beta j}r^g_{\beta j}k_j^4 P^\phi(k_j) \\ \nonumber
533:  & &  \mbox{} + q_{\beta j} A_{ji} { 2a_i \over 3 \omega_m} D_i^{-1}
534: b^g_{\alpha i}r^g_{\alpha i}k_i^4 P^\phi(k_i) \\ \nonumber
535:  & &  \left. \mbox{} + \delta_{ij} \left({2 a_i \over 3 \omega_m}\right)^2
536: D_i^{-2}\Delta\chi_i^{-1}
537: b^g_{\alpha i}b^g_{\beta i}r^{gg}_{\alpha \beta i}k_i^4
538: P^\phi(k_i) 
539: \rule[-3ex]{0ex}{6ex} \right\} + {\delta_{\alpha\beta}\over n_\alpha}\\ 
540: \label{kg1}
541: C^{\kappa g}_{\alpha\beta} \equiv \langle \kappa_\alpha g_\beta \rangle 
542:  & = & \sum_{ij} 
543: p_{\alpha i}p_{\beta j} \left\{
544:  \rule[-3ex]{0ex}{6ex} %% This is to match bracket size
545: (1+f_{\alpha i})q_{\beta j}\sum_n
546: A_{in}A_{jn}\Delta\chi_n k_n^4 P^\phi_n(k_n)
547: \right. \\
548: \nonumber
549:  & & \mbox{} + (1+f_{\alpha i}) A_{ij} { 2a_j  \over 3 \omega_m} D_j^{-1}
550: b^g_{\beta j}r^g_{\beta j}k_j^4 P^\phi(k_j) \\ \nonumber
551:  & &  \mbox{} + q_{\beta j} A_{ji} { 2a_i \over 3 \omega_m} D_i^{-1}
552: b^\kappa_{\alpha i}r^\kappa_{\alpha i}k_i^4 P^\phi(k_i) \\ \nonumber
553:  & &  \left. \mbox{} + \delta_{ij} \left({2 a_i \over 3 \omega_m}\right)^2
554: D_i^{-2}\Delta\chi_i^{-1}
555: b^\kappa_{\alpha i}b^g_{\beta i}r^{\kappa g}_{\alpha \beta i}k_i^4
556: P^\phi(k_i) 
557: \rule[-3ex]{0ex}{6ex} \right\} + {\delta_{\alpha\beta}\over n_\alpha}\\ 
558: \label{kk1}
559: C^{\kappa\kappa}_{\alpha\beta} \equiv \langle \kappa_\alpha \kappa_\beta \rangle 
560:  & = & \sum_{ij} 
561: p_{\alpha i}p_{\beta j} \left\{
562:  \rule[-3ex]{0ex}{6ex} %% This is to match bracket size
563: (1+f_{\alpha i})(1+f_{\beta j})\sum_n
564: A_{in}A_{jn}\Delta\chi_n k_n^4 P^\phi_n(k_n)
565: \right. \\
566: \nonumber
567:  & & \mbox{} + (1+f_{\alpha i}) A_{ij} { 2a_j  \over 3 \omega_m} D_j^{-1}
568: b^\kappa_{\beta j}r^\kappa_{\beta j}k_j^4 P^\phi(k_j) \\ \nonumber
569:  & &  \mbox{} + (1+f_{\beta j}) A_{ji} { 2a_i \over 3 \omega_m} D_i^{-1}
570: b^\kappa_{\alpha i}r^\kappa_{\alpha i}k_i^4 P^\phi(k_i) \\ \nonumber
571:  & &  \left. \mbox{} + \delta_{ij} \left({2 a_i \over 3 \omega_m}\right)^2
572: D_i^{-2}\Delta\chi_i^{-1}
573: b^\kappa_{\alpha i}b^\kappa_{\beta i}r^{\kappa\kappa}_{\alpha \beta i}k_i^4
574: P^\phi(k_i) 
575: \rule[-3ex]{0ex}{6ex} \right\} +
576: \delta_{\alpha\beta}\left({\sigma^2_\gamma\over n}\right)_\alpha
577: \end{eqnarray}
578: The comoving wavevector is $k_i=\ell/D_i$.
579: In each equation, note that only one of the last three terms is
580: non-zero, depending on whether $j<i$, $j>i$, or $j=i$, respectively.
581: For the shear-shear correlation $C^{\kappa\kappa}$,
582: the $i=j$ term is recognizable as the ``II''
583: intrinsic-correlation effect of \citet{Hirata2}, while the $i<j$
584: and $j>i$ terms are their ``GI'' effect.  
585: 
586: Note that
587: the last term in the density-shear expression $C^{\kappa g}$
588: is an additional intrinsic-correlation term, between the galaxy
589: density and the intrinsic shapes, which is distinct from the
590: covariance between the lensing potential and shear.  This galaxy-shear
591: correlation has been constrained in the context of systematic errors
592: to ``galaxy-galaxy'' lensing, \eg
593: \citet{BernsteinNorberg,FLMvdBYJPM,Hetal04}. 
594: 
595: Examining the density-density correlation $C^{gg}$ we find the final
596: term has the normal expected form, but the first three terms describe
597: correlations induced by lensing magnification. 
598: 
599: Finally we note that the covariance matrix manifests the same
600: symmetries for $g\leftrightarrow\kappa$ that were discussed at the end
601: of \S\ref{observsec}.
602: 
603: \subsection{Likelihood and Fisher matrix}
604: Under our Gaussian assumption, the likelihood functions for the
605: observables are independent at each multipole $\ell m$.  We define a
606: data vector
607: ${\bf d}_{\ell m}$ to be the union of the $g_\alpha$ and
608: $\kappa_\alpha$ observables at each multipole, and ${\bf C}_{\ell}$ to
609: be the covariance matrix derived above.  Under our Gaussian
610: assumption, the total likelihood for the survey is
611: \begin{equation}
612: \label{likelihood}
613: -2\ln L = \sum_{\ell m} \left[ {\bf d}_{\ell m}^T {\bf C}_\ell^{-1} {\bf
614:   d}_{\ell m} + \ln | {\bf C}_\ell | \right].
615: \end{equation}
616: Forecasts of survey performance are made using the Fisher matrix.  The
617: usual formula for zero-mean Gaussian distributions applies
618: \citep{TTH97}.  We reduce the mode sum to a series of $N_\ell$ bins centered
619: on multipoles $\ell_i$, then the Fisher matrix element for parameters
620: $p$ and $q$ is
621: \begin{equation}
622: \label{fisher}
623: F_{pq} = \sum_{i=1}^{N_\ell} {(2\ell_i+1)\Delta \ell_i\,f_{\rm
624:     sky}\over 2}\,
625: {\rm Tr} \left[{\bf C}^{-1}_{\ell_i} {\partial {\bf C}_{\ell_i} \over \partial p}
626: {\bf C}^{-1}_{\ell_i} {\partial {\bf C}_{\ell_i} \over \partial q}\right].
627: \end{equation}
628: Examination of Equations~(\ref{gg1})--(\ref{kk1}) shows that 
629: all derivatives of ${\bf C}$ with respect to parameters are very simple.
630: The calculation of the Fisher matrix is reduced to rapid linear
631: algebra, significantly accelerated by exploiting the very sparse
632: nature of most of the derivative matrices.
633: 
634: We have thus succeeded in producing a likelihood function for the most
635: general joint lensing$+$density survey, for the case of Gaussian
636: likelihoods limited to 2-point statistics.  Given a likelihood we can
637: of course form a Fisher matrix for forecasting, or we can execute a
638: maximum-likelihood analysis of real data.  Since this likelihood
639: function makes no mention of a particular dark-energy theory, we see
640: that the parameterization chosen here permits a highly flexible
641: analysis.  Indeed no theory of gravity or initial conditions of the
642: Universe have been assumed either, just the existence of a Newtonian
643: gauge metric on a RW background cosmology.  The lensing-potential
644: power spectrum $P^\phi(k,z)$ appears as a series of free parameters,
645: as do the bias and correlation coefficients of the galaxy density and
646: intrinsic alignments.
647: 
648: We have variables
649: that describe, in the most general possible fashion, the important
650: systematic errors, excepting additive shear contamination:  
651: \begin{enumerate}
652: \item Uncertainty in power-spectrum theory will be expressed through
653:   prior distributions on the $P^\phi_i$ parameters. 
654: \item Shear calibration errors arises through finite prior uncertainty
655:   on the $f_{\alpha i}$. 
656: \item Magnification-bias calibration errors arises through finite prior
657:   uncertainty on the $q_{\alpha i}$. 
658: \item Intrinsic alignments are embodied through the $b^\kappa$,
659:   $r^\kappa$, $r^{g\kappa}$, and $r^{\kappa\kappa}$ coefficients. 
660: \item Redshift-distribution errors are manifested through the
661:   uncertainties in the $p_{\alpha i}$ probabilities. 
662: \end{enumerate}
663: 
664: The cost of this great generality is that there are a huge number of
665: nuisance parameters, enough to make us doubt whether the
666: maximum-likelihood analysis---or even the Fisher-matrix analysis!---is
667: feasible.  
668: 
669: \subsection{Parameter Inventory}
670: The WL survey covariance matrix has a
671: horrendously large number of parameters.  The cosmological treasure
672: lies in these:
673: \begin{itemize}
674: \item The 2 global cosmological parameters $\omega_m$ and $\omega_k$.
675: \item The distances $D_i$, which encode the expansion history of the
676:   Universe in $N_z$ steps.  The $\Delta\chi_i$ and the Hubble
677:   parameters  $h(z_i)$ can be expressed in terms of these and $\omega_k$.
678: \item The metric-potential power spectra $P_i^\phi$, which describe
679:   the growth of 
680:   dark-matter structure.  For $N_\ell$ bins in $\ell$, there will be
681:   $N_\ell N_z$ distinct matter-power parameters in the model.  A
682:   prediction for the growth of potential fluctuations will typically be an
683:   important element of any cosmological scenario under test, so the
684:   $P^\phi_i$ can be replaced as parameters by a much smaller number of
685:   cosmological parameters.
686: \end{itemize}
687: There are then a large number of nuisance parameters.  If there are
688: $N_{ss}$ non-empty source subsets, the nuisance parameters are 
689: \begin{itemize}
690: \item The redshift-distribution parameters $p_{\alpha i}$, with
691:   $N_{ss}-N_s$ degrees of freedom.
692: \item The shear-calibration errors $f_{\alpha i}$, another $N_{ss}$
693:   degrees of freedom.
694: \item The magnification-bias coefficients $q_{\alpha i}$, another $N_{ss}$
695:   degrees of freedom.
696: \item The source-density biases $b^g_{\alpha i}$ and correlation coefficients
697:   $r^g_{\alpha i}$ with respect to $\phi$, which may be scale-dependent,
698:   yielding $2N_\ell N_{ss}$ degrees of freedom.
699: \item The intrinsic-alignment power and correlations with the mass,
700:   $b^\kappa_{\alpha i}$ and $r^\kappa_{\alpha i}$, another $2N_\ell
701:   N_{ss}$ parameters. 
702: \item The correlation coefficients $r^{gg}_{\alpha\beta i}$, 
703: $r^{\kappa g}_{\alpha\beta i}$, and $r^{\kappa\kappa}_{\alpha\beta
704:   i}$, which also may be scale-dependent.  The number of such
705: parameters is $\approx 3N_\ell N_{ss}^2/2N_z$.
706: \end{itemize}
707: The number of nuisance parameters for a non-parametric analysis is
708: enormous.  If we are analyzing a photo-z survey with typical errors
709: $\Delta z\approx 0.05(1+z)$, then we would typically want to space the
710: redshift shells logarithmically in $1+z$, with $\Delta\ln a \approx
711: 0.02$ so that we resolve the redshift distribution of each photo-z
712: bin.  In this case, $N_z\approx100$ bins span $0<z<5$, and
713: we will require $N_{ss}\gtrsim 1000$ if we track all subsets out to
714: $\pm3\sigma$ of the photo-z distribution. 
715: 
716: To reduce the dimensionality of the likelihood function, we can
717: replace many of the discrete nuisance parameters by the values of
718: parameterized functions for the nuisance variables.
719: Table~\ref{nuisancefunc} lists the variables in the WL likelihood
720: function that can be replaced by parametric functions. The nuisance
721: variables are functions of: wavevector $k$; redshift $z$; and redshift
722: difference $\Delta z = z_\alpha - z_i$ between the nominal and true
723: redshifts of a source subset.
724: In later sections we will describe the parametric functions
725: that we have implemented to reduce the number
726: of degrees of freedom in the model.  Each time we introduce a
727: parametric function, we need to choose a fiducial parameter set and a
728: prior distribution for the parameters.
729: 
730: \begin{deluxetable}{lcc}
731: \tablewidth{0pt}
732: \tablecaption{Nuisance variables that can be replaced by functions}
733: \tablehead{
734: \colhead{Description} &
735: \colhead{Discrete variables} &
736: \colhead{Parametric function}
737: }
738: \startdata
739: Lensing potential power spectrum & $P^\phi_i$ & $P^\phi(k,z)$ \\
740: Shear calibration error & $f_{\alpha i}$ & $f(z, \Delta z)$ \\
741: Magnification bias & $q_{\alpha i}$ & $q(z, \Delta z)$ \\
742: Redshift distribution & $p_{\alpha i}$ & $p(z, \Delta z)$ \\
743: Source density bias & $ b^g_{\alpha i}$ & $b^g(k, z, \Delta z)$ \\
744: Density-mass correlation & $ r^g_{\alpha i}$ & $r^g(k, z, \Delta z)$ \\
745: Intrinsic alignment bias & $ b^\kappa_{\alpha i}$ & $b^\kappa(k, z, \Delta z)$ \\
746: IA-density correlation & $ r^\kappa_{\alpha i}$ & $r^\kappa(k, z, \Delta z)$ \\
747: Density-density x-correlation & $r^{gg}_{\alpha\beta i}$ &
748: $r^{gg}(k,z,\Delta z_\alpha, \Delta z_\beta)$ \\
749: Density-IA x-correlation & $r^{g\kappa}_{\alpha\beta i}$ &
750: $r^{g\kappa}(k,z,\Delta z_\alpha, \Delta z_\beta)$ \\
751: IA-IA x-correlation & $r^{\kappa\kappa}_{\alpha\beta i}$ &
752: $r^{\kappa\kappa}(k,z,\Delta z_\alpha, \Delta z_\beta)$
753: \enddata
754: \label{nuisancefunc}
755: \end{deluxetable}
756: 
757: \section{Spectroscopic Redshift Likelihood}
758: \label{speclike}
759: If we draw a single
760: member from source set $\alpha$ and measure its spectroscopic redshift in
761: an unbiased 
762: fashion, then by definition the likelihood of the spectroscopic
763: redshift being on shell $i$ is $p_{\alpha i}.$  If we measure
764: $N_\alpha^{\rm spec}$ redshifts, and find that $N^{\rm spec}_{\alpha
765:   i}$ are on shell $i$, 
766: then the likelihood is
767: \begin{equation}
768: \ln L = \sum_i N^{\rm spec}_{\alpha i} \ln p_{\alpha i}.
769: \end{equation}
770: This is true if the redshifts are statistically independent, which
771: requires that they be dispersed across the sky to eliminate source
772: correlations.  We assume this limit.
773: 
774: Following \citet{MaBernstein}, the
775: Fisher matrix for the parameters $\{p_{\alpha i}\}$ resulting from the
776: unbiased spectroscopic observations is
777: \begin{equation}
778: F_{\alpha i \beta j} = \left\langle {\partial^2 (-\ln L) \over \partial p_{\alpha i}
779: \partial p_{\alpha j} } \right\rangle = N^{\rm spec}_\alpha {\delta_{ij} \over
780: p_{\alpha i}}.
781: \label{specfish}
782: \end{equation}
783: We add this Fisher information to the density-lensing Fisher matrix
784: (\ref{fisher}) 
785: when considering the constraints offered by a WL survey that is
786: combined with an {\em unbiased} spectroscopic redshift survey drawn
787: from one or more of the source population sets.
788: 
789: We do not in general presume any functional form for the $p_{\alpha
790:   i}$ redshift distributions when the sets are assigned from
791: photo-z's.  We adopt fiducial values either from an analytic form or from
792: a simulation of photo-z performance.  Then we leave all the $p_{\alpha
793:   i}$ as free parameters in the Fisher matrix, adding the
794: spectroscopic-survey Fisher matrix (\ref{specfish}) if appropriate to
795: the planned 
796: experiment.  Note that the cross-correlations in the WL survey data
797: offer constraints on the redshift distribution even if there is no
798: unbiased spectroscopic survey ($N^{\rm spec}_\alpha=0$).
799: 
800: \section{An Implementation}
801: A package of {\tt C++} classes implements the Fisher matrix
802: calculation for Gaussian lensing$+$density observations, the
803: spectroscopic-survey Fisher matrix, plus the
804: models for lensing-potential power and nuisance functions
805: described below.  From these classes we can construct numerous
806: applications, the most obvious being a Fisher-matrix forecast of
807: cosmological constraints from the combination of a photometric galaxy
808: lensing/density survey, plus a redshift survey to constrain the
809: photo-z distribution.
810: We list in Table~\ref{params1} input fields for this forecasting
811: implementation.  Further program inputs
812: are listed in later sections which detail the models for the lensing
813: power spectrum and nuisance parameters that we describe below and
814: adopt for this implementation.
815: 
816: In the current implementation we assume the source galaxies to be
817: binned solely by photo-z, but generalizations are possible, \eg including a
818: second population of source galaxies
819: sets that are observed spectroscopically.  Note that when additional
820: galaxy populations are introduced, we need a new set of nuisance functions
821: to describe them.  Furthermore, we need to model the
822: cross-correlations between all galaxy populations.
823: 
824: The calculation of the Fisher matrix takes $<1$ minute per
825: multipole bin using a single core of a typical current-epoch desktop
826: CPU.  Total execution time for a forecast is 10--20 minutes with the
827: default parameters, with the most time-consuming operation being the
828: marginalization over bias model parameters.  The execution time is
829: very sensitive to the redshift-shell width $\Delta\ln a$ and to the
830: width of the photo-z error distribution, as these control the number
831: of subsets and the parameter count.
832: 
833: \begin{deluxetable}{cll}
834: \tablewidth{0pt}
835: \tablecaption{Controlling inputs to Fisher forecast}
836: \tablehead{
837: \colhead{Parameter name} &
838: \colhead{Description} &
839: \colhead{Default}
840: }
841: \startdata
842: {\tt fsky} & $f_{\rm sky}$, imaging sky coverage & 0.5 \\
843: {\tt minLogL} & $\log_{10} \ell_{\rm min}$, minimum multipole & 1.0 \\
844: {\tt maxLogL} & $\log_{10} \ell_{\rm max}$, maximum multipole & 3.5 \\
845: {\tt logLStep} & $\Delta \log_{10} \ell$, multipole bin width in dex
846: & 0.3 \\
847: {\tt zmax} & $z_{\rm max}$, redshift of most distant shell & 3.5 \\
848: {\tt dlna} & $\Delta \ln a$, width of distance shells & 0.03 \\
849: {\tt coreDLna} & Maximum $|\ln (1+z_i)/(1+z_\alpha)|$ of subsets &
850: 0.15 \\
851: {\tt sigGamma} & $\sigma_\gamma$, shape noise per source galaxy & 0.24
852: \\
853: {\tt zdist} & String specifying fiducial source $n_\alpha$ and
854: $p_{\alpha i}$ & \nodata \\
855: {\tt logNSpec} & $\log_{10} N_{\rm spec}$ & 5. \\
856: {\tt outfile} & Root name for output files & \nodata 
857: \enddata
858: \label{params1}
859: \end{deluxetable}
860: 
861: 
862: 
863: \section{Power Spectrum Models}
864: \label{powermodel}
865: In most cosmological models, theory will offer strong guidance to the
866: form of the lensing-potential power spectrum $P^\phi(k,z)$.
867: In most forecasting or data-reduction codes this is a
868: fully deterministic function of a small number of cosmological
869: parameters.  In our analysis, however, the theoretical prediction is
870: taken as the mean $P^\phi$ value of a prior distribution of finite
871: uncertainty. 
872: 
873: \subsection{Central Model}
874: The WL likelihood given above can be calculated for any model that
875: predicts $P^\phi$.  We have chosen to implement a model that allows
876: for failure of General Relativity in describing growth of structure;
877: but other models are possible if one wishes to test the Poisson
878: equation or other tenets of General Relativity.  Under the conditions
879: \begin{enumerate}
880: \item The potential and the mass-energy density are related by the
881:   Poisson equation of General Relativity, and
882: \item Non-relativistic matter is the only significant inhomogeneous
883:   component of the 
884:   Universe, \ie there is no dark-energy clustering, and
885: \item Matter is conserved, $\bar \rho_m \propto a^{-3}$, and
886: \item $\Phi = -\Psi$, as in the absence of anisotropic stress for
887:   General Relativity
888: \end{enumerate}
889: then the potential power spectrum is related to the matter-density
890: fluctuation spectrum $P^m$ via
891: \begin{equation}
892: k^4P^\phi(k,a) = \left({3\omega_m \over 2a}\right)^2 P^m(k,a).
893: \label{poisson}
894: \end{equation}
895: Under these conditions, the linearized perturbations to the metric
896: grow in a scale-free manner so we can write
897: \begin{equation} 
898: P^\phi_{\rm lin}(k, a) = g_\phi^2(a) P^\phi_{\rm prim}(k) T^2(k).
899: \end{equation}
900: In our current code, the primordial power spectrum is a power law 
901: \begin{equation}
902: \Delta^2_{\rm prim}(k) \equiv {k^3 \over 2 \pi^2}P^\phi_{\rm prim}(k)
903:  = \left({3 \Delta_\zeta \over 5}\right)^2 (k/k_0)^{n_s-1}.
904: \end{equation}
905: The curvature variation $\Delta_\zeta$ and spectral index $n_s$ are
906: free parameters.
907: A running of the slope could easily be
908: added. The normalization wavenumber $k_0$ must be set by some
909: convention.  We typically adopt the 5-year WMAP parameters as fiducial
910: values \citet{WMAP5}.
911: 
912: The transfer function $T(k)$ is taken from \citet{EH}.  It is a function of
913: the matter and baryon densities $\omega_m$ and $\omega_b$.  The impact
914: of massive neutrinos could be added to the transfer function if
915: desired.  We ignore the baryon acoustic oscillations; experiments that
916: try to exploit them will generate a distinct Fisher matrix for them.
917: 
918: The map from $P^\phi_{\rm lin}$ to the nonlinear $P^\phi_{\rm nl}$ is derived
919: using the prescription of \citet{Smith} for nonlinear $P^m$, combined
920: with the Poisson equation (\ref{poisson}).  The
921: Smith {\it et al.} formula also requires knowledge of $\Omega_m$ at
922: the desired epoch, but it can be expressed in terms of other
923: quantities that are already in our model: $\Omega_m(z_i)=\omega_m
924: a_i^{-3} h^{-2}(z_i).$  We do not expect the \citet{Smith} formula to
925: describe non-linear growth to high accuracy for all (or any) cosmologies.
926: It does however capture the dependence of non-linear
927: power on cosmological parameters to a level that suffices for
928: forecasting purposes.
929: 
930: In General Relativity, the growth function $g_\phi(a)$ is determined by the
931: expansion history $H(z)$.  Defining $F=\ln (ag_\phi)$, the growth
932: equation is
933: \begin{equation}
934: \label{growtheq}
935: F^{\prime\prime} + \left( F^\prime\right)^2 + F^\prime\left(2+ {d\ln h
936:     \over d \ln a}\right) = {3 \omega_m \over 2 h^2 a^3},
937: \end{equation}
938: where a prime denotes differentiation with respect to $\ln a$.  Note
939: that $F^\prime$ is the quantity $d\ln g_m/d\ln a$ that appears in the
940: peculiar-velocity power spectrum for a tracer of mass.  
941: 
942: If GR holds, then the above relations fully specify the model for
943: $P^\phi$ given $\{\omega_m, \omega_b, n_s, \ln \Delta_\zeta\}$ plus
944: the expansion history, which in turn is given by $\{D_i,
945: \omega_k\}$. As a test of GR, we allow the growth function arbitrary
946: deviations from the $F_{\rm fid}$ that solves the GR growth equation for
947: the fiducial expansion history:
948: \begin{equation}
949: \ln ag_\phi(a_i) = F_{\rm fid}(a_i) + \delta F_i.
950: \end{equation}
951: The $\delta F_i$ become parameters of the likelihood function.
952: 
953: \subsection{Model Errors}
954: An important WL systematic is the expected finite accuracy in
955:   theoretical modeling of the power spectrum.  We hence introduce
956:   an error function to describe the describe the (logarithmic)
957:   difference between the power $P^\phi$ and the value predicted by the
958:   parametric model described in the previous paragraphs:
959: \begin{equation}
960: \ln P^\phi(k,a) = \ln P^\phi_{\rm nl}(k,a) + \delta \ln P(k,a).
961: \end{equation}
962: The nuisance function $\delta\ln P$ will be described with the
963: ``$kz$'' parametric form described in Appendix~\ref{functions}.
964: We parameterize the $\delta \ln P$ function by its values $\delta
965: P_{ij}$ at a grid of points $(k_i, a_j)$ regularly spaced in $\ln k$
966: and $\ln a$.  The $\delta\ln P$ is linearly interpolated between grid
967: points. 
968: The $\delta P_{ij}$ become free parameters of
969: the model and hence
970: parameters in the likelihood function. We then place an independent
971: Gaussian prior on each $\delta P_{ij}$ which has mean of zero and a
972: standard deviation of
973: \begin{eqnarray}
974: \sqrt{{\rm Var}(\delta P_{ij})} & = & 0.012 f_{\rm Zhan} \cases {
975: 1 + 5\log_{10}(k_i/k_1) & $k_i>k_1$ \cr
976: (k_i/k_1)^{1+a_j} & $k_i<k_1$ }  \\
977: k_1 & \equiv & a^{-2.6} \,{\rm Mpc}^{-1}.
978: \end{eqnarray}
979: This function is a fit to an estimate, supplied by Hu Zhan, of the
980: impact of baryonic physics on the mass power spectrum \citet{ZhanKnox,Jing}.
981: We scale the overall size of the theory-error systematic with the
982: control scalar $f_{\rm Zhan}$.  We can also adjust the density $\Delta
983: \ln k$ and $\Delta \ln a$ at which the $\delta P_{ij}$ grid points are
984: spaced. This corresponds to setting some coherence length for theory
985: errors in this space.  In \S\ref{tuneup} we will investigate the choice of
986: these grid spacings.
987: 
988: The procedure above means that we replace the $P^\phi_i$ as parameters in
989: our likelihood with a new (and hopefully smaller) set:
990: \begin{itemize}
991: \item The small set $\{\omega_m, \omega_b, \Delta^2_\zeta, n_s\}$ that
992:   control the linear power spectrum.
993: \item The $\{\delta F_i\}$ which define the growth function vs redshift.
994: \item A grid of theory error values $\delta P$, which are nuisance
995:   parameters to marginalize after construction of a Fisher matrix or
996:   likelihood.  We have physically-based priors to apply to these
997:   before marginalization. 
998: \end{itemize}
999: Table~\ref{params2} lists the input fields for the part of our
1000: forecasting code which constructs the power-spectrum model.  The WMAP5
1001: $\Lambda$CDM cosmology provides the fiducial values of all
1002: lensing power values; the program inputs define the behavior of the
1003: deviations from the theoretical model and the prior expectations on
1004: the size of such deviations.
1005: \begin{deluxetable}{cll}
1006: \tablewidth{0pt}
1007: \tablecaption{Power spectrum inputs to Fisher forecast}
1008: \tablehead{
1009: \colhead{Parameter name} &
1010: \colhead{Description} &
1011: \colhead{Default}
1012: }
1013: \startdata
1014: {\tt zhan} & $f_{\rm Zhan}$, power-spectrum theory uncertainty
1015: relative to baryonic effects & 0.5 \\
1016: {\tt psDlnk} & $\Delta \ln k$, node spacing of power-spectrum theory
1017: errors in $k$ & 1.0 \\
1018: {\tt psDlna} & $\Delta \ln a$, node spacing of power-spectrum theory
1019: errors in $a$ & 0.5  
1020: \enddata
1021: \label{params2}
1022: \end{deluxetable}
1023: 
1024: 
1025: \section{Non-parametric nuisance modeling}
1026: \label{nuisance}
1027: \subsection{General comments on nuisance functions}
1028: The WL likelihood contains many nuisance parameters that are discretized
1029: representations of nuisance {\em functions}.  It is common in the
1030: literature to assign some parametric form to a nuisance
1031: function, then marginalize over the parameters of the nuisance
1032: function to recover a purely cosmological likelihood.  This can be a
1033: very dangerous approach: if the nuisance function does {\em not}
1034: in actuality follow the assumed form, then the process is invalid and
1035: we may have greatly overestimated the power of the experiment to
1036: remove the systematic error from the signal.  When marginalizing over
1037: a systematic, we must be sure that the assumed parametric form is
1038: sufficiently flexible to include any expected manifestation of the
1039: systematic.  For example we should not assume that systematics scale
1040: linearly with redshift unless there is a physical reason to expect
1041: this.
1042: 
1043: It is unfortunately not possible to model a completely free function
1044: with a finite number of free parameters.  This becomes possible,
1045: however, if we limit the bandwidth of variation in the function.
1046: As an example consider our power-spectrum theory error function
1047: $\delta \ln P(k,z)$.  We could decompose $\delta \ln P$ into
1048: Fourier modes or polynomial terms over its finite $(k,z)$ domain.
1049: Retaining a finite number of modes or terms leads to a tractable
1050: parameterization, albeit with a maximum frequency or polynomial order
1051: that defines a coherence length for the reconstructed function. For
1052: $\delta \ln P$ we choose to limit the bandwidth using linear interpolation
1053: between a 2d grid of specified values.
1054: In
1055: Appendix~\ref{functions} we describe the family of functions that we
1056: use to model the nuisance functions of $(k, z, \Delta z)$ that are
1057: common in the WL likelihood analysis ({\it cf.}
1058: Table~\ref{nuisancefunc}).\footnote{
1059: In practice we use $\ln a$ and $\Delta\ln a$ to specify each subset's
1060: nominal redshift and redshift error, but in the text we will stick
1061: with $(z,\Delta z)$ to reduce the clutter.}
1062:   The Appendix describes both the
1063: functional form, and the prior likelihoods on the parameters that are
1064: used to give the nuisance function the desired $RMS$ uncertainty.
1065: 
1066: These models of nuisance functions are non-parametric in the sense of
1067: being able to reproduce very general types of behavior once the
1068: bandwidth is specified.  The question remains: what is the proper
1069: choice of bandwidth to allow the nuisance function?  Our approach is
1070: to find the bandwidth which {\em causes the most damage to
1071:   cosmological constraints} under a prior that specifies the expected RMS
1072: fluctuations in the nuisance function.  This is the most conservative
1073: approach.  Typically one finds the following: if the nuisance function
1074: is given a highly coherent, low-order functional form, then it is
1075: easily distinguished from cosmological signals and can be marginalized
1076: away with little damage to cosmological constraints.  On the other
1077: hand if the nuisance-function bandwidth is very high, then the
1078: broad WL kernel tends to average away the nuisance signal, leaving 
1079: little trace in the cosmology.  There is an intermediate point where
1080: the systematic error is most easily confused with cosmology.  The
1081: conservative approach is to find this regime and use it for modeling
1082: the systematic error.  In \S\ref{tuneup} we will find coherence lengths in $z$
1083: and $k$
1084: at which our systematics are most damaging.
1085: 
1086: \subsection{Redshift Distributions}
1087: We specify the fiducial values of the photo-z distribution $n_\alpha$
1088: and error probabilities $p_{\alpha i}$ either with analytic formulae
1089: (\eg photo-z errors Gaussian in $\ln a$), or by taking the output of
1090: a simulation of galaxy detection and photo-z assignment for the chosen
1091: survey.  For spectroscopic samples or the CMB source plane there are
1092: no photo-z errors at all.
1093: 
1094: We do not place any parametric form or prior assumption on the
1095: $p_{\alpha i}$.  All redshift constraints arise either
1096: from the lensing survey data itself or from additional spectroscopic
1097: data.  The likelihood arising from spectroscopic redshift samples is
1098: described in \S\ref{speclike}.
1099: 
1100: \subsection{Shear calibration and magnification bias}
1101: The shear calibration factors $f_{\alpha i}$ and magnification bias
1102: coefficients $q_{\alpha i}$ are, most generally, distinct in every
1103: subset.  We use the ``$z\Delta z$'' functional
1104: form described in Appendix~\ref{functions} to generate the $f_{\alpha
1105:   i}$ and $q_{\alpha i}$ from a smaller set of function parameters.
1106: Each function is specified by a polynomial function of $\Delta
1107: z$; polynomial coefficients are interpolated between grid points
1108: equally spaced in $\ln a$ at intervals $\Delta\ln a$.
1109: 
1110: We set the fiducial functions to be
1111: $f_{\alpha i}=0,$ and $q_{\alpha i}=q_{\rm fid}$ independent of
1112: $(k,z)$.  The priors on the
1113: polynomial coefficients are chosen to yield a chosen RMS variation of
1114: $f$ or of $q$.  As detailed in Appendix~\ref{functions}, we also
1115: specify whether the nuisance function varies mostly along the $z$
1116: direction or along the $\Delta z$ direction of its domain.
1117: 
1118: Table~\ref{params3} lists the program inputs necessary to specify the
1119: model for $f$ or $q$: their functional form, fiducial values, and
1120: priors on deviations from the fiducial.
1121: 
1122: \begin{deluxetable}{cll}
1123: \tablewidth{0pt}
1124: \tablecaption{Shear calibration and magnification bias inputs to Fisher forecast}
1125: \tablehead{
1126: \colhead{Parameter name} &
1127: \colhead{Description} &
1128: \colhead{Default}
1129: }
1130: \startdata
1131: {\tt fRMS} & $f_{\rm RMS}$, RMS variation of $f$ allowed under prior &
1132: 0.01 \\
1133: {\tt qFid} & $q_{\rm fid}$, fiducial value for all $q_{\alpha i}$ & 1.0 \\
1134: {\tt qRMS} & $q_{\rm RMS}$, RMS variation of $q$ allowed under prior &
1135: 0.1 \\
1136: {\tt fqDlna} & $\Delta\ln a$, node spacing for $f$ \& $q$ models & 0.5
1137: \\
1138: {\tt fqDzOrder} & Order of polynomial used to model $\Delta z$
1139: dependence of $f,q$
1140: & 2 \\
1141: {\tt fqVarFracDZ} & Fraction of $f,q$ variance that due to $\Delta z$
1142: dependence & 0.5  
1143: \enddata
1144: \label{params3}
1145: \end{deluxetable}
1146: 
1147: 
1148: \subsection{Galaxy correlation coefficients}
1149: The correlation coefficient
1150: $r^g_{\alpha i}$ is, most generally, different at each subset
1151: and at each multipole $\ell$.  We model $r^g$ using the 
1152: ``$kz\Delta z$'' function form described in Appendix~\ref{functions}.
1153: These functions are polynomial in $\Delta z$, with the polynomial
1154: coefficients linearly interpolated
1155: from a grid in $(\ln k, \ln a)$ space.  This grid of
1156: polynomial coefficients replaces the $r^g_{\alpha i}$
1157: as parameters in the likelihood.
1158: 
1159: For the fiducial correlation
1160: coefficient, we interpolate smoothly between a linear and non-linear
1161: limit according to the value of $\Delta^2_{\rm lin}(k,z)=k^3P^m_{\rm
1162:   lin}(k,z)/2\pi^2$:
1163: \begin{equation}
1164: r^g_{\rm fid}(k,z) = {r^g_{\rm NL} + r^g_{\rm L} \over 2}
1165:  + { (r^g_{\rm NL} - r^g_{\rm L})  \over \pi} \tan^{-1} \left[
1166: { \ln \Delta_{\rm lin}(k,z) \over W} \right].
1167: \label{lnlinterp}
1168: \end{equation}
1169: The constant $W$ sets the width of the transition from the linear to
1170: nonlinear regime.  We set $W=1$ unless otherwise noted.
1171: 
1172: Each polynomial coefficient at each grid point is assigned an
1173: independent Gaussian prior.  These are selected to yield a preselected
1174: RMS variation $r^g_{\rm RMS}$. The RMS prior
1175: uncertainties are interpolated in $(k,z)$ space
1176: between linear and 
1177: non-linear limiting values $r^g_{\rm RMS,L}$ and $r^g_{\rm RMS,NL}$
1178: using the same functional form (\ref{lnlinterp}).
1179: 
1180: Table~\ref{params4} lists the program inputs needed to specify the
1181: galaxy bias and correlation models.
1182: 
1183: \subsection{Galaxy bias}
1184: We expect $b^g$ to vary quite strongly with $z$, as the more distant
1185: source galaxies are likely intrinsically very bright and highly
1186: biased. There may also be a strong dependence of $b^g$ on $\Delta z$
1187: because both $\Delta z$ and the bias may couple strongly to galaxy
1188: spectral type.  Variation with $k$ should be weaker.  We hence define $b^g$
1189: to be the sum of two functions,
1190: \begin{equation}
1191: b^g = b^g_{\rm coarse}(z,\Delta z) + b^g_{\rm fine}(k,z,\Delta z).
1192: \end{equation}
1193: The fiducial values are $b^g_{\rm coarse}=b^g_{\rm fid}, b^g_{\rm
1194:   fine}=0$.  
1195: 
1196: The coarse
1197: contribution is given a very weak prior, but can only vary slowly with
1198: $z$: $\Delta\ln a=0.5$ by default for the $b^g_{\rm coarse}$ grid nodes.
1199: 
1200: The $b^g_{\rm fine}$ function is interpolated between the same $(\ln
1201: k, \ln a)$ grid points as the correlation coefficient $r^g$. The prior
1202: on each $b^g_{\rm fine}$ node is arranged to give RMS uncertainty
1203: that is interpolated between linear and nonlinear limits $b^g_{\rm
1204:   RMS,L}$ and $b^g_{\rm RMS,NL}$ just as for $r^g$.
1205: 
1206: 
1207: \begin{deluxetable}{cll}
1208: \tablewidth{0pt}
1209: \tablecaption{Galaxy bias inputs to Fisher forecast}
1210: \tablehead{
1211: \colhead{Parameter name} &
1212: \colhead{Description} &
1213: \colhead{Default}
1214: }
1215: \startdata
1216: {\tt bg} & $b^g_{\rm fid}$, fiducial galaxy bias & 1.5 \\
1217: {\tt rgL} & $r^g_{\rm L}$, fiducial galaxy correlation coeff at linear
1218: limit & 0.9 \\
1219: {\tt rgNL} & $r^g_{\rm NL}$, fiducial galaxy correlation coeff at non-linear
1220: limit & 0.6 \\
1221: {\tt biasDlnk} & $\Delta\ln k$ interpolation grid step for bias
1222: models & 1.0 \\
1223: {\tt biasDlna} & $\Delta\ln a$ interpolation grid step for bias
1224: models & 0.1 \\
1225: {\tt brgRMSL} & $b^g_{\rm RMS,L}$ and $r^g_{\rm RMS,L}$ RMS prior
1226: variation for bias and correlation, linear limit & 0.05 \\
1227: {\tt brgRMSNL} & $b^g_{\rm RMS,NL}$ and $r^g_{\rm RMS,NL}$ RMS prior
1228: variation non-linear limit & 0.10 \\
1229: {\tt brVarFracDZ} & Fraction of $b^g,r^g$ variance due to $\Delta
1230: z$ dependence & 0.2 \\
1231: {\tt bgCoarseDlna} & $\Delta\ln a$ node grid spacing for $b^g_{\rm coarse}$
1232:  & 0.5 \\
1233: {\tt bgCoarseZRMS} & RMS prior variation of $b^g_{\rm coarse}$ at
1234: $\Delta z=0$ & 0.5 \\
1235: {\tt bgCoarseDZRMS} & RMS prior variation of $b^g_{\rm coarse}$ at
1236: fixed $z$ & 0.3 \\
1237: {\tt kgg} & $K^{gg}_{\rm fid}$, fiducial value of $K^{gg}(k,z)$
1238: cross-corr spec & 2. \\
1239: {\tt kggRMS} & $K^{gg}_{\rm RMS}$, prior variation on $K^{gg}(k,z)$ &
1240: 1.
1241: \enddata
1242: \label{params4}
1243: \end{deluxetable}
1244: 
1245: 
1246: \subsection{Intrinsic Alignments}
1247: The strength of intrinsic alignments are specified by the
1248: $b^\kappa_{\alpha i}$ and $r^\kappa_{\alpha i}$ values at each
1249: multipole $\ell$.  As for the galaxy density, the free-parameter count
1250: can be reduced by specifying parametric functions $b^\kappa,
1251: r^\kappa$ of $(k, z, \Delta 
1252: z)$ instead.  Each of these two functions modeled using the
1253: ``$kz\Delta z$'' form described in Appendix~\ref{functions}, just as
1254: for $r^g$.  The fiducial $b^\kappa_{\rm fid}$ and 
1255: $r^\kappa_{\rm fid}$ are
1256: taken as constant over the entire domain.  The RMS variation
1257: $b^\kappa_{\rm RMS}$ and $r^\kappa_{\rm RMS}$ in the
1258: priors of these functions are also taken to be constant over the domain.
1259: We take $b^\kappa_{\rm RMS}=|b^\kappa_{\rm fid}|$, because we
1260: expect that the best constraints on IA to always arise from self-calibration
1261: of WL surveys rather than through any external modeling or
1262: prior.  Roughly speaking, the IA measured by \citet{Mandelbaum06} for
1263: the SDSS population corresponds to $b^\kappa \approx -0.003$
1264: \citep{BridleKing}, which we will normally take as our fiducial model
1265: for IA.  Setting $b^\kappa_{\rm fid}=0$ turns off the IA systematic entirely.
1266: 
1267: Table~\ref{params5} lists the program inputs needed to specify the
1268: functional form for intrinsic alignments, the fiducial values, and the
1269: prior constraints.  Note that we assume the intrinsic-alignment
1270: functions to be defined on the same $(\ln k, \ln a)$ grid as the
1271: galaxy bias and covariance functions.
1272: 
1273: \begin{deluxetable}{cll}
1274: \tablewidth{0pt}
1275: \tablecaption{Intrinsic-alignment (IA) inputs to Fisher forecast}
1276: \tablehead{
1277: \colhead{Parameter name} &
1278: \colhead{Description} &
1279: \colhead{Default}
1280: }
1281: \startdata
1282: {\tt bk} & $b^\kappa_{\rm fid}$, fiducial intrinsic alignment &
1283: $-0.003$ \\
1284: {\tt rk} & $r^\kappa_{\rm fid}$, fiducial correlation between mass and
1285: intrinsic alignment & 0.7 \\
1286: {\tt rkRMS} & $r^\kappa_{\rm RMS}$, RMS prior variation in $r^\kappa$ & 0.2 \\
1287: {\tt kkk} & $K^{\kappa\kappa}_{\rm fid}$, fiducial value of
1288: $K^{\kappa\kappa}(k,z)$ & 1. \\
1289: {\tt kkkRMS} & $K^{\kappa\kappa}_{\rm RMS}$, RMS prior variance of $K^{\kappa\kappa}$
1290:  & 1.\\
1291: {\tt skg} & $s^{g\kappa}_{\rm fid}$, fiducial IA-bias
1292: cross-correlation & 0. \\
1293: {\tt skgRMS} & $s^{g\kappa}_{\rm RMS}$, RMS prior on IA-bias
1294: cross-correlation & 0.3 
1295: \enddata
1296: \label{params5}
1297: \end{deluxetable}
1298: 
1299: \subsection{Cross-correlation coefficients}
1300: The cross-correlation coefficients $r^{gg}_{\alpha\beta i},
1301: r^{g\kappa}_{\alpha\beta i}$, and $r^{\kappa\kappa}_{\alpha\beta i}$ are even
1302: more complex because each depends on $k, z$, plus two subsets' $\Delta
1303: z_{\alpha i}$ and $\Delta z_{\beta i}$.  We find it infeasible to
1304: construct nuisance-function templates spanning 4 dimensions.  We
1305: therefore simplify by first writing 
1306: \begin{equation}
1307: r^{g\kappa}_{\alpha \beta i}  =  r^g_{\alpha i} r^\kappa_{\beta i} 
1308:  + s^{g\kappa}_{\alpha \beta i} 
1309: \sqrt{\left[1-(r^g_{\alpha i})^2\right]
1310:       \left[1-(r^\kappa_{\beta i})^2\right]}.
1311: \end{equation}
1312: A value $|s^{g\kappa}_{\alpha \beta i}|\le 1$ is necessary (but not
1313: sufficient) to keep the
1314: mass-galaxy covariance 
1315: matrix from acquiring non-physical negative eigenvalues.  In principle
1316: the functional form of $s^{g\kappa}$ must vary over four
1317: dimensions, but we make the gross simplification that it is constant
1318: for the survey, since we expect this type of cross-correlation to have
1319: minimal effect on cosmological constraints.  The program thus requires
1320: simply a fiducial scalar $s^{g\kappa}$ and an RMS for its Gaussian
1321: prior.
1322: 
1323: The density-density cross-correlation $r^{gg}$ may have substantial
1324: impact on cosmological constraints, so we model it with more freedom,
1325: though not full 4-dimensional behavior. We set
1326: \begin{equation}
1327: \label{kggfunc}
1328: s^{gg}_{\alpha \beta i}  =  {\rm max}\left[ 0, 1 - K^{gg}(k,z_i) { |z_\alpha - z_\beta|
1329:     \over 2 \Delta z_{\rm max} } \right].
1330: \end{equation}
1331: This
1332: functional form for $s$ gives the most closely related subsets the
1333: highest covariance.    $\Delta z_{\rm max}$
1334: is the width of the redshift distribution within a set.
1335: The function $K^{gg}(k,z)$ adjusts how quickly the
1336: subsets decorrelate as their photo-z's diverge. 
1337: $K^{gg}(k,z)$ is implemented as the ``$kz$'' functional form described
1338: in Appendix~\ref{functions}, namely a linear
1339: interpolation between a grid of values in the
1340: $(\ln k, \ln a)$ plane.  We hence parameterize the
1341: 4-dimensional cross-correlation function by a 2d grid of $K^{gg}$
1342: nodal values.  These points are all given the same fiducial value
1343: $K^{gg}_{\rm fid}$.
1344: The $K^{gg}$ nodal values are given Gaussian priors
1345: to select a range of uncertainty $K^{gg}_{\rm RMS}$.  
1346: 
1347: The cross-correlation parameters $r^{\kappa\kappa}_{\alpha\beta i}$
1348: are similarly reduced from 4-dimensional behavior by defining 
1349: values $s^{\kappa\kappa}_{\alpha\beta i}$ that are set by the nodal
1350: values of a 2-dimensional function $K^{\kappa\kappa}(k,z)$ in complete
1351: analogy with \eqq{kggfunc}.
1352: 
1353: \subsection{An apology}
1354: This section on nuisance functions is obscure and lengthy, especially
1355: regarding the cross-correlations of galaxies and intrinsic alignments.
1356: It is, unfortunately, impossible to fully describe the likelihood of
1357: lensing survey data without invoking some model for all of these functions.
1358: 
1359: Previous work has avoided these messy details and functions by making
1360: many simplifications.
1361: Most have implicitly assumed that all the correlation coefficients are
1362: unity.  Most
1363: have ignored the intrinsic-alignment signal entirely, {\it i.e.}
1364: taking $b^\kappa=0$.  All previous analyses have considered the galaxy
1365: bias to be constant within a set, and if the multiplicative error has
1366: been considered, it has also been constant within a galaxy set.  Only
1367: a few analyses have allowed galaxy bias to vary with redshift
1368: \citep{Zhan06, BJ04}.  The most sophisticated treatment to date
1369: is that of \citet{HJ04}, who take all bias and correlation
1370: coefficients to derive from a halo model of galaxies.  The
1371: redshift-distribution parameters $p_{\alpha i}$ have, in the most
1372: ambitious analyses to date, been reduced to two-parameter (Gaussian)
1373: functions. \citet{MaBernstein} consider sum-of-Gaussian models.  These
1374: assumptions can all be implemented in the present 
1375: formalism if desired, but can also be relaxed to assess their impact
1376: on the cosmological constraints.
1377: 
1378: 
1379: \section{Tuning the forecast parameters}
1380: \label{tuneup}
1381: In this section we determine the values of bin widths and
1382: nuisance-function bandwidths that are needed for reliable extraction
1383: of maximum information from lensing surveys.  Unless otherwise noted,
1384: we will derive these parameters for a canonical survey with $f_{\rm
1385:   sky}=0.5$; an effective source density of 60 galaxies per arcmin$^2$
1386: with median redshift of 1.2; $\sigma_\gamma=0.24$; and
1387: Gaussian-distributed fiducial photo-z errors of $\sigma_z=0.04(1+z)$.
1388: Except as noted, we assume $N_{\rm spec}=10^7$ so that photo-z
1389: calibration errors are negligible, and also reduce the shear
1390: calibration RMS uncertainty to $10^{-4}$ to be negligible as well.
1391: Other inputs assume the default values given in
1392: Tables~\ref{params1}--\ref{params5}. 
1393: 
1394: The information content of a survey will be gauged using the DETF
1395: figure of merit \citep{DETF}: the Fisher matrix will be marginalized
1396: over all nuisance parameters, then the $D_i$ and $\delta F_i$
1397: variables projected onto a model obeying General Relativity with
1398: homogeneous dark energy of equation of state $w=w_0+w_a(1-a)$.  A prior
1399: representing expected Planck results is added (also from the DETF
1400: report), and we marginalize over $\{\omega_m, \omega_b, \omega_k,
1401: \omega_{\rm DE}, n_s, \Delta_\zeta\}$ to yield the Fisher matrix ${\bf
1402:   F}_w$ over $\{w_o, w_a\}$.  The DETF FoM is defined as $|{\bf
1403:   F}_w|^{1/2}$.  
1404: \subsection{Multipole bin size}
1405: Since we have excluded baryon oscillations from our transfer function,
1406: we expect to find little information in the detailed shape of the
1407: lensing or density power spectra.  The broad lensing kernel in
1408: redshift also smoothes away fine structure in the convergence.  So we
1409: expect the information content in the Fisher matrix to be independent of
1410: the multipole bin width $\Delta\log_{10}\ell$ below some modest
1411: value.  Larger values of $\Delta\log_{10}\ell$ reduce the complexity
1412: and execution time of the calculations, so we seek the maximum
1413: $\Delta\log_{10}\ell$ at which nearly all the lensing information is
1414: present.
1415: 
1416: Figure~\ref{dlogl} plots the DETF FoM of the lensing$+$density survey
1417: (plus spectroscopic redshift survey and Planck prior) vs
1418: $\Delta\log_{10}\ell$ for several candidate surveys. The top line is
1419: for a very optimistic survey: $n_{\rm eff}=100\,{\rm arcmin}^{-2}$,
1420: $N_{\rm spec}=10^7$, $q_{\rm RMS}=10^{-3}$, $f_{\rm RMS}=10^{-4}$,
1421: $b^g_{\rm RMS}=r^g_{\rm RMS}=0.01$, and $b^\kappa_{\rm fid}=10^{-3}$.
1422: By reducing the systematics and the shot noise to (unrealistically)
1423: low levels, we give the lensing survey the chance to extract maximal
1424: information.  We find that the FoM gains only 2\% for
1425: $\Delta\log_{10}\ell<0.3$. 
1426: 
1427: Other lines in the plot are FoM vs $\Delta\log_{10}\ell$ for weaker
1428: surveys, with $N_{\rm spec}=10^{4.5}$ and/or  $n_{\rm eff}=60\,{\rm arcmin}^{-2}$,
1429: $q_{\rm RMS}=0.1$, $b^g_{\rm RMS}=r^g_{\rm RMS}=0.1$, $b^\kappa_{\rm
1430:   fid}=-0.003$.  In these cases we also find that the DETF FoM
1431: increases by $<2$--3\% for $\Delta\log_{10}\ell<0.3$.
1432: 
1433: We adopt $\Delta\log_{10}\ell=0.3$ for all future use.
1434: \begin{figure}
1435: \plottwo{f1a.eps}{f1b.eps}
1436: \caption[]{{\em Left:} DETF Figure of Merit resulting from Fisher analyses of
1437:   several weak lensing surveys, plotted against multipole bin width
1438:   $\Delta\log_{10}\ell$ of the analysis.  Each solid line plots a
1439:   particular survey scenario (see text for details).  The dashed lines
1440:   are horizontal, to 
1441:   help the eye judge the information degradation as we increase
1442:   $\Delta\log_{10}\ell$.  We conclude that $\Delta\log_{10}\ell\le
1443:   0.3$ retains $>97\%$ of the information for surveys of any quality
1444:   level. {\em Right:} Value of DETF FoM for the default survey vs
1445:   range of multipole used.  Choice of upper bound is more critical
1446:   than choice of lower bound on $\ell$.
1447: }
1448: \label{dlogl}
1449: \end{figure}
1450: 
1451: \subsection{Multipole range}
1452: On the right-hand side of Figure~\ref{dlogl} we plot the DETF FoM for
1453: various ranges of $\ell$.  In this study we assume a space based
1454: survey obtaining $n_{\rm eff}=60\,{\rm arcmin}^{-2}$ over $f_{\rm
1455:   sky}=0.5$.  The photo-z and shear calibration systematics are held
1456: negligible with priors but other systematics (intrinsic alignment,
1457: etc.) have default priors.
1458: 
1459: We see that the choice of $\ell_{\rm max}$ has a strong influence, as
1460: moving from $10^3$ to $10^4$ changes the FoM by $1.7\times$.  We note
1461: this is true even though we have included uncertainty in the
1462: theoretical power spectrum at high $k$ values, showing that there is
1463: still information to be gained when the theory is incomplete.  We
1464: find, in fact, that our default power-spectrum theory uncertainty of
1465: $f_{\rm Zhan}=0.5$ leads to only 6\% degradation of the DETF FoM
1466: relative to
1467: an assumption of {\em zero} uncertainty in the theory, even when
1468: $\ell_{\rm max}=10^4$.  
1469: 
1470: Unfortunately our assumption of Gaussian statistics will fail by
1471: $\ell=10^4$ \citet{CoorayHu01,LeePen}, rendering the Fisher calculation less
1472: reliable.  We will restrict our analysis to
1473: $\ell<10^{3.5}$, but additional study of the effect of
1474: non-Gaussian statistics is clearly needed.
1475: 
1476:  The flat-sky and Limber approximations will fail
1477: at low $\ell$, but the choice of $\ell_{\rm min}$ appears less
1478: critical to the $w_0/w_a$ information content, so will retain the
1479: $\ell>10$ bound in our analyses.
1480: 
1481: \subsection{Scale resolution for nuisance functions}
1482: We require choice of node spacing $\Delta\ln k$ in the nuisance
1483: functions for the power-spectrum theory errors, the calibration errors
1484: $f$ and $q$, and the bias/correlation parameters $b^g_{\rm fine}, r^g,
1485: b^\kappa, r^\kappa, K^{gg},$ and $K^{\kappa\kappa}$.  We set
1486: $\Delta\ln k$ to be equal for all nuisance functions, and find the
1487: value which minimizes the DETF FoM as the ``most damaging'' scale of
1488: variation.  We examine the default case described above, for several values
1489: of the
1490: shear calibration prior $f_{\rm RMS}$ and photo-z calibration size
1491: $N_{\rm spec}$.
1492: 
1493: Figure~\ref{dlnk} shows $\Delta\ln k\approx 0.7$--1 yields minimum
1494: information for fixed RMS priors, but the dependence is very weak.
1495: The FoM varies by only 7\% over the range $0.5<\Delta\ln k<1.5$.  We
1496: henceforth adopt $\Delta\ln k=1$.  Perhaps not surprisingly, this
1497: makes the nuisance functions have $\approx 1$ independent node in each
1498: multipole bin of $\Delta\log_{10}\ell=0.3$.
1499: 
1500: \begin{figure}
1501: \plottwo{f2a.eps}{f2b.eps}
1502: \caption[]{{\em Left:} DETF FoM vs $\Delta\ln k$, the spacing of nuisance-function
1503:   nodes in the length-scale axis.  We examine surveys with varying
1504:   strengths of priors on shear calibration and photo-z calibration.
1505:   For this plot, each is normalized to the value at $\Delta\ln k=0.7$
1506:   in order to show the (weak) dependence of FoM on $\Delta\ln k$ when
1507:   the prior on RMS nuisance-function fluctuations are held fixed.  We
1508:   adopt $\Delta\ln k=1$ as the most conservative bandwidth for
1509:   nuisance-function variation with scale. {\em Right:} DETF FoM vs
1510:   $\Delta \ln a$, the spacing of nodes in redshift for the bias,
1511:   correlation, and intrinsic-alignment nuisance functions.  The FoM is
1512:   quite sensitive to this choice, and allowing the bias/IA to vary on
1513:   $\Delta\ln a=0.1$ scales is most damaging to cosmological inference.
1514: }
1515: \label{dlnk}
1516: \end{figure}
1517: 
1518: \subsection{Redshift resolution for nuisance functions}
1519: All of the nuisance functions are dependent on $z$.  We next
1520: investigate the redshift node spacing $\Delta\ln a$ at which the
1521: nuisance functions are most damaging to the DETF FoM.  We find that
1522: the FoM is insensitive to the $\Delta\ln a$ of the power-spectrum
1523: theory errors.  The value of $\Delta\ln a$ for the calibration
1524: functions $f$ and $q$ that minimizes the FoM depends upon the strength
1525: of the prior.  The choice $\Delta\ln a=0.5$ produces a FoM
1526: that is within 2\% of the minimum, however, so we fix this value for
1527: the theory-error and calibration nuisance functions.
1528: 
1529: The redshift freedom given to the bias and intrinsic-alignment
1530: nuisance functions has a strong impact on the DETF information
1531: content.  Figure~\ref{dlnk} illustrates that, at fixed RMS prior
1532: variation, models with freedom to vary on rather fine scales,
1533: $\Delta\ln a = 0.1$ are most damaging to cosmological
1534: information.
1535: 
1536: 
1537: \section{Conclusion}
1538: The core of this paper are the expressions (\ref{gg1})--(\ref{kk1})
1539: for the two-point 
1540: correlation matrix of the lensing and density observable multipoles produced by
1541: a typical lensing survey.  This was derived under a very limited set
1542: of assumptions: a homogeneous and isotropic 4-dimensional metric
1543: Universe with scalar perturbations; plus the weak-lensing limit,
1544: the Limber and Born approximations, and an approximation that lensing
1545: magnification bias and intrinsic density fluctuations are additive.
1546: The last four assumptions could be relaxed at the expense of
1547: computational complexity.  We thus hope that data analyses based on
1548: this framework could be used to constrain a wide variety of potential
1549: explanations for the acceleration phenomenon, including gravity
1550: modifications as well as new fields in the Universe.  In the limit of
1551: Gaussian fluctuation fields, the two-point information is a complete
1552: description of the likelihood and hence can be used to construct
1553: Fisher matrices or analyze data.
1554: As currently configured, the analysis yields the survey's ability to
1555: constrain the distance function $D(z)$ and linear growth function
1556: $g_\phi(z)$, without reference to particular dark-energy models.  It
1557: would be straightforward to implement scale-dependent 
1558: linear-growth functions.
1559: 
1560: This framework subsumes all of the
1561: information (up to 2-point level) that is likely to be obtained from
1562: lensing observations: density-density, lensing-density, and
1563: lensing-lensing correlations, plus redshift distributions from
1564: unbiased spectroscopic surveys (\S\ref{speclike}).  Furthermore it
1565: allows for the most 
1566: important expected forms of systematic error: photo-z calibration
1567: errors, shear and magnification-bias calibration errors, intrinsic
1568: alignments, and inaccuracies in power-spectrum theory.  Systematics
1569: that are additive to shear (\eg uncorrected PSF ellipticity) or to
1570: density (\eg uncorrected foreground extinction) have not been
1571: included.  We have not done so since the additive errors could, in
1572: principle, exhibit almost any arbitrary signature in the covariance
1573: matrix of the observables.  Hence a completely general model for
1574: additive errors would be degenerate with almost all other signals.
1575: For the additive systematics, it is better to determine the level at
1576: which they would bias the cosmological results than to attempt to fit
1577: a model.  \citet{AmaraRefregier} is a good example of this approach.
1578: 
1579: Since the analysis framework is independent of models for dark energy,
1580: gravity, power-spectrum evolution, or galaxy bias, we get a
1581: stripped-down look at what parameters are truly constrained by the
1582: data, and what nuisance functions must be modeled in order to extract
1583: the cosmological information.  There is a substantial suite of biases
1584: and correlation functions involved in understanding the full survey
1585: data.  In other work these have been ignored, or have been quantified
1586: by reference to halo occupation models \citep{HJ04,CvdBMLMY}.
1587: Here we introduce generic functions for bias and calibration nuisance
1588: functions that are not based on any
1589: particular physical model.  
1590: 
1591: We implement one possible model for the evolution of the
1592: lensing-potential power spectrum, based on General Relativity but
1593: allowing for failure of the growth equation.  It is straightforward to
1594: implement other potential deviations from General Relativity.  In the
1595: current implementation, the end result of the Fisher analysis is a
1596: forecast of the ability to constrain the functions $D_A(z)$ and
1597: $g_\phi(z)$.
1598: 
1599: Since the analysis must be discretized in redshift and angular scale
1600: in order to be feasible, we investigated the bin sizes or bandwidths
1601: of nuisance functions that should be chosen.  We find that $\approx 3$
1602: bins per decade of angular scale suffice to extract all information
1603: (apart from baryon acoustic oscillations), and that nuisance functions
1604: should be specified no finer than this.  Nuisance functions for
1605: power-spectrum theory errors and for shear and magnification-bias
1606: calibration errors can be specified coarsely in redshift space
1607: ($\Delta \ln a \approx 1$), but the galaxy biases, correlations, and
1608: intrinsic-alignments must be modeled with potentially finer structure
1609: in redshift ($\Delta \ln a \approx 0.1$) to immunize against potential
1610: astrophysical systematics.
1611: 
1612: In future papers we will use this framework and its implementation to
1613: investigate the requirements for spectroscopic calibration of
1614: photo-z's in large lensing surveys, and other practical issues.  As a
1615: simple first application of our framework, we have shown here that
1616: power-spectrum theory uncertainty does not significantly degrade the
1617: cosmological power of a nominal lensing survey at $10<\ell<10^4$.
1618: Non-Gaussian statistics are a much more important factor to consider.
1619: 
1620: {\tt C++} Code to implement Fisher forecasting using this framework has been
1621: written and runs quickly on desktop computers despite the large number
1622: of free parameters in these general models.  Interested parties should
1623: contact the author for access to the code.
1624: 
1625: \acknowledgments
1626: This work is supported in this work by grant AST-0607667 from the
1627: National Science Foundation, Department of Energy grant
1628: DOE-DE-FG02-95ER40893, and NASA grant 
1629: BEFS-04-0014-0018.  I thank Bhuvnesh Jain, Zhaoming Ma, Chris Hirata,
1630: Ravi Sheth, Hu Zhan,
1631: and the members of the Dark Energy Task Force for helpful
1632: conversations during the long gestation of this work.
1633: 
1634: 
1635: \appendix
1636: \section{Parametric functional forms for nuisance variables}
1637: \label{functions}
1638: In modeling an experiment, we often encounter some systematic error
1639: associated with a nuisance variable $f$ about which we have little {\it a
1640:   priori} knowledge.  We would like to fit some parametric form to
1641: this variable, but would like a form that is flexible enough to
1642: describe any ``reasonable'' behavior the function might exhibit.  We
1643: also want to conveniently relate the number and prior probabilities
1644: for the parameters to the kind of variation that $f$ might exhibit.
1645: A parametric description of some nuisance function $f$ defined over a variable
1646: $x\in[-1,1]$ would ideally have the following properties:
1647: \begin{enumerate}
1648: \item $f(x)$ has a variable number $N$ of controlling parameters $\{a_0,
1649:   a_2, \ldots, a_{N-1}\}$ such that any continuous differentiable function
1650:   $F(x)$ can be approximated to any desired accuracy with a
1651:   sufficiently large choice of $N$.
1652: \item We can draw $\{a_j\}$ from independent
1653:   Gaussian distributions of zero mean and widths $\{\sigma_j\}$, with the
1654:   result that ${\rm Var}[f(x)]$ is
1655:   independent of $x$.  In other words, the nuisance value $f$ has a
1656:   uniform and well-determined variance when we apply a simple diagonal
1657:   Gaussian prior to the parameter set $\{a_j\}$.
1658: \end{enumerate}
1659: A Fourier decomposition, $f = \sum (a_j \sin j\pi x + b_j \cos j\pi
1660: x)$, exhibits these qualities, but converges poorly when $f(-1)\ne
1661: f(+1)$.  
1662: 
1663: \subsection{Linearly interpolated functions}
1664: Another approach is linear interpolation: choosing a spacing $\Delta x
1665: = 2/(N-1)$, we define $a_i$ as the value of $f$ at $x_i=i\Delta x
1666: -1$.  At some other $x_i<x<x_{i+1}$, we define
1667: \begin{equation}
1668: f(x) = wa_i + (1-w) a_{i+1}, \qquad w={x_{i+1} - x \over \Delta x}.
1669: \end{equation}
1670: If we assign an independent Gaussian prior of width $\sigma_a$ to each
1671: $a_i$, then by definition we have ${\rm Var} [f(x)] = \sigma_a^2$ if $x$
1672: coincides with a node.  But the variance of $f$ is not quite
1673: homogeneous: it drops to ${\rm Var} [f(x)] = \sigma_a^2/2$ when $x$ is
1674: halfway between two nodes.  If we want the {\em mean} variance of
1675: $f(x)$ over the interval $x\in[-1,1]$ to equal $\sigma_f^2$, then the
1676: variance of the prior on each node needs to be $\sigma_a^2=3\sigma_f^2/2$.
1677: 
1678: \subsection{Legendre polynomials}
1679: Polynomial expansions are also commonly used to model nuisance
1680: functions. The simplistic form $f(x)=\sum a_i x^i$ results in
1681: extremely non-uniform variance for $f$ with diagonal prior on
1682: $\{a_i\}$ and is hence inappropriate for our purpose.  A better
1683: choice is to expand in Legendre polynomials $P_n(x)$, which are orthogonal over
1684: $[-1,1]$.  We define
1685: \begin{equation}
1686: f(x) = \sum_{i=0}^{N-1} a_i P_i(\nu x).
1687: \end{equation}
1688: Recall that the Legendre polynomials satisfy $P_0=1, P_1=x,
1689: (n+1)P_{n+1}=(2n+1)xP_n-nP_{n-1}$.
1690: 
1691: We wish to choose priors $\{\sigma_i\}$ on $\{a_i\}$ that cause
1692: the variance of $f(x)$ be as uniform as possible for
1693: $x\in[-1,1]$.  We have not found a way to attain perfect uniformity in $x$
1694: with polynomial interpolation, however the following scheme gets
1695: usefully close.  We discover numerically that
1696: \begin{equation}
1697: \lim_{N\rightarrow\infty} {1 \over N} \sum_{i=0}^{N-1} P_n^2(\nu x) (2i+1)
1698: = {2 \over \pi} \left(1-\nu^2x^2\right)^{-1/2}.
1699: \end{equation}
1700: This implies that, if we set $\sigma_i = \sqrt{(2i+1)\pi/2N}$, then in
1701: the limit of large $N$ we will obtain ${\rm Var} [f(x)] =
1702: \left(1-\nu^2x^2\right)^{-1/2}.$  For $\nu=1$ this would diverge at
1703: the ends of our nuisance function's interval.  If, however, we choose
1704: $\nu=0.9$, the RMS is only $\approx1.5\times$ larger at the
1705: endpoints than at $x=0$.  Over the $[-1,1]$ interval, the mean
1706: variance is $(\sin^{-1} \nu)/\nu$.  Hence if we wish to have a
1707: function with $\langle {\rm Var} f(x) \rangle_x = \sigma^2_f$, we set
1708: the priors on the Legendre coefficients to be
1709: \begin{equation}
1710: \sigma_i = \sigma_f \sqrt{(2i+1)\pi \nu \over 2N \sin^{-1}\nu}.
1711: \end{equation}
1712: 
1713: \subsection{Standard multidimensional functions}
1714: The Fourier, linear-interpolation, and Legendre-polynomial functional
1715: forms can each be extended to $>1$ dimensions in straightforward
1716: fashion.  In the lens modeling, we need functions of these
1717: dimensions:
1718: \begin{enumerate}
1719: \item Comoving wavenumber or physical scale: $x_1=\ln k$;
1720: \item Redshift $z$: more precisely we will use the variable $x_2=\ln
1721:   (1+z)$;
1722: \item Photometric redshift error $\Delta z$; more precisely, our code
1723:   uses the variable $x_3 = \Delta \ln (1+z) = \ln[(1+z_\alpha)/(1+z_i)]$ for
1724:   subset $\alpha i$.
1725: \end{enumerate}
1726: 
1727: \subsubsection{$kz$ form}
1728: For functions over the $(x_1,x_2)$ space we use a simple
1729: two-dimensional version of interpolation between values on a
1730: rectangular grid.  The power-spectrum theory error $\delta\ln P$ uses
1731: this form, as do the $K^{gg}(k,z)$ and $K^{\kappa\kappa}(k,z)$
1732: nuisance functions.  The only complication of note is that the nodal
1733: point $a_{ij}$ should have a prior with variance
1734: $\sigma^2_a=(3/2)^2\sigma^2_f$ if the output function is to have
1735: variance $\sigma^2_f$.  The factor of $(3/2)^2$ is needed to
1736: counteract the reduced variance when interpolating between grid points.
1737: 
1738: The $kz$ nuisance function is specified by:
1739: \begin{itemize}
1740: \item the spacing $\Delta x_1 = \Delta \ln k$ of the nodes for
1741:   linearly interpolation in $x_1$;
1742: \item the spacing $\Delta x_2 = \Delta \ln (1+z)$ of the nodes for
1743:   linearly interpolation in $x_2$;
1744: \item the RMS variation $\sigma_f$ of the function 
1745:   allowed under the prior, which can depend on $x_1$ and $x_2$.
1746: \item the fiducial dependence of $f$ on $x_1$ and $x_2$.
1747: \end{itemize}
1748: 
1749: \subsubsection{$z\Delta z$ form}
1750: For nuisance variables over the $(x_2,x_3)$ space ($z$ and $\Delta
1751: z$), we adopt the following strategy:
1752: we choose to have variation over $x_3$ be described by
1753: polynomials since we usually define the range of non-catastrophic
1754: photo-z errors to be 
1755: bounded to some range $|x_3| \le \Delta_{\rm max}$.  We define
1756: the ``$z\Delta z$'' functional form as follows:
1757: \begin{equation}
1758: f(x_2,x_3) = \sum_{i=0}^{N-1} a_i(x_2) P_i(\nu x_3 / \Delta_{\rm max}).
1759: \end{equation}
1760: The Legendre coefficients are in turn defined to be linearly
1761: interpolated between a series of values $a_{ij}$ at redshift nodes
1762: $\{\ln (1+z_j)\}$.  The $a_{ij}$ become parameters of the model.  
1763: 
1764: The fiducial and prior values for the $i=0$ terms (constant in $x_3$)
1765: are treated differently than the $i>0$ terms.  We might, for
1766: example, expect some nuisance functions to vary strongly with $x_2$
1767: (nominal redshift) but only slightly with $x_3$ (photo-z error) at
1768: fixed $x_2$.  
1769: 
1770: We specify the total RMS fluctuation in $f$ allowed
1771: by the prior to be $\sigma_f$. But we also specify the fraction {\tt
1772:   VarFracDZ} of the variance that is due to dependence on $x_3$. If
1773: we define
1774: \begin{equation}
1775: \bar f(x_2) \equiv \int_{-\Delta_{\rm max}}^{+\Delta_{\rm max}} dx_3\,f(x_2,x_3) /
1776: 2\Delta_{\rm max},
1777: \end{equation}
1778: then we aim to achieve
1779: \begin{eqnarray}
1780: {\rm Var} [\bar f(x_2)] & = & \sigma^2_f \left(1-{\tt VarFracDZ}\right) \\
1781: {\rm Var} \left[f(x_2,x_3)-\bar f(x_2)\right] & = & \sigma^2_f \left({\tt
1782:     VarFracDZ}\right).
1783: \end{eqnarray}
1784: 
1785: This is achieved approximately by setting the priors on the constant terms as
1786: \begin{equation}
1787: \label{ap1}
1788: \sigma^2_{0j} = \frac{3}{2} \sigma^2_f \left(1-{\tt VarFracDZ}\right) 
1789: \end{equation}
1790: and the $x_3$-dependent terms as
1791: \begin{equation}
1792: \label{ap2}
1793: \sigma^2_{ij} = \frac{3}{2} \sigma^2_f \left({\tt VarFracDZ}\right) 
1794: \sqrt{(2i+1)\pi \nu \over 2N \sin^{-1}\nu}.
1795: \end{equation}
1796: 
1797: The RMS prior variation $\sigma_f$ can be made a function of $x_2$
1798: without loss of generality.
1799: We typically take fiducial values of $a_{ij}=0$ for $i>0$ in our
1800: nuisance functions, \ie no fiducial dependence upon $\Delta z$.
1801: 
1802: To summarize, the $z\Delta z$ nuisance function is specified by:
1803: \begin{itemize}
1804: \item the order $N$ of the polynomial in $x_3$;
1805: \item the maximum range $\Delta_{\rm max}$ of applicability in the
1806:   $x_3$ axis;
1807: \item the spacing $\Delta x_2 = \Delta \ln (1+z)$ of the nodes for
1808:   linearly interpolation in $x_2$;
1809: \item the RMS variation $\sigma_f$ of the function allowed under the
1810:   prior, which can depend on $x_2$;
1811: \item the fraction {\tt VarFracDZ} of this variance that is due to
1812:   $x_3$ ($\Delta z$) dependence;
1813: \item the fiducial dependence of $f$ on $x_2$.
1814: \end{itemize}
1815: 
1816: \subsubsection{$kz\Delta z$ form}
1817: The bias and correlation coefficients can, most generally, depend on
1818: scale ($x_1$) as well as subset $(x_2,x_3)$, so we generalize to the
1819: ``$kz\Delta z$'' functional form:
1820: \begin{equation}
1821: f(x_1,x_2,x_3) = \sum a_i(x_1,x_2) P_i(x_3).
1822: \end{equation}
1823: The coefficients
1824: $a_i$ are linearly interpolated
1825: between nodes in the two-dimensional space $(x_1,x_2)$.  Thus the free
1826: parameters of this model become the Legendre coefficients $a_{ijk}$.
1827: As for the $z\Delta z$ function, we specify the prior by the overall
1828: mean RMS variation $\sigma_f$ (which can be a function of $x_2$ and
1829: $x_3$), plus the {\tt VarFracDZ} specifying how much of the variance
1830: is manifested as dependence on $\Delta z$.  The formulae for the
1831: priors $\sigma_{ijk}$ on the nodal coefficients are derived exactly as
1832: in Equations~(\ref{ap1}) and (\ref{ap2}), except that we now need
1833: factors of $(3/2)^2$ to account for the reduced variance when
1834: interpolating in two dimensions.
1835: 
1836: The $kz\Delta z$ function thus requires all of the specifications as
1837: the $z\Delta z$ function, plus a spacing $\Delta\ln k$ for nodes in
1838: the $x_1$ axis.
1839: 
1840: \newpage
1841: \begin{thebibliography}{}
1842: 
1843: \bibitem[Albrecht et al.(2006)]{DETF} 
1844: Albrecht, A., et al.\ 2006, Report of the Dark Energy Task Force,
1845: arXiv:astro-ph/0609591  
1846: 
1847: \bibitem[Amara \& Refregier(2007)]{AmaraRefregier}
1848: Amara, A., \& Refregier, A.\ 2007, arXiv:0710.5171 
1849: 
1850: \bibitem[Bernstein(2006)]{curvature} 
1851: Bernstein, G.\ 2006, \apj, 637, 598 
1852: 
1853: \bibitem[Bernstein \& Jain(2004)]{BJ04}
1854: Bernstein, G. \& Jain, B., 2004, \apj, 600, 17 [BJ04]
1855: 
1856: \bibitem[Bernstein \& Norberg(2002)]{BernsteinNorberg}
1857: Bernstein, G.~M., \& Norberg, P.\ 2002, \aj, 124, 733 
1858: 
1859: \bibitem[Blandford et al.(1991)]{Blandford91} 
1860: Blandford, R.~D., Saust, A.~B., Brainerd, T.~G., \& Villumsen, J.~V.\
1861: 1991, \mnras, 251, 600  
1862: 
1863: \bibitem[Bridle \& King(2007)]{BridleKing} 
1864: Bridle, S., \& King, L.\ 2007, New Journal of Physics, 9, 444 
1865: 
1866: \bibitem[Cacciato \etal(2008)]{CvdBMLMY}
1867: Cacciato, M. \etal\ (2008), arXiv:0807.4934
1868: 
1869: \bibitem[Cooray \& Hu(2001)]{CoorayHu01} 
1870: Cooray, A., \& Hu, W.\ 2001, \apj, 554, 56 
1871: 
1872: \bibitem[Croft \& Metzler(2000)]{CroftMetzler} 
1873: Croft, R.~A.~C., \& Metzler, C.~A.\ 2000, \apj, 545, 561 
1874: 
1875: \bibitem[Eisenstein \& Hu(1999)]{EH} Eisenstein, D.~J., \& 
1876: Hu, W.\ 1999, \apj, 511, 5 
1877:  
1878: \bibitem[Faltenbacher et al.(2007)]{FLMvdBYJPM} Faltenbacher, A., 
1879: Li, C., Mao, S., van den Bosch, F.~C., Yang, X., Jing, Y.~P., Pasquali, A., 
1880: \& Mo, H.~J.\ 2007, \apjl, 662, L71 
1881: 
1882: \bibitem[Goldberg \& Bacon(2005)]{BaconFlex} 
1883: Goldberg, D.~M., \& Bacon, D.~J.\ 2005, \apj, 619, 741 
1884: 
1885: \bibitem[Hennawi \& Spergel(2005)]{Hennawi}
1886: Hennawi, J. F. \& Spergel,  D. N.\ 2005, \apj, 624, 59
1887: 
1888: \bibitem[Heymans et al.\@(2006)]{STEP1}
1889: Heymans, C., et al., 2006, \mnras\ 368 1323
1890: 
1891: \bibitem[Hirata et al.(2004)]{Hetal04} 
1892: Hirata, C.~M., et al.\ 2004, \mnras, 353, 529
1893: 
1894: % Intrinsic-alignment paper:
1895: \bibitem[Hirata \& Seljak(2004)]{Hirata2} 
1896: Hirata, C.~M., \& Seljak, U.\ 2004, \prd, 70, 063526 
1897: 
1898: %Calibration uncertainties:
1899: \bibitem[Hirata \& Seljak(2003)]{HS03}
1900: Hirata, C.~M., \& Seljak, U.\ 2003, \prd, 67, 043001 
1901: 
1902: \bibitem[Hu(1999)]{Hu} 
1903: Hu, W.\ 1999, \apjl, 522, L21 
1904: 
1905: \bibitem[Hu \& Jain(2004)]{HJ04}
1906: Hu, W, \& Jain, B., 2004, \prd, 70, 043009 
1907: 
1908: \bibitem[Hu \& Okamoto(2002)]{HO02}
1909: Hu, W. \& Okamoto, T.\ 2002, \apj, 574, 566 
1910: 
1911: \bibitem[Huterer et al.\@(2006)]{HTBJ}
1912: Huterer, D., Takada, M., Bernstein, G., \& Jain, B. 2006, \mnras, 366 101
1913: 
1914: \bibitem[Ishak \etal(2004)]{Ishak04}
1915: Ishak, M., Hirata, C. M., McDonald, P., \& Seljak, U. 20004, \prd, 69,
1916: 083514
1917: 
1918: \bibitem[Jain(2002)]{Jainmag} 
1919: Jain, B.\ 2002, \apjl, 580, L3 
1920: 
1921: \bibitem[Jain \& Seljak(1997)]{JS97} 
1922: Jain, B., \& Seljak, U.\ 1997, \apj, 484, 560 
1923: 
1924: \bibitem[Jain \& Taylor(2003)]{JainTaylor}
1925: Jain, B., \& Taylor, A.\ 2003, \prl, 91, 141302 
1926: 
1927: \bibitem[Jain et al.(2007)]{ConnollyJain} 
1928: Jain, B., Connolly, A., \& Takada, M.\ 2007, Journal of Cosmology and
1929: Astro-Particle Physics, 3, 13  
1930: 
1931: \bibitem[Jing \etal(2006)]{Jing}
1932: Jing, Y. P., Zhang, P., Lin, W. P., Gao, L., \& Springel, V. 2006,
1933: \apjl, 640, L119
1934: 
1935: \bibitem[Kaiser(1992)]{Kaiser92} 
1936: Kaiser, N.\ 1992, \apj, 388, 272 
1937: 
1938: \bibitem[King \& Schneider(2003)]{KingSchneider} 
1939: King, L.~J., \& Schneider, P.\ 2003, \aap, 398, 23 
1940: 
1941: \bibitem[Knox, Song, \& Tyson(2006)]{KST}
1942: Knox, L., Song, Y.-S., \& Tyson, J. A. 2006, \prd, 74, 023512 
1943: 
1944: \bibitem[Komatsu et al.(2008)]{WMAP5} 
1945: Komatsu, E., et al.\ 2008, arXiv:0803.0547 
1946: 
1947: \bibitem[Lee \& Pen(2008)]{LeePen}
1948: Lee, J. \& Pen, U.-L.\ 2008, arXiv:0807.1538
1949: 
1950: \bibitem[Ma \& Bernstein(2008)]{MaBernstein}
1951: Ma, Z., \& Bernstein, G.\ 2008, \apj, 682, 39 
1952: 
1953: \bibitem[Ma, Hu, \& Huterer(2006)]{MaHutererHu}
1954: Ma, Z., Hu, W., \& Huterer, D.\ 2006, \apj, 636, 21 
1955: 
1956: \bibitem[Mandelbaum et al.(2006)]{Mandelbaum06} 
1957: Mandelbaum, R., Hirata, C.~M., Ishak, M., Seljak, U., 
1958: \& Brinkmann, J.\ 2006, \mnras, 367, 611 
1959: 
1960: \bibitem[Marian \& Bernstein(2006)]{Laura}
1961: Marian, L., \& Bernstein, G.~M.\ 2006, \prd, 73, 123525 
1962: 
1963: \bibitem[Metcalf \& White(2007)]{MW07} 
1964: Metcalf, R.~B., \& White, S.~D.~M.\ 2007, \mnras, 381, 447 
1965: 
1966: \bibitem[Miralda-Escud\'{e}(1991)]{Jordi}
1967: Miralda-Escud\'{e}, J.\ 1991, \apj, 380, 1
1968: 
1969: \bibitem[Newman(2008)]{Newman} 
1970: Newman, J.~A.\ 2008, arXiv:0805.1409 
1971: 
1972: % 21 cm lensing
1973: \bibitem[Pen(2004)]{Pen04} 
1974: Pen, U.-L.\ 2004, New Astronomy, 9, 417 
1975: 
1976: \bibitem[Schneider et al.(2006)]{SKZC} 
1977: Schneider, M., Knox, L., Zhan, H., \& Connolly, A.\ 2006, \apj, 651, 14 
1978: 
1979: \bibitem[Smith \etal(2002)]{Smith}
1980: Smith, R. E., \etal, 2003, MNRAS, 341, 1311 
1981: 
1982: \bibitem[Song \& Knox(2004)]{SK04}
1983: Song, Y.-S. \& Knox, L. 2004, 
1984: 
1985: \bibitem[Takada \& Jain(2004)]{TJ04}
1986: Takada, M., \& Jain, B.\ 2004, \mnras, 348, 897 
1987: 
1988: \bibitem[Tegmark, Taylor, \& Heavens(1997)]{TTH97}
1989: Tegmark, M., Taylor, A., Heavens, A., 1997, \apj, 480, 22
1990: 
1991: \bibitem[Wang, Khoury \& Haiman(2004)]{Wang}
1992: S. Wang, J. Khoury, Z. Haiman et al.\ 2004, \prd, 70, 123008
1993: 
1994: \bibitem[Zahn \& Zaldarriaga(2006)]{ZZ06} 
1995: Zahn, O., \& Zaldarriaga, M.\ 2006, \apj, 653, 922 
1996: 
1997: \bibitem[Zhan(2006)]{Zhan06}
1998: Zhan, H.\ 2006, Journal of Cosmology and Astro-Particle Physics, 8, 8 
1999: 
2000: \bibitem[Zhan \& Knox(2004)]{ZhanKnox}
2001: Zhan, H., \& Knox, L.\ 2004, \apjl, 616, L75 
2002: 
2003: \bibitem[Zhang, Hui, \& Stebbins(2003)]{Zhang}
2004: Zhang, J., Hui, L., \& Stebbins, A. 2003, astro-ph/0312348
2005: 
2006: \end{thebibliography}
2007: 
2008: \end{document}
2009: