astro-ph0603844/ms.tex
1: %\documentclass[12pt, preprint]{aastex}
2: \documentclass{emulateapj}
3: 
4: %\usepackage{float}
5: %\usepackage{amsmath}
6: %\usepackage{epsf} 
7: %\usepackage{epsfig,floatflt}
8: %\usepackage{graphics}
9: %\usepackage{multirow}
10: 
11: 
12: \newcommand{\etal}{et~al.\ }
13: \newcommand{\etc}{etc.\ }
14: \newcommand{\ie}{i.e.,\ }
15: \newcommand{\eg}{e.g.,\ }
16: 
17: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
18: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
19: 
20: \begin{document}
21: 
22: \title{Fast and Efficient Template Fitting of Deterministic Anisotropic Cosmological Models Applied to {\em WMAP} Data}
23: \author{T. R. Jaffe\altaffilmark{1}, A. J. Banday\altaffilmark{1},
24:   H. K. Eriksen\altaffilmark{2}, K. M. G\'orski\altaffilmark{3},
25:   F. K. Hansen\altaffilmark{4} }
26: 
27: 
28: \altaffiltext{1}{Max-Planck-Institut f\"ur Astrophysik,
29: Karl-Schwarzschild-Str.\ 1, Postfach 1317, D-85741 Garching bei
30: M\"unchen, Germany; tjaffe@MPA-Garching.MPG.DE, 
31: banday@MPA-Garching.MPG.DE.}
32: 
33: \altaffiltext{2}{Institute of Theoretical Astrophysics, University of
34: Oslo, P.O.\ Box 1029 Blindern, N-0315 Oslo, Norway; Centre of
35: Mathematics for Applications, University of Oslo, P.O.\ Box 1053
36: Blindern, N-0316 Oslo; Jet Propulsion Laboratory, M/S 169/327, 4800
37: Oak Grove Drive, Pasadena CA 91109; California Institute of
38: Technology, Pasadena, CA 91125; h.k.k.eriksen@astro.uio.no} 
39: 
40: \altaffiltext{3}{JPL, M/S 169/327, 4800 Oak Grove Drive, Pasadena CA
41:   91109; California Institute of Technology, Pasadena, CA 91125;
42:   Warsaw University Observatory, Aleje Ujazdowskie 4, 00-478 Warszawa,
43:   Poland; Krzysztof.M.Gorski@jpl.nasa.gov}
44: 
45: \altaffiltext{4}{Institute of Theoretical Astrophysics, University of
46: Oslo, P.O.\ Box 1029 Blindern, N-0315 Oslo, Norway; 
47: f.k.hansen@astro.uio.no.} 
48: 
49: 
50: \begin{abstract}
51:   We explore methods of fitting templates to cosmic microwave background (CMB) data, and in
52:   particular demonstrate the application of the total convolver
53:   algorithm as a fast method of performing a search over all possible
54:   locations and orientations of the template relative to the sky.
55:   This analysis includes investigation of issues such as chance
56:   alignments and foreground residuals.  We apply these methods to
57:   compare Bianchi models of type VII$_{h}$ to \emph{WMAP} first year
58:   data and confirm the basic result of our 2005 paper.
59: \end{abstract}
60: 
61: \keywords{cosmology: cosmic microwave background -- cosmology: observations}
62: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
63: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
64: 
65: 
66: 
67: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
68: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
69: 
70: \section{Introduction}
71: 
72: The widely accepted model in cosmology, the so-called concordance
73: model, posits an isotropic and homogeneous universe with small
74: anisotropies generated by primordial fluctuations in the inflationary
75: field.  These anisotropies are present in the cosmic microwave
76: background (CMB), which should then be statistically isotropic and
77: Gaussian.  Many CMB studies therefore examine the CMB from a
78: statistical point of view with the intention of testing for violations
79: of these properties.  Alternative cosmological models have not,
80: however, been completely ruled out, and there are several anomalies in
81: the \emph{Wilkinson Microwave Anisotropy Probe} (\emph{WMAP}) data
82: that indicate that such models merit further investigation by
83: alternate means.
84: 
85: We investigate methods for testing any deterministic anisotropic
86: cosmological model.  The predicted anisotropy template can be compared
87: to the data using fitting techniques in both pixel and harmonic space
88: to search for correlations.  We present a description of these methods
89: and apply a fast and efficient algorithm for searching the full sky
90: for the best orientation of a template relative to the data.  We test
91: these methods with both full- and incomplete-sky data sets, and use
92: simulations to characterize the significance of the results.
93: 
94: Motivated by the morphology of several detected violations of
95: Gaussianity and/or isotropy in the \emph{WMAP} data \citep{de
96: Oliveira-Costa:2004, eriksen:2004a, hansen:2004a, vielva:2004}, we
97: test our methods using Bianchi type VII$_{h}$ models and the
98: \emph{WMAP} first-year data.  A preliminary analysis was published in
99: \citet{jaffe:2005}, in which we reported on a surprisingly significant
100: detection of a Bianchi model at the $99.7\%$ significance level
101: compared to simulations.  Here we present an improved search of the
102: model space, confirm the basic result, and discuss in detail issues
103: such as foreground contamination and chance alignments.
104: 
105: 
106: \section[]{Methods}\label{method}
107: 
108: 
109: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
110: \subsection{Template Fitting}\label{fitting}
111: 
112: 
113: 
114: Given any anisotropy pattern that contributes to the data as an
115: additional component of the observed microwave sky (whether
116: topological in origin, as in the case of Bianchi models, or
117: foreground), we perform a fit of the template to the \emph{WMAP} data
118: as has been done in the past by, \eg \citet{gorski:1996} and
119: \citet{banday:1996} for foreground analysis.  The best-fit amplitude
120: $\alpha$ for a template vector ${\mathbf t}$ compared to a data vector
121: ${\mathbf d}$ can be measured by minimizing
122: \begin{equation}
123: \chi^2 = ({\mathbf d}-\alpha {\mathbf t})^T  {\mathbf
124:   M}^{-1}_{\textrm{SN}} ({\mathbf d}-\alpha {\mathbf t}) = {\bf
125:   \tilde{d}}^T  {\mathbf M}^{-1}_{\textrm{SN}} {\bf
126:   \tilde{d}},
127: \label{eq:chi2}
128: \end{equation}
129: where ${\mathbf M}_{\textrm{SN}}$ is the covariance matrix including both signal
130: and noise for the template-corrected data vector ${\bf \tilde{d}}
131: \equiv {\mathbf d} - \alpha {\mathbf t}$.  Solving for $\alpha$ then becomes
132: \begin{equation}
133: \alpha = \frac{ {\mathbf t}^T{\mathbf M}^{-1}_{\textrm{SN}} {\mathbf d} }{ {\mathbf t}^T{\mathbf M}^{-1}_{\textrm{SN}} {\mathbf t} }.
134: \label{eq:cc_basic}
135: \end{equation}
136: 
137: 
138: To compare multiple template components to a given data sets, \eg
139: different foregrounds, the problem becomes a matrix equation.
140: In the case in which we have $N$ different foreground components, we
141: define
142: \begin{equation}
143: {\bf \tilde{d}} = {\mathbf d} - \sum_{k=1}^{N} \alpha_k {\mathbf t}_k 
144: \end{equation}
145: and 
146: \begin{equation}
147: {\mathbf M}_{\textrm{SN}} = \bigl< {\bf \tilde{d}}  {\bf \tilde{d}}^T \bigr> = {\mathbf M}_{\textrm{S}} + \mathbf{M}_{\textrm{N}}.
148: \end{equation}
149: In this case, minimizing  ${\bf \tilde{d}}^T  {\mathbf
150: M}^{-1}_{\textrm{SN}}  {\bf\tilde{d}}$ leads to the following set
151: of equations,
152: \begin{equation}
153:   \sum_{j=1}^N \mathbf{t}^T_k  \mathbf{M}^{-1}_{\textrm{SN}}  \mathbf{t}_j \alpha_j = \mathbf{t}^T_k  \mathbf{M}^{-1}_{\textrm{SN}}  \mathbf{d}. 
154: \end{equation}
155: This is the simple system of linear equations ${\mathbf Ax}={\mathbf b}$,
156: where 
157: %\begin{gather}
158: \begin{eqnarray*}
159: A_{kj}=\mathbf{t}^T_k  \mathbf{M}_{\textrm{SN}}^{-1}  \mathbf{t}_j, 
160: %\notag 
161: \\ 
162: b_k = \mathbf{t}^T_k  \mathbf{M}_{\textrm{SN}}^{-1}  \mathbf{d} 
163: %\notag, 
164: \\  
165: x_k =\alpha_k.
166: %\end{gather}
167: \end{eqnarray*}
168: When only one template is present, this reduces to equation
169: (\ref{eq:cc_basic}) above.
170: 
171: 
172: The errors $\delta\alpha^k_\nu$ are the square root of the diagonal of
173: ${\mathbf A}^{-1}$.  The matrix ${\mathbf A}$ gives information about
174: the cross-correlation between the templates themselves.  
175: 
176: Note that the above is equally valid in pixel space or harmonic space.
177: In the former, it is very easy to account for incomplete sky coverage
178: or to remove, for example the Galactic plane, by simply including in the data
179: vectors only the relevant pixels, and likewise by including only the
180: corresponding rows and columns of the covariance matrix.  The noise in
181: pixel space is usually well represented by a diagonal matrix
182: representing uncorrelated pixel noise.  But the signal covariance
183: matrix in pixel space is large and not sparse, which makes harmonic
184: space more convenient when this is possible.  In harmonic space and under the
185: assumption of Gaussianity, the signal covariance is diagonal, and with
186: the approximation of uncorrelated noise that is uniform over the sky,
187: the noise covariance can be made to be so as well.  The difficulty in
188: harmonic space is the sky coverage.  As discussed by
189: \citet{mortlock:2002}, the coupling matrix to cut out the Galactic
190: plane using a cut the size of $|b|>20\degr$ becomes numerically
191: singular for resolutions of $\ell_{\textrm{max}}>50$.  Cuts such as
192: the conservative Kp0 mask, defined by the \emph{WMAP} team, remove more
193: of the sky and, due to their structure, the coupling matrix is more
194: difficult to compute.
195: 
196: We define a method that applies harmonic space fitting to the
197: full-sky cases using highly processed maps discussed in \S\ \ref{maps}.  
198: %
199: This allows us to increase the computational efficiency using the
200: algorithm described in \S\ \ref{totalc}.  
201: %
202: Here, we use a uniform mean noise approximation that has a diagonal
203: covariance matrix.  We use pixel-space fitting for each band
204: separately in the cut-sky analysis in which the Galactic plane region is
205: masked out.  (At the \emph{WMAP} signal-to-noise ratio level, little would
206: be gained by simultaneously fitting the different bands, and the
207: memory and CPU requirements to invert the matrix would become
208: onerous.)  Again, we use a diagonal noise approximation that this time
209: takes into account the observation pattern but not the effects of
210: smoothing.  Comparisons of fits with fully correct noise to those
211: using these approximations show that the results do not vary
212: significantly (at most a few percent, or a small fraction of the error
213: bar).
214: %  Difference small compared to the error bar:
215: % /afs/mpa/planck/simdata/tjaffe/data/bianchi/wmap_fits/bands/logs/tpl_simulfit_smN_*.log
216: %/afs/mpa/planck/simdata/tjaffe/data/bianchi/wmap_fits/bands/fits_smN_*
217: All codes have been cross-checked with identical inputs to confirm
218: identical outputs.
219: 
220: 
221: Note that the cosmic monopole and dipole are not known, and although a
222: best-fit dipole is subtracted from the data in the map making process,
223: small residual monopole and dipole terms remain in the data.  For this
224: reason, we cannot include this component in the fit.  In harmonic
225: space, any monopole and dipole terms can simply be excluded or ignored
226: by setting, \eg $C_1=C_2=10^8\mu \textrm{K}^2$.  In pixel space, we
227: fit the monopole and dipole simultaneously as independent components.
228: See
229: \S
230: \ref{fg_bias} for discussion.
231: 
232: The above method determines the best-fit amplitude for a given
233: template at a fixed orientation relative to the sky.  For foreground
234: fitting, that is generally all that is required, but in the search for
235: an anisotropic cosmological component, there are two additional
236: complexities.  First, we have no {\em a priori} guess for where the
237: symmetry axis may be pointing and must thus search the entire sky.
238: Section
239: \ref{totalc} describes the algorithm we use to do this quickly and
240: efficiently.  Second, we may have an infinite number of possible
241: templates (e.g., parameterized as described in \S\ 
242: \ref{models}) among which we want to find the ``best'', so in
243: addition to determining the best-fit amplitude for each template, we
244: need a way to compare how well different templates fit the data and to
245: select the most interesting.
246: Section \ref{best} discusses how we address this.
247: 
248: 
249: 
250: \subsection{Total Convolver}\label{totalc}
251: 
252: The search for the best orientation of a template compared to the data
253: requires that we evaluate the statistic $\alpha$ described in the
254: previous section at every possible relative orientation of the
255: template and data.  Working in harmonic space allows us to use an
256: algorithm based on Fourier transforms to speed up this search
257: significantly.
258: 
259: In the case of full-sky analysis, the location or orientation of the
260: template does not affect the error, \ie  $\delta\alpha$ is invariant.
261: Then the maximum of $\alpha$ is found at the maximum of the numerator
262: in equation
263: (\ref{eq:cc_basic}) above, ${\mathbf t}^T {\mathbf
264: M}^{-1}_{\textrm{SN}} {\mathbf d}$.  Neglecting for the moment the
265: covariance matrix, the quantity to be maximized is simply the
266: convolution of the data with the template.  We seek the maximum over
267: all possible locations and orientations, and this can be found
268: efficiently using the total convolver algorithm described in
269: \citet{wandelt:2001}, which was originally developed for map making
270: using instrument beams.  
271: 
272: This algorithm decomposes the Euler angles into what amounts to a scan
273: pattern and then takes advantage of the form the convolution takes in
274: harmonic space to simplify the calculation.  The rotation operator
275: $D(\Phi_2,\Theta,\Phi_1)$ can be factored into
276: $D(\phi_{\textrm{E}},\theta_{\textrm{E}},0)D(\phi,\theta,\omega)$,
277: where a pre-defined scan pattern determines $\theta_{\textrm{E}}$ and
278: $\theta$, which in the case of full-sky coverage are both $\pi/2$, so
279: that the set of angles $(\phi_{\textrm{E}},\phi,\omega)$ covers the
280: full sky at all possible orientations (see \citet{wandelt:2001}
281: Figure 1).  (In {\em only} this context of total convolution on the
282: full sky, $\phi$ corresponds to the polar angle and
283: $\phi_{\textrm{E}}$ to the azimuthal angle.  Elsewhere in this paper,
284: these are represented by the more common $\theta$ and $\phi$.)
285: Defining $T(\phi_{\textrm{E}},\phi,\omega)\equiv {\mathbf
286:   t}^T{\mathbf d}$ as the quantity to be maximized, $b_{\ell m}$ as
287: the spherical harmonic coefficients of the template ${\mathbf t}$, and
288: $a_{\ell m}$ as that for the data ${\mathbf d}$, the convolution is then
289: %(Wandelt \& G\'orski  9 and 8)
290: \citep[eqs.~9~\&~8]{wandelt:2001}
291: %\begin{gather}
292: \begin{eqnarray}
293: \label{eq:tc_basic}
294: T_{mm'm''} = \sum_l a_{\ell m} d^\ell_{mm'}(\theta_{\textrm{E}}) d^\ell_{m'm''}(\theta) b^*_{\ell m''}  \\
295: T(\phi_{\textrm{E}},\phi,\omega) = \sum_{m,m',m''} T_{mm'm''}e^{im\phi_{\textrm{E}} + im'\phi + im''\omega},
296: %\end{gather}
297: \end{eqnarray}
298: where $d^\ell_{mm'}(\theta)$ is the real function such that
299: $D^\ell_{mm'}(\phi_2,\theta,\phi_1)=e^{-im\phi_2}d^\ell_{mm'}(\theta)e^{-im'\phi_1}$.
300: The problem has then become simply to calculate $T_{mm'm''}$ and
301: Fourier transform to $T(\phi_{\textrm{E}},\phi,\omega)$ to find
302: the maximum.
303: To take into account the signal and noise covariance, we simply use a
304: ``whitened'' data vector, ${\bf M}_{\textrm{SN}}^{-1}{\mathbf d}$.
305: 
306: The total convolver can find the best-fit position with an accuracy
307: limited only by the resolution of the inputs.  The positional accuracy
308: is $\pi/\ell_{\textrm{max}}$, which for our analysis is $2\degr .8$.  Note
309: that this is larger than the size of a pixel at the usual HEALPix
310: resolution of $N_{\textrm{side}}=\ell_{\textrm{max}}/2$.
311: 
312: It should also be noted that searching the full sky will {\em not}
313: return an unbiased estimate for the amplitude.  Simulations with a
314: known input value for a particular template at a known position will,
315: on average, have slightly higher amplitudes returned by the search.
316: If the correct template location is simply fit to an ensemble of
317: simulations with additional CMB and noise, the returned amplitudes
318: will have a Gaussian distribution with the correct mean and variance,
319: but the same is not true when one is searching for the best location
320: and orientation as well.  This is because the search is seeking the
321: maximum, and the resulting distribution is a form of extreme value
322: distribution\footnote{See, e.g., {\tt http://mathworld.wolfram.com/
323: ExtremeValueDistribution.html}\label{foot:evd}}, which introduces a
324: small positive bias in the results.  For realistic situations with CMB
325: and noise in addition to the component we are fitting, the total
326: convolver is likely to find a maximum amplitude a small distance away
327: from the true position.  How different the amplitudes and positions
328: are on average depends on the particular case in question, since it is
329: a function of how dominant the template is compared to the CMB and
330: noise, and how much the template structure changes over angular distance,
331: etc.  This is quantified for the particular case in question in \S
332: \ref{full_sky_accuracy} using simulations.
333: 
334: 
335: This method is approximately 2 orders of magnitude faster than
336: performing the fit in harmonic space over a grid of individual
337: rotations one at a time.  The disadvantage is the storage requirement
338: for the matrix $T$, which increases with the third power of the
339: resolution and becomes over 2GB for a HEALPix resolution of
340: $N_{\textrm{side}}=128$ or angular resolution of $42\arcmin$.
341: 
342: 
343: \subsection{Best-fit Model and Significance}\label{best}
344: 
345: As mentioned above, when it is not one unique template for which we
346: are testing but rather a set of possibilities, we need not only to
347: find the best fit of each to the data but also to find the best fit
348: among the possible models.  
349: %Bianchi models are continuously defined
350: %over a two-dimensional model space, so 
351: Depending on how the model space is parameterized, there can be an
352: infinite number of possibilities.  Previous studies seeking upper
353: limits on shear and vorticity \citep{kogut:1997,bunn:1996} used two
354: different statistics to determine the ``best''-fit model.
355: 
356: Given a model, Kogut \etal define the best-fit position and amplitude
357: in terms of $\Gamma=\alpha/\delta\alpha$.  They used
358: \emph{Cosmic Background Explorer} (\emph{COBE}) data, for which no full-sky analysis was possible.  In
359: the case of incomplete sky analysis, the amount of template structure
360: that is masked changes the significance of the fit.  A large amplitude
361: in which most of the structure is masked by the Galactic plane cut is not
362: as interesting as a lower amplitude fit in which the structure is
363: included.  By finding the maximum not of $\alpha$ but of $\Gamma$,
364: they attempt to find the most significant fit rather than simply the
365: maximum amplitude.
366: 
367: Bunn \etal (1996) use a different statistic to accomplish the same effective
368: selection.  They define $\eta_1 \equiv (\chi^2_0 -
369: \chi^2_1)/\chi^2_0$, where $\chi^2_1$ is as in equation \ref{eq:chi2},
370: and $\chi^2_0$ the corresponding statistic for the data by itself,
371: uncorrected for any anisotropic component.  The difference is then an
372: indication of how much better the data fit the (statistically
373: isotropic CMB) theory after correction for the anisotropic model.
374: 
375: Finding the maximum of $\Gamma$ is equivalent to finding the maximum
376: of $\eta_1$ (although Bunn et al. use a different statistical method).
377: So for a given model, either statistic can be used to find the
378: best-fit amplitude and position.  But it becomes more complicated to
379: compare one model to another in order to determine which model fits
380: the data better.
381: 
382: The problem with the simple approach, used by \citet{kogut:1997} as
383: well as in our preliminary analysis \citep{jaffe:2005}, of using
384: $\Gamma$ or $\eta_1$ to find the best model is that the distribution
385: of these values for chance alignments is not the same for all models.
386: Although they are generally quite similar, differences in the tails of
387: the distributions mean that a given value of $\Gamma$ has a slightly
388: different significance for different models.  This means that finding
389: the maximum of $\Gamma$ might have missed other models that are
390: significant but in which the tail of the distribution does not reach as
391: high in $\Gamma$.  In other words, the significance of the fit found
392: in our original result is not incorrect, but it is possible that such
393: an analysis fails to detect another significant model.
394: 
395: For this more complete analysis, we analyze a set of LILC simulations
396: \citep{eriksen:2004b,eriksen:2005}, using the above formalism to
397: characterize the distributions of $\alpha$ values for a given model.
398: %Figure
399: %\ref{fig:grid_sig3_contours} shows the $\Gamma$ value that represents
400: %the $3\sigma$ level chance alignment; \ie $99.7\%$ of simulations'
401: %best-fit chance alignments have a lower value of $\Gamma$.  This shows
402: %that different regions of the parameter space would be preferred by the
403: %simple $\Gamma$ test.  
404: In this analysis, for a given model, we compare the $\alpha$ value
405: (equivalently $\Gamma$, since $\delta\alpha$ does not change for a
406: given model on the full sky) for the \emph{WMAP} data against the
407: ensemble of simulations.  We can then quantify the significance of a
408: given model fit to the data based on the percentage of LILC
409: simulations in which the model fits with a lower amplitude.  This gives
410: clearer indication of which are the most interesting models than a
411: simple $\Gamma$ or $\chi^2$ statistic.
412: %  See /afs/mpa/planck/simdata/tjaffe/data/bianchi/wmap_fits_second/grids/left_big_grid_new/contours_chi{.pro,s.eps}
413: Comparison of the results using $\alpha$ or $\eta_1$ in this way show
414: there is little difference between the two in terms of how significant
415: a given fit to the data is against the simulations.  In the following
416: analysis, we use the numbers for $\alpha$ only.
417: 
418: 
419: 
420: \subsection{Visualization:  Cross-Correlation Signal Maps}\label{cc_maps}
421: 
422: It is helpful to be able to visualize what parts of the sky are
423: driving a particular fit.  To do this, we simply note that the
424: numerator of equation \ref{eq:cc_basic}, ${\mathbf t}^T{\mathbf
425: M}^{-1}_{\textrm{SN}} {\mathbf d}$, can be rewritten in pixel space
426: as $\sum_p [{\mathbf L}^{-1} {\mathbf t}]_p [{\mathbf L}^{-1} {\mathbf
427: d}]_p$, where ${\mathbf L}$ is the ``square root'' of the covariance
428: matrix ${\mathbf M}_\textrm{SN}$, or its lower triangular decomposition found
429: from, for example Cholesky decomposition.  A simple visualization is to turn this
430: into a map, where each pixel contains the product of $[{\mathbf
431: L}^{-1}_{\textrm{SN}} {\mathbf t}]$ and $[{\mathbf L}^{-1}_{\textrm{SN}}
432: {\mathbf d}]$ at that pixel.  This map shows exactly what regions on
433: the sky drive the fit at a given orientation.  This is particularly
434: important when certain regions of the sky are known to be
435: contaminated; these plots show whether or how much those regions
436: affect the fit.  Examples will be shown in \S\ \ref{two_models}.
437: 
438: 
439: 
440: 
441: \section{Data and simulations}
442: 
443: Here we describe the particular class of models we investigate and the
444: data sets used in the analysis.
445: 
446: 
447: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
448: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
449: \subsection{Bianchi Models}\label{models}
450: 
451: 
452: Bianchi type VII$_{h}$ refers to the class of spatially homogeneous
453: generalizations of Friedmann universes that include small vorticity
454: (universal rotation) and shear (differential expansion) components.
455: (Type VII$_0$ includes the flat Friedmann-Robertson-Walker model, and
456: VII$_{h}$ includes that with negative spatial curvature as special
457: cases.)
458: \citet{barrow:1985} solve the geodesic equations to derive the induced
459: CMB anisotropy by linearizing the anisotropic perturbations about the
460: Friedmann models.  Their solution does not include any dark energy
461: component, which is a significant shortcoming considering the
462: preponderance of evidence that now points to $\Omega_\Lambda \sim
463: 0.7$.  
464: But we examine them first as a test of our template-fitting 
465: methods and second because of the intriguing possibility
466: that they may explain several anomalies in the data.
467: 
468: Following the prescription in \citet{barrow:1985}, we construct a
469: template for the anisotropy induced by vorticity ($\omega$) and shear
470: ($\sigma$).  Bianchi type VII$_{h}$ models are parameterized by the
471: current total energy density $\Omega_0$ and a parameter $x$
472: \citep{collins:1973}, 
473: \begin{equation}
474:  x = \sqrt{ \frac{h}{1-\Omega_0} }, 
475: \end{equation}
476: where $h$ is related to the canonical structure constants and is that
477: to which the type VII$_{h}$ refers (see
478: \citealt{kogut:1997,bunn:1996,barrow:1985}). This parameter can be
479: understood as the ratio of the scale on which the basis vectors change
480: orientation to the Hubble radius (present 
481: values).  The resulting temperature anisotropy pattern is then
482: described by
483: \citep[ eq.~4.11]{barrow:1985}
484: %\begin{equation}
485: %\begin{split}
486: \begin{eqnarray}
487:   \frac{\Delta T}{T} = \left(\frac{\sigma}{H}\right)_0 \{ & [ B(\theta_R) + A(\theta_R)  ]\sin(\phi_R)  \nonumber \\
488: &  \pm  [ B(\theta_R) - A(\theta_R) ]\cos(\phi_R) \}, 
489: %\end{split} 
490: \label{eq:bianchi}
491: %\end{equation}
492: \end{eqnarray}
493: where $A$ and $B$ are also functions of $x$ and $\Omega_0$ and include
494: integrals over conformal time that trace the geodesic from the
495: surface of last scattering to observation.  The angles $\theta_R$ and
496: $\phi_R$ are not the observing angles; those are rather
497: $\theta_{\textrm{ob}}=\pi-\theta_R$ and $\phi_{\textrm{ob}}=\pi+\phi_R$.
498: The sign on the $\cos(\phi_R)$ term (or alternatively, the $\phi_R$ to
499: $\phi_\textrm{obs}$ transformation) determines the handedness. Then
500: $\sigma$ determines the amplitude of the fluctuation and $x$ the
501: pitch angle of the spiral.  The vorticity is then
502: \begin{equation}
503:  \left(\frac{\omega}{H}\right)_0 = \frac{  \sqrt{2(1+h)(1+9h)} }{
504: 6x^2\Omega_0 } \left(\frac{\sigma}{H}\right)_0.
505: \label{eq:vorticity}
506: \end{equation}
507: Note that the shear and vorticity values in our original paper 
508: \citep{jaffe:2005} contain an error in amplitude, although the basic conclusions are
509: not affected.
510: 
511: Equation \ref{eq:bianchi} can be rewritten as
512: \begin{equation}
513: \frac{\Delta T}{T} \propto \cos(\phi_R \pm \tilde{\phi}).
514: \end{equation}
515: 
516: In other words, for a given $\theta_R$, the temperature variation
517: follows a $\cos(\phi_R)$ dependence.  The phase shift $\tilde{\phi}$ is
518: ultimately a function of $\theta_R$ and the two physical parameters,
519: $x$ and $\Omega_0$.  The result is a spiral pattern with approximately
520: $N=2/\pi x$ twists.  The smaller the $x$, the smaller the
521: scale at which the basis vectors change their orientations and the
522: tighter the resulting spiral.  In the case of $\Omega_0<1$ models,
523: geodesic focusing leads to an asymmetry wherein the spiral structure
524: appears compressed in one direction along the rotation axis.
525: 
526: 
527: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
528: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
529: 
530: %  Figure 1:  Bianchi examples  
531: \begin{figure}
532: %{\epsfxsize=\linewidth \epsfbox{f1.eps}}
533: \plotone{f1.eps}
534: \figcaption{\small Examples of left-handed Bianchi anisotropy templates in
535: orthographic projection, all on a common color scale to show the
536: relative amplitudes.  These must be multiplied by a factor of
537: $\alpha=(\sigma/H)_0$, \ie the shear (realistically of order
538: $< 10^{-9}$), in order to give the amplitude of the observed
539: anisotropy in $\mu \textrm{K}$.  Note that these have been rotated by
540: $\beta=-90\degr$ to move the center of the structure from the
541: $-\hat{z}$ pole (as defined in equation
542: \ref{eq:bianchi}) to the Galactic Center. \label{fig:examples}}
543: \end{figure}
544: 
545: 
546: The template is calculated as $\frac{\Delta T}{(\sigma/H)_0}$, i.e. the
547: contents of the curly brackets in equation \ref{eq:bianchi} times the
548: average CMB temperature, so that the shear $(\sigma/H)_0$ is the
549: amplitude of the template to be found by fitting it against the CMB
550: anisotropies, $\Delta T$.  Examples are shown in Figure
551: \ref{fig:examples}, where the template is plotted without the
552: normalization by the shear.  In generating all of our Bianchi
553: templates, we have taken the redshift to the surface of last
554: scattering, or recombination, as $z_{\textrm{rec}}=1100$.  (Changing
555: to, for example, $z_{\textrm{rec}}=1000$ lowers the amplitude of the
556: anisotropy by $\sim15\%$, implying a corresponding increase in the
557: value of the shear $(\sigma/H)_0$ for a given $\Delta T$.)
558: 
559: We make the simple and pragmatic assumption that the anisotropy
560: induced by the geometry simply adds to the statistically isotropic and
561: Gaussian component.  
562: %There is certainly an inconsistency in assuming
563: %an inflationary scenario for generating the statistically isotropic
564: %component while testing for an anisotropic component due to a geometry
565: %that is incompatible with inflation.  However, inflation is only one
566: %example of how to generate such fluctuations.  Regardless of which
567: %mechanism generates such Gaussian fluctuations, \citet{bunn:1996}
568: %point out briefly why this additive simplification is likely to be
569: %reasonable, lacking a more self-consistent prescription.
570: 
571: We examine a grid of such models over $0.1\le\Omega_0\le1.0$ in
572: increments of $0.05$ and over $0.1\le x \le 10.0$ in increments of
573: $0.05$ in the interval $0.1\le x \le 1.0$ and then
574: logarithmically sampled up to $x=10.$  A finer grid was also examined
575: surrounding the best-fit model, $0.52\le x\le 0.68$ and $0.42\le
576: \Omega_0 \le 0.58$ in increments of $0.02$.  
577: For the largest values of $x$, the spiral has almost disappeared
578: (because the scale on which the basis vectors change orientations
579: becomes larger than the horizon size), and so models of higher $x$
580: are self-similar.  Smaller values of $x$ start to become physically
581: unrealistic.
582: \citet{collins:1973} point out that for $x\sim 0.05$, the
583: characteristic length scale over which basis vectors change
584: orientation becomes comparable to the size of large-scale structure,
585: which means that lower values are ruled out by observations of large
586: scale homogeneity.  Furthermore, as discussed in \S
587: \ref{loc_accuracy}, small values of $x$ require higher precision
588: analysis than is feasible.
589: 
590: 
591: 
592: 
593: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
594: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
595: \subsection{Data}\label{maps}
596: 
597: 
598: For this work, we are interested only in large scale structure.  In
599: all of the following, unless otherwise noted, we use maps in
600: HEALPix\footnote{{\tt http://healpix.jpl.nasa.gov/}} format
601: \citep{healpix} at a resolution of $N_{\textrm{side}}=32$ and smoothed
602: to an effective beam of FWHM $5\degr.5$, with harmonics up to
603: $\ell_{\textrm{max}}=64$.
604: 
605: The following full-sky maps are used in this analysis (where all
606: \emph{WMAP} data products are from the first-year data release):
607: \begin{itemize}
608: 
609: \item The full-sky \emph{WMAP} Internal Linear Combination (WILC) map released
610: by the \emph{WMAP} team (see \citealt{bennett:2003b}).  This map is
611: formed by taking linear combinations of the different bands such that
612: the foregrounds, each of which has a different spectral dependence
613: from the CMB, are removed leaving only CMB.  The different weights of
614: the linear combination are determined solely by the data, via minimum
615: variance, rather than by any prior assumptions about the foreground
616: behavior.
617: 
618: \item The Lagrange Internal Linear Combination (LILC) map of
619:   \citet{eriksen:2004b,eriksen:2005}.  The weights used to form the
620:   WILC map are slightly sub-optimal with respect to the
621:   minimum-variance criterion \citep{eriksen:2005}, and this is
622:   corrected in the LILC map, which uses Lagrange multipliers to
623:   compute the ILC weights.
624: 
625: \item The foreground-cleaned map of \citet{tegmark:2003}, hereafter
626: TOH.  This map is also generated by a linear combination of bands,
627: where in this case, the weights are determined in harmonic space.  
628: 
629: \end{itemize}
630: 
631: All of these maps contain residual foreground emission, some of which
632: is visible by eye along the Galactic plane and some of which extends
633: to high latitudes.  It should be noted that none of these maps is
634: intended for high-precision CMB analysis, but we nevertheless use them
635: in the following to locate the best-fit Bianchi template by full-sky
636: convolution.  Simulations show that these fits are affected by two
637: opposing biases (see \S\ \ref{totalc} and \S\ \ref{loc_accuracy}) that
638: are larger than the effects of the foreground residuals (see \S
639: \ref{full_sky_accuracy}),  thus justifying our use of these maps
640: despite their known disadvantages.  In general, we use the full-sky
641: maps initially to locate best-fit axis for each Bianchi model (see \S
642: \ref{full_sky_accuracy}), and then verify the amplitude using
643: partial-sky algorithms on the following additional data:
644: 
645: \begin{itemize}
646: 
647: \item \emph{WMAP} uncorrected maps for each of the five frequency
648: bands, co-added from each differencing assembly using noise weighting
649: (see \citealt{bennett:2003a}) and lso noise-weighted, coadded combinations
650: of bands Q+V, V+W, Q+V+W, Q-V, V-W, Q-W.
651: 
652: %\item \emph{WMAP} Q, V, and W band foreground cleaned maps per
653: %differencing assembly, coadded into one map per band;  see
654: %\citet{bennett:2003b}. 
655: 
656: \item Kp0 intensity mask, excluding $23.2\%$ of the pixels in which the
657: K-band intensity is high and also $0\degr.6$ around known point
658: sources, downgraded to $N_{\textrm{side}}=32$.
659: 
660: \end{itemize}
661: Finally, we use observations at other wavelengths as foreground
662: templates:
663: \begin{itemize}
664: 
665: \item the \citet{finkbeiner:1999} model for thermal dust emission
666: (hereafter FDS); 
667: 
668: \item the \citet{schlegel:1998} $100 \mu \textrm{m}$ intensity dust
669: template (hereafter SFD), which is used an alternative to the FDS
670: model (see discussion in \S\ \ref{best_result});
671: 
672: %  HKE had the Finkbeiner version.
673: %\item the \citet{dickinson:2003} H$\alpha$ template with a dust
674: %correction of $f_d=0.5$;
675: 
676: \item the \citet{finkbeiner:2003a} H$\alpha$ template, with dust
677:   correction $f_{\textrm{d}}=0.5$;   
678: 
679: \item the \citet{dickinson:2003} H$\alpha$ template with no dust
680: correction, which is used as an alternative to the Finkbeiner
681: template;
682: 
683: %  According to Tony, HKE had the de-striped version:
684: \item and the \citet{haslam:1982} 408MHz map of synchrotron emission
685: processed by \citet{davies:1996}.
686: 
687: \end{itemize}
688: 
689: These foreground components are fit simultaneously to each band over
690: the incomplete sky using the Kp0 mask, which reduces the effects of
691: foreground contamination on the fit amplitude (see \S\ \ref{fg_bias}).
692: Note that although we are simultaneously fitting the foreground
693: components, these templates are not accurate enough in the Galactic
694: plane region for full-sky fits to be reliable.
695: 
696: \subsection{Gibbs Samples}\label{gibbs}
697: 
698: In addition to the \emph{WMAP} data products, we also analyze a set of
699: Gibbs sampled maps that were generated by the method described by
700: \citet{jewell:2004},\citet{wandelt:2004}, and \citet{eriksen:2004c}.  Effectively, this method samples the space of
701: CMB signal maps that are consistent with the data, taking into account
702: both noise characteristics and limited sky coverage. Thus, each single
703: Gibbs sample represents a full-sky, noiseless CMB signal consistent
704: with the data assuming Gaussianity, and the distribution of such maps
705: describes the full CMB signal posterior distribution.
706: 
707: Such sampled maps can thus be analyzed very efficiently using the
708: total convolver method described above, since neither sky cut nor
709: non-uniform noise complicate the analysis.  These allow us to avoid
710: the problem of foreground residuals in the Galactic plane, since this
711: region of the Gibbs samples contains only CMB signal that is either
712: consistent with the structure outside the plane, in the case of large
713: enough scales, or entirely Gaussian random, in the case of smaller
714: scales.  The ensemble of fit results then reflects how well the
715: template fits the CMB signal posterior distribution.
716: In the following, we analyze ensembles of 1000 samples
717: corresponding to each of the three cosmologically important
718: \emph{WMAP} Q, V, and W bands.  
719: %The Kp0 mask that excludes point
720: %sources was imposed in all cases.
721: 
722: 
723: 
724: \subsection{Assumed Signal Covariance}\label{sig_covar}
725: 
726: Given that we are searching for evidence of anisotropy, the
727: description of the expected signal covariance is not trivial.  Bianchi
728: models in particular are not compatible with inflation theory and do
729: not make any prediction for fluctuations at the surface of last
730: scattering.  Clearly, a self-consistent theory is required to explain
731: the observed anisotropies in addition to the Bianchi component, and in
732: particular, that theory must be consistent with the acoustic peaks now
733: detected at smaller scales.  No such theory currently exists, but we
734: note that the Harrison-Zel'dovich power-law spectrum prediction
735: predates inflation theory.
736: %, and that that theory remains rather {\it ad hoc}.
737: Because it has been shown to match the data very well on small scales,
738: we use the inflationary prediction as a starting point.
739: 
740: The signal covariance expected after subtraction of any Bianchi
741: component is then assumed to be that of Gaussian, isotropic CMB
742: fluctuations fully characterized by the power spectrum.  We use the
743: best-fit \emph{WMAP} theoretical power-law spectrum to perform our
744: fit.  One could then refine the input spectrum based on the result
745: (\ie do a new parameter estimation using the corrected sky) and
746: iterate.  In the present analysis, however, we do not aim to improve
747: the power spectrum estimation.  Template fitting proves to be
748: insensitive to the assumed power spectrum.  (The fit result changes by
749: % From 
750: % /afs/mpa/planck/simdata/tjaffe/data/bianchi/wmap_fits/fits/ilc_unrot_fits.out
751: % /afs/mpa/planck/simdata/tjaffe/data/bianchi/wmap_fits/ilc_unrot_clsf18uK_fits.out
752: less than $3\%$ when using a flat, $Q=18\mu \textrm{K}$ power spectrum
753: instead.)
754: %,
755: %and by less than $4\%$ using a flat power spectrum with a completely
756: %implausible normalization a factor of ten different.
757: So for the
758: purposes of this analysis, the best-fit \emph{WMAP} theoretical power-law 
759: spectrum is sufficient.
760: 
761: 
762: 
763: \section{Performance, bias, and accuracy}\label{simulations}
764: 
765: In order to interpret the results of the analysis using real data, we
766: need first to quantify the effects described above.  The model
767: selection accuracy, the bias due to the maximization over rotations,
768: any bias due to foreground residuals, and the distribution of chance
769: alignments are all effects that we can quantify using simulations.
770: 
771: 
772: These are generated by the LILC simulation pipeline of
773: \citet{eriksen:2004b,eriksen:2005}.  The simulations start with a Gaussian CMB
774: signal generated from an assumed power spectrum and are then smoothed
775: to the beam width of each \emph{WMAP} differencing assembly.  Pixel noise is
776: added, uncorrelated and following the instrument properties and
777: observation pattern described in
778: \citet{bennett:2003a}.  Finally, the three foreground components above
779: are added to create simulated raw data for each of the 10 
780: differencing assemblies.  The LILC algorithm is then used to
781: reconstruct the corresponding processed, foreground-cleaned sky.
782: Although these are known to underestimate somewhat the amount of
783: residual emission along the Galactic plane, they provide a vital
784: indication of the morphology and approximate amount of such residuals
785: that may be present in the WILC or LILC maps.  
786: 
787: 
788: We apply the fitting methods outlined above to the ensemble of LILC
789: simulations, with and without an additional known anisotropic signal,
790: to characterize how well the methods perform.  In most of the analysis
791: below, a set of 1000 simulations were used in the full grid searches
792: and cut-sky pixel space fitting.  An expanded ensemble of 10,000 LILC
793: reconstructions was used to refine the significance measures for the
794: two best-fit models found as described in \S\ \ref{fit_results}.
795: 
796: 
797: 
798: \subsection{Model Selection Accuracy}\label{model_selection}
799: 
800: First, we add a known Bianchi component (the particular
801: template and amplitude found in our initial analysis
802: \citealt{jaffe:2005}) to a set of LILC simulations
803: and perform the full sky search over all rotations (using the total
804: convolver) and over the grid of models.  We find that the most
805: significant model returned is close ($\pm\sim0.1$ in $x$ and
806: $\Omega_0$) to the correct model in $\sim50\%$ of cases.  Among the
807: other $\sim50\%$, a qualitatively different model was found to be the
808: best-fit, but the correct model was still found to be over $99\%$
809: significant in most cases.  In other words, only in $\sim23\%$ of
810: realizations was the correct model not detected.
811: %  From 500 simulations here:
812: %/afs/ipp/bc-b/tjaffe/work/bianchi/ilc_sims_2nd_all_10k_better/rsurfs_pb_idl.dat
813: % /afs/ipp/bc-b/tjaffe/work/bianchi/ilc_sims_2nd_all_10k_better/contours_pb.pro
814: % /afs/ipp/bc-b/tjaffe/work/bianchi/ilc_sims_2nd_all_10k_better/idl_journal.pro.bak3
815: 
816: We must then see if we can distinguish the correct model from a false
817: detection by other means such as incomplete sky fits with simultaneous
818: foreground template fitting.  These give an idea how much the full-sky
819: fit is affected by residuals in the Galactic plane.  
820: %In cases where an
821: %incorrect model was selected as the best-fit, the cut-sky fit
822: %amplitude drops twice as much as in cases where the approximately
823: %correct model was the best-fit.  
824: %
825: % From 500 examples here:
826: % /afs/ipp/bc-b/tjaffe/work/bianchi/ilc_sims_2nd_kp0s_grid/set_0001/kp0_grid_best_?/grid_best_qvw_?_fits.out
827: %
828: % looked at here:
829: % /afs/ipp/bc-b/tjaffe/work/bianchi/ilc_sims_2nd_all_10k_better/fits_pb_compare.pro
830: % /afs/ipp/bc-b/tjaffe/work/bianchi/ilc_sims_2nd_all_10k_better/idl_journal.pro.bak2
831: %
832: Furthermore, models that appear far apart in the model space may in
833: fact be fitting to the same CMB structure.  We therefore select the
834: several most significant models to examine in more detail.  Then we
835: look at what structures are driving the fits and how they behave when
836: the Galactic plane is excluded and foreground templates simultaneously
837: fit.  These tools give an additional qualitative way to compare
838: different model fits.
839: 
840: 
841: \subsection{Full-Sky Fitting Accuracy}\label{full_sky_accuracy}
842: 
843: Next, we consider a known Bianchi component added to the input
844: noiseless, pure CMB realization (as opposed to the LILC
845: reconstruction) and see how well its position and amplitude are
846: recovered by the full-sky fit. For 1000 simulations, a Bianchi
847: component (at the same position and amplitude as our best-fit against
848: the real data) is added to the input CMB sky and then fit using the
849: total convolver method described above.  In $\sim80\%$ of
850: realizations, the returned fit is within $5\degr$ (approximately the
851: beam width) of the correct location.  
852: %
853: (In the orientation angle, it is less accurate due to the
854: self-similarity of the spiral structure under such rotations.  The
855: returned orientation is within $10\degr$ in $52\%$ of the
856: simulations.)
857: %
858: The amplitudes average $\sim7\%$ {\em higher} than the input value (as
859: noted in \S
860: \ref{totalc}), with an rms error of about $80\%$ the calculated error.
861: Neither of these facts is unexpected, since these values are the
862: selected maxima, and their distribution is not Gaussian.  The results
863: are quantitatively the same for the LILC reconstructed skies,
864: indicating that the foreground residuals do not introduce a
865: significant additional bias in the case in which a real Bianchi component
866: is being fitted.  Note that simulations in which the input Bianchi model
867: has an amplitude a factor of $\sim3$ higher show a much smaller {\it
868: relative} bias ($\sim1\%$), as one would expect.
869: % /afs/ipp/bc-b/tjaffe/work/bianchi/ilc_sims_2nd_all_10k_better/two_models_all_10k/compare_locations_*pro
870: 
871: 
872: \subsection{Cut-Sky Fitting Accuracy}\label{fg_bias}
873: 
874: The cut-sky fits are performed with the Bianchi model at the fixed
875: location found as the best-fit using the full-sky total convolver
876: method.  As described in \S\ \ref{totalc}, there is a bias
877: introduced by the selection of the maximum amplitude position.  This
878: bias will also be reflected in the cut-sky fits, although masking out
879: the Galactic plane should remove some of the bias due to residual
880: foreground emission.
881: 
882: For fits to the raw data outside the Kp0 mask, eight template
883: components are fit simultaneously to each band,  the three foreground
884: templates described in \S\ \ref{maps}, a monopole term, the three
885: spherical harmonics representing the real-valued dipole terms, and the
886: Bianchi template.
887: 
888: For simulations with no additional Bianchi component, the results show
889: amplitudes on average $6\%$ lower than those from the LILC fits.  This
890: is further indication that chance alignments are affected by residuals
891: in the plane, since the exclusion of that region tends to lower the
892: fit amplitude.
893: 
894: Simulations with an additional Bianchi component at a known position
895: and amplitude were run through the same pipeline, \ie first the
896: full-sky LILC reconstruction was used to find the best-fit location,
897: then that location used to fit the template to the cut sky in pixel
898: space.  As described above, the total convolver will return a position
899: that is very close to the true position but one where the fit amplitude
900: happens to be highest due to CMB and noise contributions.  These will
901: also affect the cut-sky fits, which also show a bias of $\sim3\%$.
902: This is lower than the bias in the full-sky fits, showing that a few
903: percent of the full-sky bias is due to residuals in the Galactic plane
904: region.  The relative drop in amplitude between the full- and cut-sky
905: fits for true detections is on average half the drop in the case of
906: chance alignment detections.
907: % From 500 examples here:
908: % /afs/ipp/bc-b/tjaffe/work/bianchi/ilc_sims_2nd_kp0s_grid/set_0001/kp0_grid_best_?/grid_best_qvw_?_fits.out
909: %
910: % looked at here:
911: % /afs/ipp/bc-b/tjaffe/work/bianchi/ilc_sims_2nd_all_10k_better/fits_pb_compare.pro
912: % /afs/ipp/bc-b/tjaffe/work/bianchi/ilc_sims_2nd_all_10k_better/idl_journal.pro.bak2
913: %
914: % /afs/ipp/bc-b/tjaffe/work/bianchi/ilc_sims_2nd_kp0s_grid/set_0001/fits/fits_pb_bfit_v_hketpls.out2
915: % /afs/ipp/bc-b/tjaffe/work/bianchi/ilc_sims_2nd_kp0s_grid/set_0001/fits/fits_pb_fixed_v_hketpls.out2
916: % /afs/ipp/bc-b/tjaffe/work/bianchi/ilc_sims_2nd_kp0s_grid/set_0001/fits/grid_right104_v_fits.out2
917: % /afs/ipp/bc-b/tjaffe/work/bianchi/ilc_sims_2nd_kp0s_grid/set_0001/idl_journal.pro.bak
918: 
919: 
920: 
921: % These were here but aren't related?
922: %/afs/ipp/bc-b/tjaffe/work/bianchi/ilc_sims_2nd_kp0s_grid/set_9001/grid_right104_qvw_fits.out
923: %/afs/ipp/bc-b/tjaffe/work/bianchi/ilc_sims_2nd_all_10k_better/two_models_all_10k/fits/lilcs_pbalms_noise3uK_right_fits.out
924: %/afs/ipp/bc-b/tjaffe/work/bianchi/ilc_sims_2nd_all_10k_better/two_models_all_10k/fits/inputs_noise3uK_right_fits.out
925: 
926: Figure \ref{fig:fit_histos} shows what these distributions look like
927: for the fit to 1000 simulations in the V band, both in the case where
928: a Bianchi component is added (\emph{red histogram}) and where it is
929: not (\emph{black histogram}).  Also plotted as vertical lines are the
930: mean and rms errors on the distributions, and the true
931: value and expected errors plotted in green.  The small bias in the value of the Bianchi
932: fit is seen in the distance between the vertical red and green lines.
933: 
934: 
935: %  Figure 2:  Histograms
936: % /afs/ipp/bc-b/tjaffe/work/bianchi/ilc_sims_2nd_kp0s_grid/set_0001/idl_scripts/read_three_fits_v_onlyb_fudge.pro
937: \begin{figure}
938: %{\epsfxsize=\linewidth \epsfbox{f2.eps}}
939: \plotone{f2.eps}
940: \figcaption{\small Distributions of fit results for the Bianchi component
941: for 1000 simulations without (\emph{black lines}) and with (\emph{red lines}) a Bianchi component
942: added.  Vertical solid lines show the means, and vertical dashed show the
943: actual errors.  The vertical green line shows the true value and
944: expected errors.  \label{fig:fit_histos}}
945: \end{figure}
946: 
947: 
948: 
949: 
950: Note that in all these cases, the bias in the fits affects the
951: absolute amplitude (\ie shear) estimate, but not the significance of
952: the fit, since the ensemble of simulations used to estimate the
953: significance is also affected by such a bias.  The expected bias in
954: the amplitude is also much smaller than the error bar.  Therefore this
955: does not affect our basic results, namely the particular best-fit
956: model, its location, its approximate amplitude, and its approximate
957: significance relative to chance alignments.
958: 
959: 
960: \subsection{Chance Alignments}\label{chance_alignments}
961: 
962: 
963: For a given sky realization, we find the best model as described in
964: \S\ \ref{best} and then simply compare the amplitude of that fit
965: against the ensemble of amplitudes for that model relative to Gaussian
966: simulations to estimate the significance.  Visual inspection of the
967: \emph{WMAP} sky maps shows no obvious Bianchi component, so any such
968: signal must remain at or below the level of the stochastic component.
969: Chance alignments may therefore either {\em cancel} a Bianchi-induced
970: signal or give a false positive .  The former effect was quantified in
971: \S\ \ref{model_selection} at $\sim23\%$, but the latter is more
972: difficult to quantify.
973: 
974: The family-wise error rate (FWER), the expected number of false
975: detections when testing $m$ hypotheses, is $\sim m p$ when $p$ is
976: the probability of one false detection.  If $3\sigma$ is the detection
977: threshold (implying $p\sim0.003$) and one tests 100 different
978: hypotheses (or models), the FWER is then $0.3$, meaning one gets a
979: false detection somewhere in the model space one-third of the time.
980: Over our grid of Bianchi parameters, the models are not independent
981: (since models close in $(x,\Omega_0)$ space will resemble each other
982: closely), so we cannot determine {\em a priori} what the true
983: frequency of false detections would be, but we can get this from the
984: ensemble of Gaussian simulations.
985: 
986: We perform the full-sky search using the total convolver over the grid
987: of Bianchi models and find the best-fit model for each realization.
988: We find that a false detection due to a chance alignment that has a
989: significance of $99.7\%$ occurs in $\sim17\%$ of the cases.  A better
990: comparison might be to use the $\chi^2$ representing the goodness of
991: the fit.  We then compare the statistic $\eta_1 \equiv (\chi^2_0
992: -\chi^2_1)/\chi^2_0$ (defined above in \S \ref{best}) , namely the
993: relative improvement in the $\chi^2$ when the Bianchi model is
994: subtracted.  We find that by this measure $\sim10\%$ of the best
995: chance alignments fit their respective realizations as well as our
996: best-fit model does the \emph{WMAP} data (see \S\ \ref{fit_results}).
997: Note, however, that these statistics are dependent on the assumed
998: amount of large-scale power.
999: The above numbers simply imply that a detection of a Bianchi model
1000: with an amplitude higher than in $99.7\%$ of simulations is more than
1001: 4 times as likely to be real as it is to be a chance alignment, in
1002: the absence of all other information.
1003: % about the fit morphology.
1004: 
1005: 
1006: %%  For the full-sky fits:
1007: %%
1008: %/afs/ipp/bc-b/tjaffe/work/bianchi/ilc_sims_2nd_all_10k_better/chis_best.pro
1009: %%  
1010: %%  And it's no improvement for cut-sky fits, unfortunately.  
1011: 
1012: 
1013: 
1014: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1015: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1016: \section{Application to the first-year \emph{WMAP} data}\label{fit_results}
1017: 
1018: Armed with the information gained from the analyses of simulations, we
1019: can now examine the fits to the real data.
1020: 
1021: \subsection{Fits Over Model Space Grid}\label{results_full}
1022: 
1023: 
1024: Using the total convolver to find the best orientation, we fit the
1025: grid of Bianchi models to each of the WILC, LILC, and TOH full-sky
1026: processed maps.  Figure \ref{fig:contours} shows filled contours over
1027: this grid for the LILC.  (The results for the WILC and TOH look very
1028: similar.)  For each point on the grid corresponding to a model of the
1029: given $(x,\Omega_0)$, the template is fit to the LILC map, and the
1030: color indicates the significance estimate of the resulting amplitude,
1031: \ie the fractional number of Gaussian LILC simulations (out of 1000)
1032: with lower amplitude.  As discussed in \S\ \ref{best}, we use a finer
1033: grid and better method for determining the best-fit model and thereby
1034: select a slightly different model than the analysis in
1035: \citet{jaffe:2005}.  But it is apparent from the right panels of
1036: Figure \ref{fig:contours} that the significance as a function of the
1037: Bianchi parameters $x$ and $\Omega_0$ is flat in the region 
1038: $\pm 0.1$ in both $x$ and $\Omega_0$ about the maximum.  
1039: % at $(x,\Omega_0)=(0.62\pm 0.1,0.5\pm 0.1 )$.
1040: 
1041: 
1042: %  Figure 3:  contours
1043: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1044: %	See 
1045: % /afs/ipp/bc-b/tjaffe/work/bianchi/ilc_sims_new/full_sky/read_fits2.pro
1046: % /afs/ipp/bc-b/tjaffe/work/bianchi/ilc_sims_new/full_sky/create_contours2.pro
1047: % /afs/ipp/bc-b/tjaffe/work/bianchi/ilc_sims_new/full_sky/plot_contours2.pro
1048: \begin{figure}
1049: %{\epsfxsize=\linewidth \epsfbox{f3.eps}}
1050: \plotone{f3.eps}
1051: \figcaption{\small Significance as percentage of LILC simulations whose
1052: best-fit chance alignment amplitude is lower.  Left panels show 
1053: left-handed models, and right panels show right-handed models.  Over plotted contours are at $99.3\%$,
1054: $99.5\%$, $99.7\%$, and $99.9\%$.  Two color scales are used to show the
1055: global structure ([empha{top panels}) as well as that near the peaks (\emph{bottom panels}).\label{fig:contours}}
1056: \end{figure}
1057: 
1058: 
1059: 
1060: We find that the most significant fit is found with a right-handed
1061: Bianchi template of $x=0.62$ and $\Omega_0=0.5$ when that template is
1062: rotated to a position and orientation given by Euler angles (following
1063: the total convolver's ``$zyz$'' convention about fixed axes)
1064: $(\Phi_2,\Theta,\Phi_1)=(42\degr,28\degr,-51\degr)$.
1065: As defined in \S\ \ref{models}, the spiral structure of the
1066: unrotated model is centered on the south pole (or $-\hat{z}$ axis), so
1067: this rotation places the center of that structure at Galactic
1068: longitude and latitude of $(l,b)=(222\degr,-62\degr)$ and changes it's
1069: orientation about that location by $\Phi_1=-51\degr$.  This model fits
1070: at an amplitude of $\left(\frac{\sigma}{H}\right)_0=4.29\times
1071: 10^{-10}$, which is higher than $99.7\%$ of the 10\,000 simulations.  
1072: This model and the best-fit from previous work \citep{jaffe:2005} at
1073: $x=0.55$ are almost identical.
1074: 
1075: 
1076: All models near this best-fit $(x,\Omega_0)$ return the same location
1077: for the center of the spiral within $3\degr$ but vary the orientation
1078: (Euler angle $\Phi_1$) up to $36\degr$. The broad spiral in all of
1079: these models is very self-similar under these rotations, so the change
1080: is driven largely by the precise locations of the paired hot and cold
1081: spots.
1082: 
1083: Looking at Figure \ref{fig:contours}, one can see that more than one
1084: model appears ``significant'' in the sense of fitting with an
1085: amplitude above $99\%$ of the amplitudes found fitting that
1086: same model to Gaussian simulations.  As discussed above in \S
1087: \ref{model_selection}, this is not surprising, and we must examine each
1088: of these models in more detail.  
1089: 
1090: 
1091: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1092: %  Table 1:
1093: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1094: %
1095: \begin{deluxetable}{lcccc}
1096: %
1097: \tablecaption{Fitted template amplitudes \label{tab:fits_summary} }
1098: 
1099: %
1100: \tablewidth{0pt}
1101: %
1102: \tablecolumns{4}
1103: %
1104: \tablehead{  & $(\sigma/H)_0$ & $(\omega/H)_0$ &  $P\left(|\alpha_{\textrm{sim}}| < |\alpha_{\textrm{obs}}|\right)$   \\
1105: Map  & ${\scriptstyle (\times10^{-10})}$ & ${\scriptstyle (\times10^{-10})}$ & \%  }
1106: %
1107: \startdata
1108: %  Full sky fits from:
1109: % /afs/ipp/bc-b/tjaffe/work/bianchi/wmap_fits_individual2/write_table.pro
1110: %
1111: % Cut sky fits from:
1112: % /afs/ipp/bc-b/tjaffe/work/bianchi/wmap_fits_individual2/kp0s/read6.pl
1113: %
1114: % Gibbs from:
1115: % /afs/ipp/bc-b/tjaffe/work/bianchi/wmap_fits_individual2/gibbs/read2.pro
1116: \cutinhead{Right-handed $(x,\Omega_0)=(0.62,0.5)$} \\
1117: %
1118: % Full sky:
1119: WILC & $ 4.33\pm 0.82$ & $ \phm{-}9.58$ & $ 99.8$ \\
1120: LILC & $ 4.29\pm 0.82$ & $ \phm{-}9.49$ & $ 99.7$ \\
1121: TOH & $ 4.03\pm 0.82$ & $ \phm{-}8.92$ & $ 98.6$ \\
1122: %
1123: % Cut sky:
1124: K\tablenotemark{a} & $ 2.59( 4.13)\pm 0.83$ & $\phm{-} 5.72$ & $ 16.7( 99.1)$ \\
1125: Ka\tablenotemark{a} & $ 3.50( 4.09)\pm 0.83$ & $\phm{-} 7.74$ & $ 86.9( 99.0)$ \\
1126: Q\tablenotemark{a} & $ 3.76( 4.11)\pm 0.83$ & $\phm{-} 8.31$ & $ 95.6( 99.1)$ \\
1127: V\tablenotemark{a} & $ 3.99( 4.19)\pm 0.83$ & $\phm{-} 8.82$ & $ 98.1( 99.5)$ \\
1128: W\tablenotemark{a} & $ 4.08( 4.35)\pm 0.82$ & $\phm{-} 9.03$ & $ 99.1( 99.8)$ \\
1129: QVW\tablenotemark{a} & $ 3.84( 4.15)\pm 0.83$ & $\phm{-} 8.49$ & $ 96.8( 99.2)$ \\
1130: VW\tablenotemark{a} & $ 3.99( 4.22)\pm 0.83$ & $\phm{-} 8.84$ & $ 98.2( 99.6)$ \\
1131: Q-V\tablenotemark{a} & $ 0.06( 0.11)\pm 0.02$ & $\phm{-} 0.13$ & $ 99.0(100.0)$ \\
1132: V-W\tablenotemark{a} & $-0.05(-0.08)\pm 0.02$ & $\phm{-} 0.11$ & $ 93.8( 99.0)$ \\
1133: Q-W\tablenotemark{a} & $ 0.01( 0.04)\pm 0.02$ & $\phm{-} 0.02$ & $ 25.0( 83.2)$ \\
1134: 
1135: %
1136: % Gibbs samples:
1137: Q$^{b}$ & $ 4.09\pm 0.10^c$ & $\phm{-}9.04$ & - \\
1138: V$^{b}$ & $ 4.11\pm 0.10^c$ & $\phm{-}9.09$ & - \\
1139: W$^{b}$ & $ 4.12\pm 0.11^c$ & $\phm{-}9.12$ & - \\
1140: 
1141: 
1142: \cutinhead{Left-handed $(x,\Omega_0)=(0.62,0.15)$} \\
1143: %
1144: % Full sky:
1145: WILC & $ 2.39\pm 0.47$ & $ \phm{-}22.31$ & $ 97.8$ \\
1146: LILC & $ 2.49\pm 0.47$ & $ \phm{-}23.29$ & $ 99.4$ \\
1147: TOH & $ 2.45\pm 0.47$ & $ \phm{-}22.94$ & $ 99.0$ \\
1148: %
1149: %  Cut sky:
1150: K\tablenotemark{a} & $ 2.33( 3.31)\pm 0.50$ & $\phm{-}21.76$ & $ 96.3( 99.9)$ \\
1151: Ka\tablenotemark{a} & $ 2.24( 2.63)\pm 0.50$ & $\phm{-}20.93$ & $ 94.8( 99.3)$ \\
1152: Q\tablenotemark{a} & $ 2.29( 2.50)\pm 0.50$ & $\phm{-}21.42$ & $ 96.0( 98.9)$ \\
1153: V\tablenotemark{a} & $ 2.33( 2.44)\pm 0.50$ & $\phm{-}21.81$ & $ 96.7( 98.6)$ \\
1154: W\tablenotemark{a} & $ 2.32( 2.46)\pm 0.49$ & $\phm{-}21.69$ & $ 96.3( 98.4)$ \\
1155: QVW\tablenotemark{a} & $ 2.30( 2.48)\pm 0.50$ & $\phm{-}21.46$ & $ 96.1( 98.8)$ \\
1156: VW\tablenotemark{a} & $ 2.34( 2.44)\pm 0.50$ & $\phm{-}21.85$ & $ 96.7( 98.6)$ \\
1157: Q-V\tablenotemark{a} & $ 0.03( 0.02)\pm 0.02$ & $\phm{-} 0.27$ & $ 78.6( 63.1)$ \\
1158: V-W\tablenotemark{a} & $-0.06(-0.10)\pm 0.02$ & $\phm{-}0.55$ & $ 96.4( 99.9)$ \\
1159: Q-W\tablenotemark{a} & $-0.03(-0.08)\pm 0.02$ & $\phm{-}0.29$ & $ 73.1( 99.3)$ \\
1160: 
1161: 
1162: 
1163: %
1164: % Gibbs samples:
1165: Q$^{b}$ & $ 2.10\pm 0.11^c$ & $\phm{-}19.67$ & - \\
1166: V$^{b}$ & $ 2.08\pm 0.11^c$ & $\phm{-}19.46$ & - \\
1167: W$^{b}$ & $ 2.09\pm 0.09^c$ & $\phm{-}19.53$ & - \\
1168: 
1169: \enddata
1170: %
1171: \tablecomments{ Amplitudes of the best-fit model derived from various
1172:   combinations of data and various methods as described in the
1173:   text. The full sky was used in the analysis of the WILC, LILC, TOH,
1174:   and Gibbs samples, while the Kp0 mask was imposed for the remaining
1175:   maps.}
1176: 
1177: \tablenotetext{a}{Simultaneous fits with foreground components.  In
1178: parentheses are the values using the SFD dust template instead of the
1179: FDS, and the \cite{dickinson:2003} H$\alpha$ instead of
1180: \citet{finkbeiner:2003a}.}
1181: \tablenotetext{b}{Average over 1000 Gibbs samples.}
1182: \tablenotetext{c}{Errors are rms variation over Gibbs samples.}
1183: %\tablenotetext{c}{\emph{WMAP} template corrected maps.}
1184: %
1185: \end{deluxetable}
1186: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1187: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1188: 
1189: 
1190: 
1191: The full resolution ILC map is shown along with the best-fit Bianchi
1192: model on the same scale and the corrected ILC map in Figure
1193: \ref{fig:ilc_b_bilc}.
1194: A summary of all fit results is shown in Table \ref{tab:fits_summary}.
1195: The expected bias in these results is discussed in \S
1196: \ref{fg_bias}.  The following sections describe the two most
1197: interesting models in more detail.  
1198: 
1199: 
1200: 
1201: % Figure 4:  three maps
1202: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1203: %	See
1204: %  /afs/mpa/planck/simdata/tjaffe/data/bianchi/wmap_fits/ilc_bmap_ns256_idl.dat
1205: %  /afs/mpa/planck/simdata/tjaffe/data/bianchi/wmap_fits/idl_journal.pro.cbar5
1206: \begin{figure}\begin{centering}
1207: %\mbox{\epsfig{figure=ilc_b_bilc_cbar_new.eps,width=\linewidth,clip=}}
1208: %
1209: % /afs/mpa/planck/simdata/tjaffe/data/bianchi/wmap_fits/three_plots.pro
1210: % Then autocrop and rotate them in Gimp.
1211: %\mbox{\epsfig{figure=f4a.eps,width=0.7\linewidth}}
1212: %\mbox{\epsfig{figure=f4b.eps,width=0.7\linewidth}}
1213: %\mbox{\epsfig{figure=f4c.eps,width=0.7\linewidth}}
1214: \epsscale {.70}
1215: \plotone{f4a.eps}
1216: \plotone{f4b.eps}
1217: \plotone{f4c.eps}
1218: \figcaption{\small  \emph{Top}:  \emph{WMAP} Internal Linear Combination map. \emph{Middle}:  Best-fit Bianchi VII$_{h}$ template (enhanced by a factor of
1219:   4 to bring out structure). {\it Bottom}:  Difference between WILC
1220:   and best-fit Bianchi template; the ``Bianchi-corrected'' ILC map.
1221:   Over-plotted on each as a dotted line is the equator in the reference
1222:   frame that maximizes the power asymmetry as described in \S\ 
1223: \ref{sec:asymmetry}. \label{fig:ilc_b_bilc}}
1224: \end{centering}
1225: \end{figure}
1226: 
1227: 
1228: \subsection{Two Best Fits}\label{two_models}
1229: 
1230: 
1231: 
1232: 
1233: 
1234: %\subsubsection{Left-handed model $(x,\Omega_0)=(2.,0.7)$}
1235: %
1236: %
1237: %Though not as significant as either of the two other best-fit models,
1238: %this model fits at almost $99\%$ significance.  The cut-sky amplitude
1239: %increases slightly to $3.4\times10^{-10}$ (for Q-V-W) compared to the
1240: %%/afs/ipp/bc-b/tjaffe/work/bianchi/ilc_sims_2nd_all_10k_better/grid_left_199/lilcs_grid_left_noise3uKfits.out
1241: %full-sky value of $3.2\pm0.7\times10^{-10}$.  From Figure
1242: %\ref{fig:three_fit_maps}, it is apparent that this model fits much the same
1243: %structure that the spiral of the best-fit right-handed model fits, but
1244: %this model does not account for the cold spot, and in fact makes it
1245: %worse.  The region of model space of high $x$ and $\Omega_0$ (red
1246: %areas on the upper right of both the top left and top right plots in
1247: %Figure \ref{fig:contours}) both correspond to similar structures,
1248: %since the higher the $x$, the less vorticity and the closer these are
1249: %to pure focused quadrupoles.
1250: 
1251: 
1252: 
1253: \subsubsection{Left-handed Model $(x,\Omega_0)=(0.62,0.15)$}
1254: 
1255: The most significant left-handed model, at $99.4\%$, is at
1256: $(x,\Omega_0)=(0.62,0.15)$.  This model was not found in our earlier
1257: work \citep{jaffe:2005}, because it is only in a fairly small region
1258: of the model space that this fits with any significance, and our
1259: previous, coarser grid effectively straddled the peak in $\Omega_0$.
1260: The best-fit location for this model puts the center of the structure
1261: at $(l,b)=(320,-20)$, which is closer to the Galactic Center region
1262: than the best-fit right-handed model, raising the question of how
1263: much it
1264: %, but it is not clear how much the fit
1265: is driven by foreground residuals.
1266: 
1267: Cut-sky fits give fit amplitudes for this component that are 8\%
1268: lower and significances of  $\sim96\%$ in most cases.  Furthermore,
1269: the Galactic center region tends to draw the template in simulations;
1270: the best-fit location among the simulated LILC maps for this model is
1271: twice as likely to be found in the area around $(0\degr,-20\degr)$ as
1272: should be expected from a uniform distribution.  The only thing that
1273: all of the LILC simulations have in common is foregrounds, so this is
1274: an indication that there is some residual there that is a weak
1275: attractor.  One possibility is the ``free-free haze'' described by
1276: Finkbeiner (2004, see also Patanchon et al.\ 2005), although 
1277: this haze does not match up well with the template structure, the two
1278: show little cross-correlation, and inclusion of Finkbeiner's haze
1279: template in the simultaneous fitting does not alter the fit amplitude
1280: of the Bianchi model.
1281: %/afs/mpa/planck/simdata/tjaffe/data/bianchi/wmap_fits_second/grids/right_big_grid_new/idl_journal.pro.bak
1282: 
1283: In Figure \ref{fig:three_fit_maps}, it looks like the fit should be
1284: largely driven by the cold region below the Galactic center.  The
1285: cross-correlation maps described in \S\ \ref{cc_maps} do show
1286: correlation there but also indicate that the fit is largely driven by
1287: a very strong signal in the Galactic plane.  Figure
1288: \ref{fig:three_fit_maps} shows these maps for both this model and the
1289: best-fit right handed model.  Where the right handed model shows
1290: relatively uniform correlation over the hemisphere about the best-fit
1291: axis, this model shows a rather concentrated region including a very
1292: strong driver on the Galactic plane.
1293: 
1294: The Gibbs samples throw further doubt on this model.  Among the 1000
1295: Gibbs samples in each of Q, V, and W bands, this model fits at the
1296: same approximate location as for the LILC map less than half of the
1297: time.  Where the location was the same, the amplitude of the best-fit
1298: is significantly lower for the ensemble of Gibbs-sampled maps, which
1299: drop over $15\%$ in amplitude to a mean of $2.1\times10^{-10}$,
1300: indicating that some of the structure in the data that drives the fits
1301: is not consistent with the posterior CMB distribution as determined by
1302: the Gibbs sampling technique.  Furthermore, this model is almost as
1303: likely to fit near the location of the best-fit right-handed model
1304: instead of near the Galactic center.  This is largely driven by the
1305: cold spot.
1306: 
1307: In summary, this model is quantitatively less significant than the
1308: best-fit right-handed model based on the cut sky and Gibbs sample fit
1309: values.  Furthermore, the morphology indicates that foreground
1310: residuals drive the full-sky fit.
1311: 
1312: 
1313: % Figure 5:  best 2
1314: % /afs/mpa/planck/simdata/tjaffe/data/bianchi/wmap_fits_second/bands/simulations/plot_maps4.pro
1315: \begin{figure}\begin{centering}
1316: %\mbox{\epsfig{figure=wmap_qvw_crop.eps,width=75mm,clip=}}
1317: %\mbox{\epsfig{figure=f5.eps,width=\linewidth,clip=}}
1318: \epsscale{1.0}
1319: \plotone{f5.eps}
1320: \figcaption{\small   Two significant models.  The left panels show the
1321: best-fit left handed model with $(x,\Omega_0)=(0.62,0.15)$, while the
1322: right panels show the best-fit overall model, right-handed with
1323: $(x,\Omega_0)=(0.62,0.5)$.  The top panels show the template
1324: amplified by a factor of three to bring out the structure.  The middle
1325: panels show the corresponding cross correlation map (see \S
1326: \ref{cc_maps}) scaled from $-1\%$ to $2\%$.  The bottom panels show the 
1327: ``corrected'' \emph{WMAP} Q+V+W map scaled from $-150$ to $150\mu K$.
1328: The grey region is the excluded region of the Kp0 mask.\label{fig:three_fit_maps}}
1329: \end{centering}
1330: \end{figure}
1331: 
1332: 
1333: 
1334: \subsubsection{Right-handed Model $(x,\Omega_0)=(0.62,0.5)$}\label{best_result}
1335: 
1336: Figure \ref{fig:contours} shows that the best-fit model is this
1337: right-handed model.
1338: 
1339: The amplitude of the best-fit Bianchi component varies somewhat across
1340: the different frequencies, in all cases lower than the full-sky
1341: amplitude fit with the LILC.  As discussed in \S
1342: \ref{fg_bias}, this is likely due to small foreground residuals, but does not mean that the detection is a false positive;  the 
1343: same effect occurs in simulations that include a Bianchi component.
1344: The amplitude in the W band, in which the least foreground residuals are
1345: expected, is still higher than $\sim99\%$ of simulations.  The K and
1346: Ka band fits are significantly lower when the FDS dust and Finkbeiner
1347: H$\alpha$ templates are used, but are consistent with the other bands
1348: when the SFD dust and Dickinson H$\alpha$ templates are used instead.
1349: It is known that foreground subtraction is a problem even at high
1350: latitudes in the K and Ka bands, and these residuals are clearly
1351: affecting the low frequency fits.  Looking at the residuals of the two
1352: fits shows that the difference may be driven by a small region around
1353: $(l,b)=(300\degr,-15\degr)$ where the dust templates differ strongly.
1354: The higher frequency fits, however, are more consistent.  The
1355: difference maps, \eg Q-V, should contain no CMB component but only
1356: foreground residuals and noise.  The fact that the Bianchi component
1357: amplitude found from these maps is less than $2\%$ of the co-added map
1358: amplitude is an indication that such residuals are not contributing
1359: significantly to the fit.
1360: 
1361: The results of fitting the Gibbs-sampled maps show that for this
1362: model, the amplitude is quite stable over the ensemble of Gibbs
1363: samples, with, \eg a mean of $(4.12\pm0.1)\times10^{-10}$ in the W band
1364: compared to $4.08\times10^{-10}$ for the cut-sky fit to the raw data.
1365: Since the Gibbs samples represent the posterior CMB distribution,
1366: taking into account foregrounds and iterating over the power spectrum,
1367: these results are a strong indication that the fit is due primarily to
1368: CMB signal.  
1369: 
1370: Figure \ref{fig:three_fit_maps} shows the cross-correlation map as
1371: described in \S\ \ref{cc_maps}, which give a visual indication of what
1372: regions drive the fit.  Unlike the left-handed model (\emph{left}), which
1373: shows one concentrated region in the Galactic plane to be driving the
1374: fit, this model correlates over more than half the sky at moderate
1375: levels.  One can see that the cold spot does partly drive the fit, but
1376: no particular region can be said to dominate.  
1377: % /afs/ipp/bc-b/tjaffe/work/bianchi/wmap_fits_individual2/masked_coldspot
1378: Fits to the combined QVW and VW data where the cold spot is excluded
1379: (in a $10\degr$ radius around $(l,b)=(209,-57)$) have comparable
1380: amplitudes to fits where the region is included (only $6\%$ lower) .
1381: Further Gibbs samples were also computed while masking this region.
1382: Full sky searches using these samples show that fewer than $20\%$
1383: return positions more than $10\degr$ from the original location, and
1384: amplitudes that are on average $15\%$ lower (which is within the
1385: calculated error bar).  These results confirm that the cold spot does
1386: affect but does not exclusively drive the fit amplitude.
1387: 
1388: 
1389: \subsection{Location and Orientation Accuracy}\label{loc_accuracy}
1390: 
1391: As mentioned above, where a Bianchi component was added to simulations
1392: at a known location, the full-sky search with the total convolver
1393: returned the correct position within $5\degr$ in $\sim80\%$ of
1394: realizations.  The uncertainty in the location is due to the CMB
1395: fluctuations, which are quite comparable to the Bianchi component at
1396: the amplitude detected.  %of the best-fit to the real data.
1397: 
1398: To determine how the amplitude changes with the position and
1399: orientation of the template compared to the data, we take the best-fit
1400: Bianchi model and fit it to the LILC on a grid of fixed positions
1401: within $20\degr$ of the best-fit position.  Results are shown in
1402: Figures
1403: \ref{fig:sig_by_dist} and
1404: \ref{fig:sig_by_alpha}.  The orientation is not very sensitive in this
1405: model, whose spiral structure is self similar under rotations about
1406: its symmetry axis; only the precise positions of the hot and cold
1407: spots affect the variation with orientation angle.  The amplitude
1408: drops by $1\%$ when the orientation is $4\degr$ off.  The location of
1409: the symmetry axis is a bit more sensitive, where the amplitude drops
1410: by $3\%$ at $2\degr$.  The fact that the total convolver at this
1411: resolution uses steps of $2\degr.8$ means that its best-fit
1412: amplitude can be several percent off of the actual maximum.  All the
1413: fits to the simulations as well as the data are subject to this same
1414: uncertainty.  If we assume the worst, that the LILC amplitude was
1415: found at its true maximum (\ie the true axis of symmetry happened to
1416: lie exactly on the center of one of the total convolver's bins) and
1417: the simulations are all at $1\degr .4$ away from their true maxima (\ie the
1418: axis exactly between bins) and have true values correspondingly
1419: higher, the comparative significance could then be over-estimated by
1420: only $0.5\%$.  The likely effect is of course much smaller.
1421: %  See: /afs/mpa/planck/simdata/tjaffe/data/bianchi/wmap_fits_second/grids/left_big_grid_new/idl_journal.pro.bak3
1422:  
1423: % Figure 6:  amplitude with angular distance
1424: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1425: %	See
1426: %  /afs/mpa/planck/simdata/tjaffe/data//bianchi/ilc_sims/smoothed_5.5deg_10k/lilc_angle_grid/grid_angles.pro
1427: %  /afs/mpa/planck/simdata/tjaffe/data//maps/bianchi/ilc_fit/create_grid.pl
1428: %
1429: \begin{figure}
1430: %{\epsfxsize=\linewidth \epsfbox{f6.eps}}
1431: \plotone{f6.eps}
1432: \figcaption{\small Average fit amplitude as the location
1433: of the template varies from the best-fit position.  For these fits,
1434: the orientation of the template is unchanged.\label{fig:sig_by_dist}}
1435: \end{figure}
1436: \begin{figure}
1437: %{\epsfxsize=\linewidth \epsfbox{f7.eps}}
1438: \plotone{f7.eps}
1439: \figcaption{\small Fit amplitude as the orientation of the
1440: template, \ie the Euler angle $\gamma$, is changed.  For these fits,
1441: the location in longitude and latitude is unchanged.  Note that this
1442: grid finds a preferred orientation $2\degr$ from that found by the
1443: total convolver (due to the slightly different grids used.) \label{fig:sig_by_alpha}}
1444: \end{figure}
1445: 
1446: 
1447: 
1448: 
1449: Using the LILC map at higher resolution, $\ell_{\textrm{max}}=128$, gives an
1450: accuracy in the total convolver of $\pi/\ell_{\textrm{max}}=1\degr .4$.  The
1451: position returned is identical, with only the orientation one step of
1452: $1\degr .4$ different.  
1453: %  See:  /afs/mpa/planck/simdata/tjaffe/data/bianchi/wmap_fits_second/fits/lilc_ns64_r104_fits.out
1454: 
1455: 
1456: The above applies to the best-fit model at $(x,\Omega_0)=(0.62,0.5)$,
1457: but other models have structure at different angular scales.  In
1458: particular, for the region of small $x$ and $\Omega_0$, where a
1459: tightly wound spiral is even more tightly focused in one hemisphere,
1460: the fit amplitudes are far more dependent on the exact position.
1461: Because the total convolver resolution is $\pi/\ell_{\textrm{max}}$, our
1462: analysis is not as sensitive for this region of model space as it
1463: would be for a higher resolution analysis.  In these cases, the
1464: difference of a few degrees can mean a large difference in amplitude.
1465: Simulations show that, although the location returned is the closest bin
1466: to the true location, the amplitude of a model
1467: $(x,\Omega_0)=(0.1,0.1)$ is underestimated by $\sim20\%$ on average
1468: due to the limited resolution.  Increasing the resolution of the
1469: analysis to HEALPix $N_{\textrm{side}}=64$ increases the mean and
1470: brings it closer to the correct value, but it is still underestimated.
1471: (Higher resolution analysis with the total convolver is not feasible
1472: due to the memory and CPU requirements.)
1473: In the region of model space where $x>0.25$ and $\Omega_0>0.25$, this
1474: effect drops to less than a few percent.
1475: 
1476: A more detailed look at these models at increased resolution
1477: ($N_{\textrm{side}}=64$) shows no evidence that the lower resolution
1478: analysis missed a significant detection.  But the limits placed on
1479: shear and rotation are less stringent than they would be were a higher
1480: resolution analysis feasible.
1481: % see $rzgdata/work/bianchi/small_x_test/
1482: 
1483: 
1484: 
1485: 
1486: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1487: \subsection{DMR Fit}
1488: 
1489: Our best-fit amplitude is below the upper limit DMR could place on the
1490: shear.  Using this model, a fit to the DMR data gives
1491: % /afs/ipp/bc-b/tjaffe/work/bianchi/wmap_fits_individual2/kp0s/logs/tpl_simulfit_right_dmr.log
1492: $\left(\frac{\sigma}{H}\right)_0=3.38\pm.98\times10^{-10}$, which is
1493: within our best-fit error bar for the \emph{WMAP} data, but which is
1494: not distinguishable from a chance alignment for DMR.
1495: \citet{kogut:1997} report a distribution of $\Gamma$ values for chance
1496: alignments up to $4.5$.  Our fit value and error give $\Gamma=3.4$,
1497: and although this value comes from different methods and assumptions, it
1498: is roughly comparable.
1499: 
1500: 
1501: 
1502: \subsection{Sensitivity to Assumed Power Spectrum}\label{assumptions}
1503: 
1504: 
1505: As mentioned in \S\ \ref{sig_covar}, assumptions about the
1506: cosmological parameters go into this analysis from the beginning with
1507: the choice of the signal covariance matrix.  In effect, we are
1508: assuming that the CMB signal consists of an anisotropic
1509: Bianchi-induced component plus a statistically isotropic, Gaussian
1510: random field described completely by its power spectrum, which is 
1511: taken to be the \emph{WMAP} best-fit theoretical power law spectrum.
1512: As we are searching for evidence of a model that affects the power
1513: spectrum at large scales and that is inconsistent with inflation,
1514: this approach obviously lacks consistency.  
1515: 
1516: We have verified, however, that changing the assumed parameters and
1517: using, for example, a flat $Q=18\,\mu$K power law spectrum, or a completely
1518: implausible spectrum, has little effect (less than $3\%$) on the
1519: resulting best-fit amplitude and position for the Bianchi
1520: component. In fact, the power spectrum affects only the estimated
1521: significance of the result, as that significance is dependent on the
1522: expected level of large-scale CMB structure that drives chance
1523: alignments.  As shown in Figure \ref{fig:cls_compare}, correction for
1524: this Bianchi model lowers the large-scale power.  Our significance
1525: estimates are based on simulations generated assuming a higher level
1526: of large scale power, so the significance of the detection would
1527: increase when compared to an ensemble consistent with the corrected
1528: power spectrum.
1529: 
1530: 
1531: % Figure 8:  Power spectra
1532: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1533: %  From Frode
1534: \begin{figure}\begin{centering}
1535: %{\epsfxsize=\linewidth \epsfbox{f8.eps}}
1536: \plotone{f8.eps}
1537: \figcaption{\small Comparison of power spectra. The gray and black solid lines show the power spectrum estimated from the co-added V+W map before and
1538: after correcting for the Bianchi template, respectively.  The 
1539: dotted gray and black lines shows the theoretical best-fit
1540: power-spectra from the \emph{WMAP}-team analysis and
1541: \citet{hansen:2004b} respectively. The latter is a fit to the northern
1542: hemisphere data alone.  The dashed grey line is the power in the
1543: Bianchi template alone.\label{fig:cls_compare}}
1544: \end{centering}
1545: \end{figure}
1546: 
1547: 
1548: 
1549: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1550: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1551: \section{Implications}
1552: 
1553: There are several interesting results based on \emph{WMAP} first-year
1554: data that are inconsistent with the assumptions of isotropy and
1555: Gaussianity and that are immediately relevant to this study.  De
1556: Oliveira-Costa \etal (2004), \citet{land:2005}, and \citet{copi:2005}
1557: (and sources therein) examine the low-$\ell$ multipoles of the
1558: foreground-cleaned data and find that, in addition to the anomalously
1559: low quadrupole amplitude, the preferred axes of the quadrupole and
1560: octopole are anomalously well aligned in the direction of
1561: $(l,b)=(-110\degr,60\degr)$.  
1562: %
1563: \citet{eriksen:2004a} and \citet{hansen:2004a} find a system of
1564: reference (roughly aligned with the ecliptic) in which there is a
1565: significant difference in large-scale power between the two
1566: hemispheres at the 98\%-99\% level, with significantly more power in the
1567: south.
1568: %
1569: \citet{vielva:2004} and \citet{cruz:2005} detect non-Gaussianity in
1570: the \emph{WMAP} combined Q-V-W map using spherical wavelets; they find
1571: significant kurtosis in the wavelet coefficients at a scale of
1572: $10\degr$ and identify a cold spot at $(l,b)=(209\degr,-57\degr)$ as
1573: the probable source.  Our choice of models was partly motivated by the
1574: morphology of these anomalies, and indeed, subtracting for our
1575: best-fit Bianchi template corrects them.
1576: 
1577: 
1578: \subsection{Quadrupole Amplitude}\label{sec:quadrupole}
1579: 
1580: The quadrupole amplitude has been considered anomalously low since
1581: \emph{COBE} (see \citealt{de Oliveira-Costa:2004} and references
1582: therein).  As pointed out by \citet{jaffe:2005}, the correction for
1583: this Bianchi component raises the low quadrupole amplitude to a value
1584: more consistent with the theoretical power spectrum.  This result is
1585: unchanged with the best-fit model of this work, since the models are
1586: almost the same.  Should this be considered ``fine tuning''?  
1587: 
1588: We can simulate the situation by taking as the the primordial
1589: quadrupole the \emph{WMAP} quadrupole (as derived by
1590: \citealt{bielewicz:2004}) minus the quadrupole of our best-fit Bianchi
1591: model.  If we then add the Bianchi quadrupole at random orientations,
1592: we can see how likely it is that the resulting total quadrupole be as
1593: low as the observed \emph{WMAP} quadrupole.  We find that the
1594: likelihood is
1595: % email from HKE
1596: $\sim5\%$.  This implies that the level of ``fine tuning'' required to
1597: end up with the low observed quadrupole is not exceptional.
1598: % if one assumes
1599: %that it is the addition of a Bianchi component and a randomly oriented
1600: %primordial component.  In other words, the Bianchi correction does not
1601: %replace the anomaly of a low quadrupole with the anomaly of unusually
1602: %anti-aligned primordial and Bianchi quadrupoles.
1603: 
1604: We further take a set of 1000 simulated Gaussian CMB skies, with and
1605: without a Bianchi component, and fit our best-fit Bianchi model to
1606: them.  The ``corrected'' quadrupole is on average $\sim5\%$ lower than
1607: the original, which is to be expected considering that the fit is a
1608: least squares solution. In contrast, using the real LILC data, the
1609: correction has the effect of raising the quadrupole.  This happens in
1610: over $\sim20\%$ of the simulations, so while this is not the average
1611: behavior, it is not extraordinary.
1612: 
1613: 
1614: % From:
1615: %  /afs/ipp/bc-b/tjaffe/work/bianchi/ilc_sims_2nd_all_10k_better/two_models_all_10k/idl_journal.pro.bak_quadrupole
1616: %
1617: %  which uses
1618: %  /afs/ipp/bc-b/tjaffe/work/bianchi/ilc_sims_2nd_all_10k_better/two_models_all_10k/fits/quadrupole_differences.out 
1619: %  /afs/ipp/bc-b/tjaffe/work/bianchi/ilc_sims_2nd_all_10k_better/two_models_all_10k/fits/quadrupole_differences_pb.out 
1620: %  /afs/ipp/bc-b/tjaffe/work/bianchi/ilc_sims_2nd_all_10k_better/two_models_all_10k/compare_locations_lilcs_nocl2.pro
1621: 
1622: 
1623: 
1624: 
1625: \subsection{Low-$\ell$ Alignment and Planarity}\label{sec:lowl}
1626: 
1627: 
1628: De Oliveira-Costa et al.\ (2004), \citet{land:2005}, and
1629: \citet{copi:2005} discuss the statistically anomalous alignment of the
1630: quadrupole and octopole in the \emph{WMAP} data.  The preferred axes
1631: of the $\ell =2$ and $\ell =3$ multipoles are only $7\degr$ apart
1632: (roughly in the direction of $(l,b)=(-110\degr,60\degr)$), which is
1633: anomalous at the $99.3\%$ level compared to simulations.  After
1634: subtracting the best-fit Bianchi template, these axes lie $74\degr$
1635: apart, consistent (at $27\%$) with the statistically isotropic
1636: simulations (see Fig. \ref{fig:low_ls}).
1637: 
1638: The planarity of the low-$\ell$ multipoles has also been considered
1639: somewhat anomalous (see \citealt{de Oliveira-Costa:2004} and
1640: \citealt{land:2005} for a discussion).  The {\it t}-statistic defined by
1641: \citet{de Oliveira-Costa:2004} provides a measure of this planarity.
1642: Again, subtracting the Bianchi template lowers the significance of the
1643: the low-$\ell$ multipoles.  The planarity of the octopole in
1644: particular drops from a significance of $\sim 90\%$ (depending on
1645: whether the WILC or LILC is used) to $\sim 50\%$.  Figures
1646: \ref{fig:low_ls} (b) and (d) shows how the planarity of the octopole
1647: is disrupted.  This will also impact the results of multipole vector
1648: analyses such as that of \cite{copi:2005}.
1649: 
1650: %  Figure 9:  low-l multipoles
1651: %  From 
1652: % /afs/ipp/bc-b/tjaffe//work/bianchi/low_ls/
1653: \begin{figure}
1654: %{\epsfxsize=\linewidth \epsfbox{f9.eps}}
1655: \plotone{f9.eps}
1656: \figcaption{\small Low-$\ell$ multipoles of the WILC corrected ({\it bottom panels}) and uncorrected ({\it top panels}) for the Bianchi component. \label{fig:low_ls}}
1657: \end{figure}
1658: 
1659: 
1660: 
1661: \subsection{Large-Scale Power Asymmetry}\label{sec:asymmetry}
1662: 
1663: 
1664: \citet{eriksen:2004a} and \citet{hansen:2004a} reported that the
1665: large-scale power ($\ell \lesssim 40$) in the \emph{WMAP} data is
1666: anisotropically distributed over two opposing hemispheres (in the
1667: reference frame in which the z-axis points toward
1668: % Hansen 2004a says in abstract:
1669: %By reorienting the coordinate axes, we partition the sky into
1670: %different hemispheres and search for the reference frame which
1671: %maximizes the asymmetric distribution of power. The North Pole for
1672: %this coordinate frame is found to intersect the sphere at (80°, 57°)
1673: %in Galactic colatitude and longitude over almost the entire multipole
1674: %range \u2113= 5\u201340. 
1675: %
1676: %  Originally, had (10,57) but longitude comes first, so (57,10)
1677: $(l,b)=(57\degr,10\degr)$; see Fig. \ref{fig:power_asymmetry}), with
1678: a significance of $3\sigma$ compared with simulations.  Repeating the
1679: analysis and adopting the Kp2 sky coverage, we compare the corrected
1680: V+W \emph{WMAP} map with 2048 simulations. We find that $\sim 14\%$ of
1681: the simulations have a larger maximum power asymmetry ratio than the
1682: Bianchi-corrected map, whereas only $0.7\%$ have a larger ratio than
1683: the uncorrected data (see Fig.
1684: \ref{fig:power_asymmetry}).  It is apparent that the maximum power
1685: ratio between any two hemispheres is significantly suppressed after
1686: subtracting the Bianchi template, as no asymmetry axis is found at any
1687: statistically significant level.  It is apparent from that figure,
1688: however, that some residual power asymmetry remains.  This comes
1689: largely from the range $20<l<40$, where the Bianchi template has little
1690: power, indicating that a model with more small-scale structure may be
1691: needed.
1692: 
1693: 
1694: 
1695: %  Figure 10:  power ratio
1696: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1697: %       From Frode
1698: \begin{figure} \centering
1699: %{\epsfxsize=6cm \epsfbox{ratmap_rot_merged.ps}}
1700: %{\epsfxsize=6cm \epsfbox{f10.eps}}
1701: \plotone{f10.eps}
1702: %\psfig{file=ratmap_rot_merged.ps, height=\linewidth, angle=90}
1703: \figcaption{\small Power ratio between hemispheres in \emph{WMAP} ILC, corrected (\emph{bottom}) and uncorrected (\emph{bottom}) for the best-fit Bianchi component.
1704: \label{fig:power_asymmetry}}
1705: \end{figure}
1706: 
1707: 
1708: 
1709: 
1710: \subsection{Wavelet Kurtosis}\label{sec:wavelet}
1711: 
1712: 
1713: \citet{vielva:2004} and \citet{cruz:2005} used a wavelet technique to detect an excess of
1714: kurtosis in the wavelet coefficients and isolate an unusually cold
1715: spot ($\sim 3\sigma$ significance relative to Gaussian simulations) at
1716: Galactic coordinates $(l,b)=(209^{\circ},-57^{\circ})$.  Referring
1717: again to Figure
1718: \ref{fig:ilc_b_bilc}, we see that a cold spot is indeed present at the
1719: right location, in the form of the center of the spiral.
1720: 
1721: We therefore also repeat the analysis of \citet{vielva:2004}, and
1722: compute the kurtosis of the wavelet coefficients as a function of
1723: scale from both the WILC and the corresponding Bianchi-subtracted
1724: map. A $|b|<20\degr$ galactic cut is imposed in this case, for
1725: computational convenience.
1726: 
1727: The results from this exercise are reported in Figure
1728: \ref{fig:kurtosis} After subtracting the Bianchi template, the significance
1729: of the southern hemisphere anomaly is greatly reduced, and no new
1730: non-Gaussian features have been introduced.
1731: 
1732: 
1733: % Figure 11:  kurtosis
1734: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1735: %       From Frode
1736: \begin{figure}
1737: %{\epsfxsize=\linewidth \epsfbox{f11.eps}}
1738: \plotone{f11.eps}
1739: \figcaption{\small Kurtosis in wavelet coefficients. The  boxes and
1740:  crosses show the kurtosis before and after subtracting the
1741: Bianchi template, respectively, computed from the southern ({\it
1742: dotted line}) and northern ({\it solid line}) Galactic hemispheres.\label{fig:kurtosis}}
1743: \end{figure}
1744: %
1745: 
1746: 
1747: 
1748: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1749: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1750: \section{Discussion and Conclusions}
1751: 
1752: We have considered a fast and efficient method for fitting a template
1753: to the full sky in harmonic space and finding the best-fit location
1754: and orientation.  The total convolver algorithm evaluates the
1755: correlation between the sky and the template at every possible
1756: relative orientation using fast Fourier transforms.  With this
1757: algorithm, the search for the best-fit becomes 2 orders of magnitude
1758: faster than the corresponding search performed one rotation at a time.
1759: This method, along with pixel-space simultaneous foreground fitting,
1760: provides a powerful tool for testing any deterministic model for
1761: anisotropy in the CMB.  Simulations generated by the LILC pipeline
1762: allow us to quantify the bias, investigate the effects of
1763: foreground contaminants, and show how well each of these methods
1764: detects a known input.
1765: 
1766: 
1767: We have applied this method to the first-year \emph{WMAP} data to
1768: search for evidence of shear and vorticity using templates derived for
1769: Bianchi type VII$_{h}$ universes.  We find a surprisingly significant
1770: correlation between the \emph{WMAP} data and a right-handed Bianchi model
1771: with $x=0.62$, $\Omega_0=0.5$, and shear of
1772: $\left(\frac{\sigma}{H}\right)_0=4.3\pm 0.8\times 10^{-10}$,
1773: implying a vorticity of
1774: $\left(\frac{\omega}{H}\right)_0=9.5\times10^{-10}$.  The center of
1775: the spiral structure lies at approximately
1776: $(l,b)=(222\degr,-62\degr)$.  Simulations show that this amplitude is
1777: likely to be biased by $\sim7\%$, implying a true amplitude closer to
1778: $4.0\times10^{-10}$.  Incomplete sky fits, simultaneous foreground
1779: fitting, and fits to a set of Gibbs samples are all consistent with
1780: this amplitude and indicate that confusion with Galactic emission is
1781: unlikely to contribute significantly to this detection.
1782: 
1783: 
1784: Correcting the \emph{WMAP} data for the effect of the best-fit model
1785: solves several problems seen in the data.  The corrected maps show
1786: significantly reduced power asymmetry between any two hemispheres.
1787: The correction also eliminates the non-Gaussian kurtosis in the
1788: wavelet coefficients detected by
1789: \citet{vielva:2004} and \citet{cruz:2005}, raises the low measured
1790: quadrupole by a factor of 2, and disrupts the planarity of the
1791: octopole and its anomalous alignment with the quadrupole.  In short,
1792: the data appear far more Gaussian and isotropic after correction.
1793: 
1794: 
1795: The original analyses by \citet{kogut:1997} and \citet{bunn:1996} were
1796: limited by the signal-to-noise ratio level in the DMR instrument.  Our best-fit
1797:  result is just under their upper limit but still significant due
1798: to \emph{WMAP}'s greatly improved signal-to-noise ratio.  Furthermore, the
1799: Kogut analysis searched a coarse ($\sim 10\degr$) grid of possible
1800: locations and orientations, while with the total convolver, we can
1801: efficiently search a finer grid.  
1802: 
1803: How likely is it that our best-fit model is a true detection rather
1804: than a chance alignment?  Considering the best-fit model by itself and
1805: comparing its fit amplitude to simulations, it is higher than $99.7\%$
1806: of simulations.  However, the simulations also show that $10\%-20\%$ of
1807: Gaussian, statistically isotropic skies will have one of the Bianchi
1808: models appear as significant.  Considering the fact that the sky is
1809: approximately Gaussian and isotropic, one would not expect to find a
1810: more definitive detection based on template fitting alone.  But the
1811: distribution of chance alignments in the simulations is sensitive to
1812: the amount of large-scale power assumed, and that is lowered by the
1813: Bianchi correction to the \emph{WMAP} data.  Furthermore, the
1814: cumulative probability that a chance alignment not only fits at the
1815: level of our best-fit model but also has the effect of resolving the
1816: several anomalies in the data must also be considered in any
1817: qualitative judgment of the significance of this result.
1818: 
1819: Further improvement to the data will not refine these measures
1820: significantly, because at the \emph{WMAP} sensitivity level, the
1821: analysis is already very close to the expected distribution of chance
1822: alignments in the absence of noise.  Improved foreground subtraction
1823: will, however, remove some of the possible confusion and bias,  but
1824: neither higher resolution nor higher signal-to-noise ratio data should
1825: change this result nor be able to provide additional information
1826: concerning the question of whether the fit is a real detection of
1827: vorticity and shear.  Answering that question will require additional
1828: verifiable predictions for the effects of vorticity and shear on other
1829: observables.
1830: 
1831: However, in the context of the anomalies that this hypothesis can explain,
1832: the possible detection is certainly provocative.  The most important
1833: result of this analysis is that a model with vorticity and shear can
1834: explain the observed asymmetry in the CMB anisotropies and the
1835: non-Gaussian cold spot.  Note that this asymmetry exists only in the
1836: $\Omega_0<1$ versions of these Bianchi models.  Significant evidence
1837: currently indicates that $\Omega_0$ is very close to 1, so our best-fit
1838:  model cannot be considered physically realistic.  However, as
1839: mentioned in \S\ \ref{models}, Barrow \etal (1985) did not include any dark
1840: energy component.  Furthermore, the Bianchi model does not include a
1841: mechanism to generate structure at the surface of last scattering.  A
1842: self-consistent theory is required that can explain the small scale
1843: fluctuations, and in particular the acoustic peaks, in the context of
1844: a Universe with shear and vorticity.  But from a pragmatic point of
1845: view, one can conclude that, regardless of the viability of the
1846: particular Bianchi model, this result gives a measure of the
1847: significant deviation from isotropy in the data.
1848: 
1849: 
1850: We consider this result to be further motivation for considering ideas
1851: outside of the so-called concordance model of cosmology.  There are
1852: anomalies in the data that are inconsistent with the theory of a
1853: Gaussian, statistically isotropic universe, and Bianchi models are
1854: only one such anisotropic model that merits investigation.  We have
1855: demonstrated a method of template fitting that can be applied to test
1856: any model that makes a deterministic prediction for an anisotropy
1857: pattern in the CMB.  The best-fit Bianchi model provides a template
1858: temperature pattern that can explain the observed anomalies in the
1859: data and that describes the morphology theorists may need to reproduce
1860: in considering alternatives to the standard cosmological model.
1861: 
1862: 
1863: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1864: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1865: \section*{Acknowledgments}
1866: 
1867: We are grateful to M.~Demianski, S.~Hervik, and S.~D.~M.~White for
1868: useful discussions, and to J.~McEwen for finding errors in the
1869: original numbers.  H.~K.~E. acknowledges financial support from the
1870: Research Council of Norway, including a Ph.\ D. scholarship. F.~K.~H. was
1871: supported by a Marie Curie Re-integration Grant within the 6th
1872: European Community Framework Programme.  We acknowledge use of the
1873: HEALPix software %(G\'orski, Hivon \& Wandelt 1998)
1874: \citep{healpix} and analysis package for deriving the results in this
1875: paper.  We also acknowledge use of the Legacy Archive for Microwave
1876: Background Data Analysis (LAMBDA).
1877: 
1878: 
1879: 
1880: \begin{thebibliography}{}
1881: 
1882: 
1883: \bibitem[Banday et al.(1996)]{banday:1996} Banday, A.~J., G\'orski, K.~M., Bennett, C.~L., Hinshaw, G., Kogut, A., Smoot, G.~F., 1996, \apj, 468, L85
1884: 
1885: \bibitem[Barrow et al.(1985)]{barrow:1985} Barrow, J. D., Juszkiewicz,  R., \& Sonoda, D. H. 1985, \mnras, 213, 917
1886: 
1887: \bibitem[Bennett et al.(2003a)]{bennett:2003a} Bennett, C.~L.~\etal\ 
1888: 2003a, \apjs, 148, 1 
1889: 
1890: \bibitem[Bennett et al.(2003b)]{bennett:2003b} Bennett, C.~L.~\etal\ 
1891: 2003b, \apjs, 148, 97
1892: 
1893: \bibitem[Bielewicz et al.(2004)]{bielewicz:2004}
1894: Bielewicz, P., G\'orski, K. M., \& Banday, A. J., 2004, \mnras, 355, 1283
1895: 
1896: \bibitem[Bunn et al.(1996)]{bunn:1996}Bunn E.~F., Ferreira P.~G., \& Silk J., 1996, PRDL, 77, 14, 2283 
1897: 
1898: \bibitem[Collins \& Hawking (1973)]{collins:1973} Collins, C.~B. \& 
1899: Hawking, S.~W., 1973, MNRAS, 162, 307
1900: 
1901: %\bibitem[Copi et al.(2004)]{copi:2004} 
1902: %Copi, C. J., Huterer, D., \&  Starkman, G. D.\ 2004, \prd, 70, 043515
1903: %%%CITATION = ASTRO-PH 0310511;%%
1904: 
1905: \bibitem[Copi et al.(2006)]{copi:2005}
1906: Copi, C.~J., Huterer, D., Schwarz, D.~J. \& Starkman, G. D., 2005,
1907: MNRAS, 367, 79
1908: %astro-ph/0508047
1909: 
1910: 
1911: \bibitem[Cruz et al.(2005)]{cruz:2005}
1912: Cruz, M., Martinez-Gonzalez, E., Vielva, P. \& Cayon, L. 2005, \mnras,
1913: 356, 29
1914: 
1915: \bibitem[Davies et al.(1996)]{davies:1996} 	
1916: Davies, R. D., Watson, R. A., \& Gutierrez, C. M., 1996, \mnras, 278, 925
1917: 
1918: \bibitem[de Oliveira-Costa et al.(2004)De Oliveira-Costa et al.(2004)]{de Oliveira-Costa:2004}
1919: de Oliveira-Costa, A., Tegmark, M., Zaldarriaga, M., \& Hamilton,
1920: A. 2004, \prd, 69, 063516
1921: 
1922: \bibitem[Dickinson et al.(2003)]{dickinson:2003} Dickinson ,C.,
1923: Davies, R.~D., \& Davis, R.~J., 2003, MNRAS, 341, 369 
1924: 
1925: 
1926: \bibitem[Eriksen et al.(2004a)]{eriksen:2004b}
1927: Eriksen, H. K., Banday, A. J., G\'orski, K. M., \& Lilje, P. B. 2004a,
1928: \apj, 612, 633
1929: 
1930: \bibitem[Eriksen et al.(2005)]{eriksen:2005}
1931:   Eriksen, H. K., Banday, A. J., G\'{o}rski, K. M., \& Lilje,
1932:   P. B. 2005, [astro-ph/0508196]
1933: 
1934: \bibitem[Eriksen et al.(2004b)]{eriksen:2004a} Eriksen, H.~K., Hansen, 
1935: F.~K., Banday, A.~J., G{\' o}rski, K.~M., \& Lilje, P.~B.\ 2004b, \apj, 605, 
1936: 14 
1937: 
1938: \bibitem[Eriksen et al.(2004c)]{eriksen:2004c} Eriksen, H. K., O'Dwyer, I. J., Jewell, J. B., Wandelt, B. D., Larson,D. L., Górski, K. M., Levin, S., Banday, A. J., \& Lilje, P. B., 2004c, ApJS, 155, 227 
1939: 
1940: \bibitem[Finkbeiner et al.(1999)]{finkbeiner:1999} % FDS models for thermal dust emission
1941: Finkbeiner, D. P., Davis, M., \& Schlegel, D. J. 1999, ApJ, 524, 867
1942: 
1943: %\bibitem[Finkbeiner (2003a)]{finkbeiner:2003a} Finkbeiner, D.~P., 2003, ApJ, 146, 407
1944: 
1945: \bibitem[Finkbeiner (2004)]{finkbeiner:2004} Finkbeiner, D.~P., 2004,
1946: ApJ, 614, 186
1947: 
1948: \bibitem[Finkbeiner(2003)]{finkbeiner:2003} Finkbeiner, D.~P.\ 2003, \apjs, 146, 407
1949: 
1950: %\bibitem[G\'orski, Hivon, \& Wandelt (1999a)]{gorski:1999} G\'orski, K.  M., Hivon,
1951: % E., \& Wandelt, B.  D., 1999, in Evolution of Large-Scale Structure:
1952: % from Recombination to Garching, ed.  A.  J.  Banday, R.  K.  Sheth,
1953: % \& L.  N.  da Costa (Garching, Germany: European Southern
1954: % Observatory), 37
1955: 
1956: 
1957: \bibitem[G\'orski et al.(1996)]{gorski:1996} G\'orski, K. M., Banday
1958:   A. J., Bennett C. L., Hinshaw G., Kogut A., Smoot G. F., \& Wright
1959:   E. L. 1996, \apj, 464, L11
1960: 
1961: \bibitem[G{\' o}rski et al.(2005)]{healpix} G{\' o}rski, K.~M., 
1962: Hivon, E., Banday, A.~J., Wandelt, B.~D., Hansen, F.~K., Reinecke, M., \& 
1963: Bartelmann, M.\ 2005, \apj, 622, 759 
1964: 
1965: %\bibitem[G\'orski et al.(1999)]{healpix}  G\'orski,
1966: %K.~M.,  Hivon, E., \& Wandelt, B.~D., 1999, ``Analysis Issues for Large
1967: %CMB Data Sets'', in {\it Proceedings of the MPA/ESO Cosmology
1968: %Conference "Evolution of Large-Scale Structure"}, eds. A.J. Banday,
1969: %R.S. Sheth and L. Da Costa, PrintPartners Ipskamp, NL, pp. 37-42 
1970: 
1971: \bibitem[Haslam et al.(1982)]{haslam:1982} % ubiquitous 408 MHz reference
1972: Haslam, C. G. T., Salter, C. J., Stoffel, H., \& Wilson, W.\ 1982, \aaps, 47, 1 
1973: 
1974: \bibitem[Hansen et al.(2004a)]{hansen:2004b} 
1975: Hansen, F. K., Balbi, A., Banday, A.~J., \& G{\' o}rski, K.~M.\ 2004a, 
1976: \mnras, 354, 905
1977: 
1978: \bibitem[Hansen et al.(2004b)]{hansen:2004a} 
1979: Hansen, F. K., Banday, A.~J., \& G{\' o}rski, K.~M.\ 2004b, 
1980: \mnras, 354, 641
1981: 
1982: %\bibitem[Hivon et al.(2002)]{hivon:2002} Hivon, E., G{\' o}rski, 
1983: %K.~M., Netterfield, C.~B., Crill, B.~P., Prunet, S., \& Hansen, F.\ 2002, 
1984: %\apj, 567, 2 
1985: 
1986: %\bibitem{Jaffe et al.(2004)]{jaffe:2004} Jaffe, A.~H., Balbi,A., Bond,J.~R., Borrill,J., Ferreira,P.~G., Finkbeiner, D., Hanany, S., Lee, A.~T., Rabii, B., Richards,P.~L., Smoot,G.~F.,Stompor, R., Winant, C.~D.,\&  Wu,J.~H.~P., 2004, ApJ, 615, 55
1987: 
1988: \bibitem[Jaffe et al.(2005)]{jaffe:2005} Jaffe, T.~R., Banday, A.~J., Eriksen, H.~K., G\'orski, K.~M., \& Hansen, F.~K., 2005, ApJ, 629, L1
1989: 
1990: \bibitem[Jewell et al.(2004)]{jewell:2004} Jewell, J., Levin, S., \& Anderson, C. H., 2004, \apj, 609, 1
1991: 
1992: \bibitem[Kogut et al.(1997)]{kogut:1997} 
1993:   Kogut, A., Hinshaw, G., Banday, A. J. 1997, \prd, 55, 4, 1901 
1994: 
1995: \bibitem[Land \& Magueijo(2005)]{land:2005}
1996: Land, K., \& Magueijo, J. 2005, preprint (astro-ph/0502237)
1997: 
1998: \bibitem[Mortlock et al.(2002)]{mortlock:2002} Mortlock, D.~J., Challinor, A.~D., \& Hobson, M.~P., 2002, MNRAS, 330, 405
1999: 
2000: %\bibitem[O'Dwyer et al.(2004)]{odwyer:2004} O'Dwyer, I.~J., \etal\ 
2001: %2004, \apjl, 617, L99 
2002: 
2003: \bibitem[Patanchon et al.(2004)]{patanchon:2004}  Patanchon, G. ,
2004: Cardoso, J.-F., Delabrouille, J., \& Vielva, P., 2006, \mnras, 364, 1185
2005: 
2006: %\bibitem[Slosar \& Seljak(2004)]{slosar:2004}
2007: %Slosar, A., \& Seljak, U. 2004, \prd 70, 083002
2008: 
2009: \bibitem[Schlegel et al.(1998)]{schlegel:1998} Schlegel D. J., Finkbeiner, D. P., \& Davis, M. 1998, \apj, 500, 525 
2010: 
2011: \bibitem[Tegmark et al.(2003)]{tegmark:2003}
2012: Tegmark, M., \& de Oliveira-Costa A., \& Hamilton A., 2003, \prd 68,
2013: 123523
2014: 
2015: \bibitem[Vielva et al.(2004)]{vielva:2004}
2016: Vielva, P., Mart\'\i nez-Gonz\'alez, E., Barreiro, R. B., Sanz, J. L., \&
2017: Cay\'on, L. 2004, \apj, 609, 22
2018: 
2019: \bibitem[Wandelt \& G\'orski(2001)]{wandelt:2001} Wandelt, B. D. \&
2020:   G\'orski, K. M. 2001, \prd, 63, 123002 
2021: 
2022: 
2023: \bibitem[Wandelt et al.(2004)]{wandelt:2004} Wandelt, B. D., Larson, D. L., \& Lakshminarayanan, A., 2004, \prd, 70, 083511
2024: 
2025: 
2026: \end{thebibliography}
2027: 
2028: 
2029: 
2030: 
2031: 
2032: 
2033: 
2034: 
2035: 
2036: 
2037: 
2038: \end{document}
2039: 
2040: 
2041: 
2042: 
2043: 
2044: %  Table 1:
2045: %
2046: \begin{deluxetable}{lcccc}
2047: %
2048: \tablecaption{OLD Fitted template amplitudes \label{tab:fits_summary} }
2049: 
2050: %
2051: \tablewidth{0pt}
2052: %
2053: \tablecolumns{4}
2054: %
2055: \tablehead{  & $\left(\frac{\sigma}{H}\right)_0$ & $\left(\frac{\omega}{H}\right)_0$ &  $P\left(|\alpha_{\textrm{sim}}| < |\alpha_{\textrm{obs}}|\right)$   \\
2056: Map  & ${\scriptstyle (\times10^{-10})}$ & ${\scriptstyle (\times10^{-10})}$ & \%  }
2057: %
2058: \startdata
2059: %  Full sky fits from:
2060: %  /afs/mpa/planck/simdata/tjaffe/data/bianchi/wmap_fits_second/write_table.pro
2061: % using
2062: %  /afs/mpa/planck/simdata/tjaffe/data/bianchi/wmap_fits_second/fits/{wilc,lilc,tcm}_{left,right}_noise3uK_fits.out
2063: %  /afs/ipp/bc-b/tjaffe//work/bianchi/ilc_sims_2nd_all_10k_better/two_models_all_10k/fits/lilcs_noise3uK_{left,right}_fits.out
2064: 
2065: %
2066: %  Cut-sky fits from 
2067: %/afs/ipp/bc-b/tjaffe/work/bianchi/wmap_fits_individual/dust_sfd/scripts/read.pro
2068: %
2069: % Gibbs fits from 
2070: %/afs/ipp/bc-b/tjaffe/work/bianchi/gibbs_fits/read2.pro
2071: %
2072: %\multicolumn{4}{c}{Right-handed $(x,\Omega_0)=(0.62,0.5)$} \\
2073: \cutinhead{Right-handed $(x,\Omega_0)=(0.62,0.5)$} \\
2074: 
2075: WILC & $ 2.57\pm 0.49$ & $ \phm{-}5.69$ & $ 99.8\phn$ \\
2076: LILC & $ 2.55\pm 0.49$ & $ \phm{-}5.64$ & $ 99.7\phn$ \\
2077: TOH & $ 2.40\pm 0.49$ & $ \phm{-}5.30$ & $ 98.6\phn$ \\
2078: 
2079: % Cut-sky fits
2080: K\tablenotemark{a} & $ 1.65( 2.39)\pm 0.49$ & $\phm{-}3.65$ & $ 28.90$ \\
2081: Ka\tablenotemark{a} & $ 2.16( 2.45)\pm 0.49$ & $\phm{-}4.78$ & $ 91.90$ \\
2082: Q\tablenotemark{a} & $ 2.31( 2.49)\pm 0.49$ & $\phm{-}5.11$ & $ 97.10$ \\
2083: V\tablenotemark{a} & $ 2.45( 2.54)\pm 0.49$ & $\phm{-}5.41$ & $ 99.20$ \\
2084: W\tablenotemark{a} & $ 2.50( 2.61)\pm 0.49$ & $\phm{-}5.52$ & $ 99.50$ \\
2085: QVW\tablenotemark{a} & $ 2.36( 2.51)\pm 0.49$ & $\phm{-}5.22$ & $ 98.10$ \\
2086: VW\tablenotemark{a} & $ 2.45( 2.55)\pm 0.49$ & $\phm{-}5.42$ & $ 99.20$ \\
2087: Q-V\tablenotemark{a} & $ 0.02( 0.04)\pm 0.001$ & $\phm{-} 0.05$ & $ 91.80$ \\
2088: V-W\tablenotemark{a} & $ -0.04( -0.06)\pm 0.001$ & $ -0.10$ & $ 98.40$ \\
2089: Q-W\tablenotemark{a} & $ -0.02( -0.02)\pm 0.001$ & $ -0.04$ & $ 78.40$ \\
2090: 
2091: 
2092: %  Gibbs samples:
2093: Q\tablenotemark{b} & $ 2.43\pm 0.06^c$\tablenotemark{c} & \phm{-} 5.37 & - \\
2094: V\tablenotemark{b} & $ 2.44\pm 0.06^c$\tablenotemark{c} & \phm{-} 5.40 & - \\
2095: W\tablenotemark{b} & $ 2.45\pm 0.06^c$\tablenotemark{c} & \phm{-} 5.42 & - \\
2096: 
2097: %\multicolumn{4}{c}{Left-handed $(x,\Omega_0)=(0.62,0.15)$} \\
2098: \cutinhead{Left-handed $(x,\Omega_0)=(0.62,0.15)$} \\
2099: WILC & $ 2.11\pm 0.42$ & $ \phm{-}19.70$ & $ 97.9\phn$ \\
2100: LILC & $ 2.20\pm 0.42$ & $ \phm{-}20.57$ & $ 99.4\phn$ \\
2101: TOH & $ 2.17\pm 0.42$ & $ \phm{-}20.26$ & $ 99.0\phn$ \\
2102: 
2103: % Cut-sky fits:
2104: 
2105: K\tablenotemark{a} & $ 2.07( 3.10)\pm 0.44$ & $ \phm{-}19.31$ & $ 95.10$ \\
2106: Ka\tablenotemark{a} & $ 1.98( 2.38)\pm 0.44$ & $ \phm{-}18.55$ & $ 92.70$ \\
2107: Q\tablenotemark{a} & $ 2.03( 2.26)\pm 0.44$ & $ \phm{-}18.94$ & $ 93.90$ \\
2108: V\tablenotemark{a} & $ 2.06( 2.19)\pm 0.44$ & $ \phm{-}19.29$ & $ 95.40$ \\
2109: W\tablenotemark{a} & $ 2.06( 2.22)\pm 0.43$ & $ \phm{-}19.21$ & $ 95.60$ \\
2110: QVW\tablenotemark{a} & $ 2.03( 2.23)\pm 0.44$ & $ \phm{-}18.99$ & $ 94.60$ \\
2111: VW\tablenotemark{a} & $ 2.07( 2.20)\pm 0.44$ & $ \phm{-}19.33$ & $ 95.50$ \\
2112: Q-V\tablenotemark{a} & $ 0.03( 0.04)\pm 0.002$ & $ \phm{-}\phn0.28$ & $ 84.00$ \\
2113: V-W\tablenotemark{a} & $ -0.04( -0.07)\pm 0.002$ & $ -0.39$ & $ 88.80$ \\
2114: Q-W\tablenotemark{a} & $ -0.01( -0.03)\pm 0.002$ & $ -0.11$ & $ 35.40$ \\
2115: 
2116: %  Gibbs samples:
2117: Q\tablenotemark{b} & $ 1.86\pm 0.09^c$\tablenotemark{c} & \phm{-} 17.40 & - \\
2118: V\tablenotemark{b} & $ 1.84\pm 0.10^c$\tablenotemark{c} & \phm{-} 17.21 & - \\
2119: W\tablenotemark{b} & $ 1.85\pm 0.08^c$\tablenotemark{c} & \phm{-} 17.28 & \phn\phn- 
2120: 
2121: \enddata
2122: %
2123: \tablecomments{{\bf OLD.}}
2124: 
2125: 
2126: \end{deluxetable}
2127: 
2128: 
2129: 
2130: