1: \documentclass[useAMS,usenatbib]{mn2e}
2: \usepackage{graphicx}
3: \bibliographystyle{mn2e}
4: %%%%% AUTHORS - PLACE YOUR OWN MACROS HERE %%%%%
5: %Reference list commands
6: \def\aj{AJ}
7: \def\araa{Ann. Rev. Astron. Astrophys.}
8: \def\apj{ApJ}
9: \def\apjl{Astrophys. J. Lett.}
10: \def\apjs{ApJS}
11: \def\ao{{Appl.~Opt.}}
12: \def\aap{A\&A}
13: \def\aapr{{A\&A~Rev.}}
14: \def\aaps{{A\&AS}}
15: \def\azh{{AZh}}
16: \def\baas{{BAAS}}
17: \def\jrasc{{JRASC}}
18: \def\memras{{MmRAS}}
19: \def\mnras{MNRAS}
20: \def\pra{{Phys.~Rev.~A}}
21: \def\prb{{Phys.~Rev.~B}}
22: \def\prc{{Phys.~Rev.~C}}
23: \def\prd{{Phys.~Rev.~D}}
24: \def\pre{{Phys.~Rev.~E}}
25: \def\rmp{{Rev.~Mod.~Phys.}}
26: \def\prl{{Phys.~Rev.~Lett.}}
27: \def\pasp{{Publ. Astron. Soc. Pacific}}
28: \def\pasj{{Publ. Astron. Soc. Japan}}
29: \def\qjras{{QJRAS}}
30: \def\skytel{{S\&T}}
31: \def\solphys{{Sol.~Phys.}}
32: \def\sovast{{Soviet~Ast.}}
33: \def\ssr{{Space~Sci.~Rev.}}
34: \def\zap{{ZAp}}
35: \def\nat{{Nat}}
36: \def\iaucirc{{IAU~Circ.}}
37: \def\aplett{{Astrophys.~Lett.}}
38: \def\apspr{{Astrophys.~Space~Phys.~Res.}}
39: \def\bain{{Bull.~Astron.~Inst.~Netherlands}}
40: \def\fcp{{Fund.~Cosmic~Phys.}}
41: \def\gca{{Geochim.~Cosmochim.~Acta}}
42: \def\grl{{Geophys.~Res.~Lett.}}
43: \def\jcp{{J.~Chem.~Phys.}}
44: \def\jsp{{J.~Statrecent study.~Phys.}}
45: \def\jgr{{J.~Geophys.~Res.}}
46: \def\jqsrt{{J.~Quant.~Spec.~Radiat.~Transf.}}
47: \def\memsai{{Mem.~Soc.~Astron.~Italiana}}
48: \def\nphysa{{Nucl.~Phys.~A}}
49: \def\physrep{Phys.~Rep.}
50: \def\physscr{{Phys.~Scr}}
51: \def\planss{{Planet.~Space~Sci.}}
52: \def\procspie{{Proc.~SPIE}}
53: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
54: \newcommand{\beq}{\begin{equation}}
55: \newcommand{\eeq}{\end{equation}}
56:
57:
58: \title[ ]{Multi-scale morphology of the galaxy distribution}
59: \author[ ]{Enn Saar$^{1}$, Vicent J. Mart\'{\i}nez$^{2}$,
60: Jean-Luc Starck$^{3}$ and David L. Donoho$^{4}$\\
61: $^{1}$Tartu Observatoorium, T\~oravere, 61602, Estonia\\
62: $^{2}$ Observatori Astron\`omic, Universitat de Val\`encia, Apartat
63: de Correus 22085, E-46071 Val\`encia, Spain \\
64: $^{3}$ CEA-Saclay, DAPNIA/SEDI-SAP, Service d'Astrophysique, F-91191 Gif
65: sur Yvette, France \\
66: $^{4}$ Department of Statistics, Stanford University, Sequoia Hall,
67: Stanford, CA 94305, USA}
68: \begin{document}
69: %\date{Accepted 1988 December 15. Received 1988 December 14; in original form 1988 October 11}
70:
71: \pagerange{\pageref{firstpage}--\pageref{lastpage}} \pubyear{2002}
72:
73: \maketitle
74:
75: \label{firstpage}
76:
77: \begin{abstract}
78: Many statistical methods have been proposed in the last years for
79: analyzing the spatial distribution of galaxies. Very few of them,
80: however, can handle properly the border effects of complex
81: observational sample volumes. In this paper, we first show how to
82: calculate the Minkowski Functionals (MF) taking into account these
83: border effects. Then we present a multiscale extension of the MF
84: which gives us more information about how the galaxies are spatially
85: distributed. A range of examples using Gaussian random fields
86: illustrate the results. Finally we have applied the Multiscale
87: Minkowski Functionals (MMF) to the 2dF Galaxy Redshift Survey data.
88: The MMF clearly indicates an evolution of morphology with scale. We
89: also compare the 2dF real catalog with mock catalogs and found that
90: $\Lambda$CDM simulations roughly fit the data, except at the finest
91: scale.
92: \end{abstract}
93:
94: \begin{keywords}
95: methods: data analysis -- methods: statistical -- large-scale structure of universe
96: \end{keywords}
97: \section{Introduction\label{sec:intro}}
98:
99: One of the main tenets of the present inflationary paradigm is the
100: assumption of Gaussianity for the primordial density perturbations.
101: This postulate forms the basis of present theories of formation and
102: evolution of large-scale structure in the universe, and of its
103: subsequent analysis. But it remains a hypothesis that needs to be
104: checked.
105:
106: The most straightforward way to do that would be to follow the
107: definition of Gaussian random fields (see, e.g, \citet{adler}) --
108: their one-point probability distribution and all many-point joint
109: probability distributions of field amplitudes have to be Gaussian.
110: This is clearly a too formidable task. Another way is to check the
111: relationships between the correlation functions and power spectra of
112: different orders, which are well-defined for Gaussian random fields.
113: This approach is frequently used (see, e.g., a review in
114: \citet{martinezsaar}). A third method is to study the morphology of
115: the cosmological (density) fields. One approach to the morphological
116: description relies on the so-called Minkowski functionals and is
117: complementary to the moment-based methods because these functionals depend
118: on moments of all orders. This procedure has been usually referred to
119: as topological analysis. It has a quite long history
120: already, starting with the seminal paper by \citet{gott86}, that
121: deals with the genus, a quantity closely related to one of the four
122: Minkowski functionals. The approach in the present paper lies within
123: this latter framework; we describe it in detail in
124: Sec.~\ref{sec:morph}.
125:
126: There are two different possibilities to develop a morphological
127: analysis of galaxy catalogues based on Minkowski functionals. First,
128: we can dress all points (galaxies) with spheres of a given radius,
129: and study the morphology of the surface that is generated by the
130: convex union of these spheres, as a function of the radius, which
131: acts here as the diagnostic parameter. An appropriate theoretical
132: model to compare with in this case is a Poisson point process. On
133: the other hand, if we wish to study the morphology of the underlying
134: realization of a random field, we have to restore the (density)
135: field first, to choose an isodensity surface corresponding to a
136: given density threshold and to calculate its morphological
137: descriptors. In this approach the density threshold (or a related
138: quantity) acts as the diagnostic parameter. The theoretical
139: reference model is that of a Gaussian random field, and the crucial
140: point here is to properly choose a restoration method that
141: provides a smoothed underlying density field that should be fairly
142: sampled by the observed discrete point distribution.
143:
144: Starting from the original paper on topology by \citet{gott86}, this
145: task has been typically done by smoothing the point distribution with a
146: Gaussian kernel. The choice of the optimal width of this kernel has
147: been widely discussed; it is usually taken close to the correlation
148: length of the point distribution. However, Gaussian smoothing is not
149: the best choice for morphological studies. As we have shown recently
150: \citep{mart05}, it tends to introduce additional Gaussian features
151: even for manifestly non-Gaussian density distributions. Minkowski
152: functionals are very sensitive to small density variations, and the
153: wings of Gaussian kernels could be wide enough to generate a
154: small-amplitude Gaussian ripple that is added to the true density
155: distribution. Such effect could be alleviated by using compact
156: adaptive smoothing kernels, as we show below.
157:
158: It is well known that large-scale cosmological fields have a
159: multi-scale structure. A good example is the density field; it
160: includes components that vary on widely different sales. The
161: amplitudes of these components can be characterized by the power
162: spectrum; the present determinations encompass the frequency
163: interval 0.01--0.8 $h$/Mpc,\footnote{As usual, $h$ is the present
164: Hubble parameter, measured in units of 100 km/sec/Mpc.} which
165: corresponds to the scale range from 8 to over 600 Mpc/$h$.
166: Fig.~\ref{fig:2dfpk} shows the power spectrum for the 2dF galaxy
167: redshift survey (2dFGRS).
168:
169:
170:
171: As cosmological densities have many scales and widely varying
172: amplitudes, density restoration should be adaptive. Different
173: methods exist to adaptively smooth point distributions to estimate
174: from them the underlying density field. \citet{schaap00} have
175: introduced the Delaunay Tessellation Field Estimator (DTFS) which
176: adapts itself to the point configuration even when anisotropies are
177: present. The method starts by considering the Delaunay tessellation
178: of the point process, then we can estimate the density at those
179: points using the contiguous Voronoi cells, and finally, we should
180: interpolate to obtain the density in the whole volume. Intricate
181: point patterns have been successfully smoothed using this method
182: and applications to particle
183: hydrodynamics provide good performance \citep{pelupessy03}.
184: \citet{ascasibar05} have recently introduced a novel technique based
185: on a different partition of the embedding space. These authors use
186: multidimensional binary trees to make the partition and latter
187: apply adaptive kernels within the resulting cells.
188: Finally, it is well known that wavelets provide a
189: localized (compact-kernel) adaptive restoration method
190: \citep{starck:book02}. We have applied, in a previous paper
191: \citep{mart05}, a wavelet based denoising technique to the 2dFGRS.
192: As a result we found that the morphology
193: of the galaxy density distribution in the survey volume does not
194: follow a Gaussian pattern, in contrast to the usual results in which
195: deviations of Gaussianity are not clearly detected (see,
196: e.g., \citet{hoyle02} for the 2dFGRS and \citet{park05} for the
197: Sloan Digital Sky Survey (SDSS)).
198:
199: \begin{figure}
200: \centering
201: \resizebox{.48\textwidth}{!}{\includegraphics*{2dfpk.eps}}
202: \caption{The matter density power spectrum for the 2dF
203: GRS. Data courtesy of W.~Percival and the 2dF GRS team.
204: \label{fig:2dfpk}}
205: \end{figure}
206:
207: By the way, adaptive density restoration methods are probably the
208: best for calculating partial Minkowski functionals, to describe the
209: morphology of single large-scale density enhancements (superclusters;
210: see, e.g., \cite{shandarin04}). Partial functionals can be used to characterize
211: the inner structure (clumpiness) and shapes (via shapefinders
212: \citep{sahni98}) of superclusters. As Minkowski functionals are
213: additive, partial functionals can be, in principle, combined to
214: obtain global Miknkowski functionals for the whole catalogue volume.
215: But if we want to check for non-Gaussianity, direct calculation of
216: global Minkowski functionals is more simple and straightforward.
217: When combining partial functionals, estimating the mean densities
218: and volume distributions for the full sample is a difficult
219: problem.
220:
221: Now, although a single adaptively found density distribution
222: represents the cosmological density field better, it could not be
223: the best tool for comparing theories with observations. Theories of
224: evolution of structure predict that Gaussianity of the original
225: density distribution is distorted during evolution, and this
226: distortion is scale-dependent. According to the present paradigm,
227: evolution of structure should proceed with different pace at
228: different scales. At smaller scales signatures of gravitational
229: dynamics should be seen, and traces of initial conditions could be
230: discovered at larger scales. In a single density field, containing
231: contributions from all scales, these effects are mixed. Thus, a
232: natural way to study cosmological density fields is the multi-scale
233: approach, scale by scale. This has been done in the past by using a
234: series of kernels of different width \citep{park05}, but this method
235: retains considerable low-frequency overlap. A better way is to
236: decompose the density field into different frequency (scale)
237: subbands, and to study each subband separately.
238:
239: The simplest idea of separation of scales by using different Fourier
240: modes does not work well, at least for morphological studies
241: (studies of shapes and texture). Describing texture requires
242: knowledge of positions, but Fourier modes do not have positions,
243: their position is the whole sample space. A similar weakness, to a
244: smaller extent, is shared by discrete orthogonal wavelet expansions
245: -- their localisation properties are better, but vary with scale,
246: and large-scale modes remain badly localised.
247:
248: This leads to the conclusion that the natural candidates for scale
249: separation are shift-invariant wavelet systems, where wavelet amplitudes
250: of all scales are calculated for each point of the coordinate grid. These wavelet
251: decompositions are redundant -- each subband has the same data
252: volume as the original data.
253:
254: For such a scheme, direct calculation of
255: low-frequency subbands would require convolution with wide wavelet
256: profiles, that could be numerically expensive. A way out is the
257: \emph{\`a trous} (with holes) trick, where convolution kernels for
258: the next dyadic scale are obtained by inserting zeros between the
259: elements of the original kernel. In this way, the total number of
260: non-zero elements is the same for all kernels, and all wavelet
261: transforms are equally fast.
262:
263: Now, as usual with wavelets, there is considerable freedom in
264: choosing the wavelet kernel. A particularly useful choise is the
265: wavelet based on a $B_3$ spline scaling function (see Appendix~A).
266: In this case, the original data can be reconstructed as a simnple
267: sum of subbands, without extra weights, so the subband decomposition
268: is the most natural.
269:
270: In the present paper we combine the \`a trous representation of
271: density fields with a grid-based algorithm to calculate the
272: Minkowski functionals (MF for short), and apply it to the
273: \mbox{2dFGRS} data.
274:
275:
276: Section~\ref{sec:morph} describes how to compute the MF for complex
277: data volumes and how to extend such an approach in a multiscale
278: framework using the wavelet transform. Section~\ref{sec:data}
279: describes our observational data while section~\ref{sec:gauss}
280: evaluates the multiscale MF on Gaussian random field realizations
281: for which the analytical results are known.
282: Section~\ref{sec:datamorph}
283: presents our results for the 2dFGRS data. These results are compared
284: with the multiscale MF calculated for 22 mock surveys and about 100
285: Monte-Carlo simulations of Gaussian random fields. We list the
286: conclusions in section~\ref{sec:concl}.
287:
288: \section{Morphological analysis}
289: \label{sec:morph}
290:
291: \subsection{Definition}
292: An elegant description of morphological characteristics of density
293: fields is given by Minkowski functionals \citep{cf:mecke94}. These
294: functionals provide a complete family of morphological measures. In
295: fact all additive, motion invariant and conditionally continuous
296: \footnote{The functionals are required to be continuous only for
297: compact convex sets; we can always represent any hypersurface as
298: unions of such sets.} functionals defined for any hypersurface are
299: linear combinations of its Minkowski functionals.
300:
301: The Minkowski functionals describe the morphology of iso-density
302: surfaces \citep{minkowski,tomita}, and depend thus on the specific
303: density level (see \citet{sheth05} for a recent review). Of course,
304: when the original data are galaxy positions, the procedure chosen to
305: calculate densities (smoothing) will also determine the result
306: \citep{mart05}. Generally, convolution of the data with a Gaussian
307: kernel is applied to obtain a continuous density field from the
308: point distribution. An alternative approach starts from the point
309: field, decorating the points with spheres of the same radius, and
310: studying the morphology of the resulting surface
311: \citep{ker96var,ker2}. This approach does not refer to a density; we
312: cannot use that for the present study.
313:
314: The Minkowski functionals are defined as follows. Consider an
315: excursion set $F_{\phi_0}$ of a field $\phi(\mathbf{x})$ in 3-D (the
316: set of all points where $\phi(\mathbf{x}\ge\phi_0$). Then, the first
317: Minkowski functional (the volume functional) is the volume of the
318: excursion set:
319: \[
320: V_0(\phi_0)=\int_{F_{\phi_0}}d^3x.
321: \]
322: The second MF is proportional to the surface area
323: of the boundary $\delta F_\phi$ of the excursion set:
324: \[
325: V_1(\phi_0)=\frac16\int_{\delta F_{\phi_0}}dS(\mathbf{x}).
326: \]
327: The third MF is proportional to the integrated mean curvature
328: of the boundary:
329: \[
330: V_2(\phi_0)=\frac1{6\pi}\int_{\delta F_{\phi_0}}
331: \left(\frac1{R_1(\mathbf{x})}+\frac1{R_2(\mathbf{x})}\right)dS(\mathbf{x}),
332: \]
333: where $R_1$ and $R_2$ are the principal curvatures of the boundary.
334: The fourth Minkowski functional is proportional to the integrated
335: Gaussian curvature (the Euler characteristic) of the boundary:
336: \[
337: V_3(\phi_0)=\frac1{4\pi}\int_{\delta F_{\phi_0}}
338: \frac1{R_1(\mathbf{x})R_2(\mathbf{x})}dS(\mathbf{x}).
339: \]
340: The last MF is simply related to other known morphological
341: quantities
342: \[
343: V_3=\chi=\frac12(1-G),
344: \]
345: where $\chi$ is the Euler characteristic and $G$ is the
346: topological genus, widely used in the past study of cosmological
347: density distributions. The
348: functional $V_3$ is a bit more comfortable to use -- it is additive,
349: while $G$ is not, and it gives just twice the number of isolated
350: balls (or holes).
351: Instead of the functionals, their spatial densities $V_i$ are
352: frequently used:
353: \[
354: v_i(f)=V_i(f)/V, \quad i=0,\dots,3,
355: \]
356: where $V$ is the total sample volume. The densities allow us to
357: compare the morphology of different data samples.
358:
359: The original argument of the functionals, the density level
360: $\rho_0$, can have different amplitudes for different fields, and
361: the functionals are difficult to compare. Because of that,
362: normalised arguments are usually used; the simplest one is the
363: volume fraction $f_v$, the ratio of the volume of the excursion set
364: to the total volume of the region where the density is defined.
365: Another, similar argument is the mass ratio $f_m$, which is very
366: useful for real, positive density fields, but is cumbersome to apply
367: for realizations of Gaussian fields, where the density may be
368: negative. The most widely used argument is the Gaussianized volume
369: fraction $\nu$, defined as \beq \label{nu}
370: f_v=\frac1{\sqrt{2\pi}}\int_\nu^\infty\exp(-t^2/2)\,dt. \eeq For a
371: Gaussian random field, $\nu$ is the density deviation from the mean,
372: divided by the standard deviation. This argument was introduced
373: already by \citet{gott86}), in order to eliminate the first trivial
374: effect of gravitational clustering, the deviation of the 1-point pdf
375: from the (supposedly) Gaussian initial pdf. Notice that using this
376: argument, the first Minkowski functional is trivially Gaussian by
377: definition.
378:
379: All the Minkowski functionals have analytic expressions for
380: iso-density slices of realizations of Gaussian random fields.
381: For three-dimensional space they are \citep{tomita}:
382: \begin{eqnarray}
383: \label{gaussv}
384: v_0&=&\frac12-\frac12\Phi\left(\frac{\nu}{\sqrt2}\right),\\
385: v_1&=&\frac23\frac{\lambda}{\sqrt{2\pi}}\exp\left(-\frac{\nu^2}2\right),\\
386: v_2&=&\frac23\frac{\lambda^2}{\sqrt{2\pi}}\nu\exp\left(-\frac{\nu^2}2\right),\\
387: v_3&=&\frac{\lambda^3}{\sqrt{2\pi}}(\nu^2-1)\exp\left(-\frac{\nu^2}2\right),
388: \end{eqnarray}
389: where $\Phi(\cdot)$ is the Gaussian error integral, and $\lambda$
390: is determined by the correlation function $\xi(r)$ of the field:
391: \beq
392: \label{lambda}
393: \lambda^2=\frac1{2\pi}\frac{\displaystyle\xi''(0)}{\displaystyle\xi(0)}.
394: \eeq
395:
396:
397: \subsection{Numerical algorithms}
398:
399: Several algorithms are used to calculate the Minkowski functionals
400: for a given density field and a given density threshold. We can
401: either try to follow exactly the geometry of the iso-density
402: surface, e.g., using triangulation \citep{surfgen}, or to
403: approximate the excursion set on a simple cubic lattice. The
404: algorithm that was proposed first by \citet{gott86}, uses a
405: decomposition of the field into filled and empty cells, and another
406: popular algorithm \citep{coles96} uses a grid-valued density
407: distribution. The lattice-based algorithms are simpler and faster,
408: but not as accurate as the triangulation codes.
409:
410: We use a simple grid-based algorithm, that makes use of integral
411: geometry (Crofton's intersection formula, see \citet{jens97}). We
412: find the density thresholds for given filling fractions by sorting
413: the grid densities, first. Vertices with higher densities than the
414: threshold form the excursion set. This set is characterised by its
415: basic sets of different dimensions -- points (vertices), edges
416: formed by two neighbouring points, squares (faces) formed by four
417: edges, and cubes formed by six faces. The algorithm counts the
418: numbers of elements of all basic sets, and finds the values of the
419: Minkowski functionals as
420: \begin{eqnarray}
421: \label{crofton}
422: V_0(f)&=&a^3N_3,\nonumber\\
423: V_1(f)&=&a^2\left(\frac29N_2(f)-\frac23N_3(f)\right),\nonumber\\
424: V_2(f)&=&a\left(\frac29N_1(f)-\frac49N_2(f)+\frac23N_3(f)\right),\nonumber\\
425: V_3(f)&=&N_0(f)-N_1(f)+N_2(f)-N_3(f),
426: \end{eqnarray}
427: where $a$ is the grid step, $f$ is the filling factor, $N_0$ is the
428: number of vertices, $N_1$ is the number of edges, $N_2$ is the
429: number of squares (faces), and $N_3$ is the number of basic cubes in
430: the excursion set for a given filling factor (density threshold).
431: The formula (\ref{crofton}) was first used in
432: cosmological studies by \citet{coles96}.
433:
434: \subsection{Biases}
435: The algorithm described above is simple to program, and is very fast,
436: allowing the use of Monte-Carlo simulations for error estimation.
437:
438: However, it suffers from discreteness errors, which are not large,
439: but annoying, nevertheless. An example of that is given in
440: Fig.~\ref{fig:v1shift}, where we show the $V_1$ functional,
441: calculated by the above recipes for a periodic realization of a
442: Gaussian field (the dashed line). As we see, it has a constant
443: shift in $\nu$ over the whole range. This shift is due to the fact
444: that when we approximate iso-density surfaces by a discrete grid,
445: the vertices that compose the surface lie in a range of densities
446: starting from the nominal one. This effect can easily be calculated
447: because this bias will show up as a constant shift in $\nu$ for a
448: Gaussian density field, as observed. Other functionals ($V_2$ and
449: $V_3$) suffer similar shifts, with smaller amplitudes, and these are
450: not easy to explain.
451:
452: \begin{figure}
453: \centering
454: \resizebox{.48\textwidth}{!}{\includegraphics*{v1shift.eps}}
455: \caption{The $V_1$ functional for a realization of a Gaussian
456: random density field in a periodic $256^3$ cube (upper panel).
457: Full line shows the theoretical prediction for this realization,
458: dotted line -- the standard one-excursion-set estimate,
459: and dashed line -- the average of the functional over two excursion sets,
460: The lower panel shows the difference between the estimates and
461: the theoretical prediction; the lines encode the estimates as above.
462: \label{fig:v1shift}}
463: \end{figure}
464:
465: There is, fortunately, another and simple possibility to fight these
466: errors. The standard way is to approximate an iso-density surface by
467: the collection of vertices that have densities $\rho\ge\rho_l$,
468: where $\rho_l$ is the threshold density. But another surface, formed
469: by the vertices with $\rho<\rho_l$, is as good an approximation to
470: the iso-density surface as the first one. Thus, the natural way to
471: calculate the Minkowski functionals is to run the algorithm twice,
472: swapping the marks for the excursion set, and averaging the values
473: of the functionals obtained. The last step is justified, as the
474: Minkowski functionals are additive. The averaging rules are:
475: \begin{eqnarray*}
476: V_0&=&\left(V_0^{(1)}+V_{\mbox{\scriptsize tot}}-V_0^{(2)}\right)/2,\\
477: V_1&=&\left(V_1^{(1)}+V_1^{(2)}\right)/2,\\
478: V_2&=&\left(V_2^{(1)}-V_2^{(2)}\right)/2,\\
479: V_3&=&\left(V_3^{(1)}+V_3^{(2)}\right)/2,\\
480: \end{eqnarray*}
481: Here the upper indices $(1)$ and $(2)$ denote the original and
482: complementary excursion sets, respectively, and
483: $V_{\mbox{\scriptsize tot}}$ is the total number of grid cubes in
484: the data brick. The minus sign in the formula for the third
485: functional ($V_2$) accounts for the fact that the curvature of the
486: second surface is opposite to that of the first one.
487:
488: The $V_1$ functional calculated this way is shown in
489: Fig.~\ref{fig:v1shift} by the full line; we can compare it with the
490: theoretical prediction for Gaussian fields ((\ref{gaussv}), the
491: dotted line). The coincidence of the two curves is very good; the
492: only slight deviation is at $\nu\approx0$, where the Gaussian
493: surface is more complex. We have to stress that the Gaussian curve
494: is not a fit; the parameter $\lambda$ that determines the amplitude
495: of the curve was found directly from the data, using the relations
496: $\xi(0)=\langle\rho^2\rangle$ and
497: $\xi''(0)=\langle\rho_{,i}^2\rangle$, where $\xi(r)$ is the
498: correlation function and $\rho_{,i}$ is the derivative of density at
499: a grid vertex in one of the coordinate directions. The good match of
500: these curves shows also that the Gaussian realization is good, which
501: is not simple to model. The averaging works as well for the two
502: other functionals; this is shown in Fig.~\ref{fig:v23shift}. There
503: are slight deviations from the theoretical curve for $V_2$ around
504: $\nu\approx1$ and for $V_3$ at $\nu\approx0$; these may be intrinsic
505: to the particular realization, as the number of 'resolution details'
506: diminishes when the order of the functional increases.
507:
508: \begin{figure}
509: \centering
510: \resizebox{.48\textwidth}{!}{\includegraphics*{v2shift.eps}}\\
511: \resizebox{.48\textwidth}{!}{\includegraphics*{v3shift.eps}}\\
512: \caption{The $V_2$ (upper panels) and $V_3$ (lower panels)
513: functionals for a realization of a Gaussian
514: random density field in a periodic $256^3$ cube.
515: In the larger panels
516: full lines show the theoretical predictions for this realization,
517: dotted lines -- the standard one-excursion-set estimates,
518: and dashed lines -- the averages of the functionals over two excursion sets,
519: The smaller panels show the differences between the estimates and
520: the theoretical predictions; the lines encode the estimates as above.
521: \label{fig:v23shift}}
522: \end{figure}
523:
524: We also see that the higher the order, the closer are the
525: one- and two- excursion set estimates. So, even if we are interested
526: only in the topology of the density iso-surfaces, we should
527: correct for the border effects, all Minkowski functionals are used,
528: and it is important that they were unbiased.
529:
530: There is a natural restriction on the grid steps -- the grid has to
531: be fine enough to resolve the details of the density field. The
532: previous figures (Figs.~\ref{fig:v1shift}, \ref{fig:v23shift}) show
533: the Minkowski functionals obtained for the case of the Gaussian
534: field smoothed by a Gaussian filter with $\sigma=3$ grid steps,
535: being the smoothing radius $R\approx2\sigma=6$.
536:
537: \subsection{Border corrections\label{subsec:bound}}
538: As we have seen above, we can obtain good estimates of the Minkowski
539: functionals for periodic fields. The real data, however, is always
540: spatially limited, and the limiting surfaces cut the iso-density
541: surface. An extremely valuable property of Minkowski functionals is
542: that such cuts can be corrected for. Let us assume that the data
543: region (window or mask) is big enough relative to the typical size
544: of details, so that one can consider the field inside the mask
545: homogeneous and isotropic. For this case, \citet{ker96var} show that
546: the observed Minkowski functionals for the masked iso-surface
547: $M_i(D\cap W)$ can be expressed as a combination of the true
548: functionals $M_i(D)$ and those of the mask $M_i(W)$: \beq
549: \label{kinform} M_i(D\cap W)=\frac1{\cal V}\sum_{j=0}^i
550: \left(\begin{array}{c}i\\j\end{array}\right)
551: M_j(D)M_{i-j}(W),
552: \eeq where $\cal V$ is the total volume inside the mask. Note that
553: the functionals $M_i$ differ from the usual $V_i$ by normalisation.
554: \citet{ker96var} derive the relation (\ref{kinform}) for a
555: collection of balls. Here we have applied it to iso-density surfaces
556: and for the true values of the functionals, we get \beq
557: \label{mcorr} \frac{M_i(D)}{\cal V}=\frac{M_i(D\cap W)}{M_0(W)}-
558: \sum_{j=0}^{i-1}\left(\begin{array}{c}i\\j\end{array}\right)
559: \frac{M_j(D)}{\cal V}\frac{M_{i-j}(W)}{M_0(W)}.
560: \eeq
561:
562: The relation between $M_i$-s and the usual $V_i$-s is
563: \[
564: M_i=\frac{\omega_{d-i}}{\omega_d}V_i,
565: \]
566: where $\omega_j$ is the volume of a $j$-dimensional unit ball,
567: and $d$ is the dimension of the space.
568: For Minkowski functionals in three-dimensional space,
569: the explicit relations are:
570: \[
571: M_0=V_0;\quad M_1=\frac34 V_1;\quad M_2=\frac3{2\pi}V_2;\quad M_3=\frac3{4\pi}V_3.
572: \]
573: Using (\ref{mcorr}) and replacing $M_i$-s by $V_i$, we arrive
574: at the following correction chain:
575: \begin{eqnarray}
576: \label{corrchain}
577: v_0(\nu)&=&\frac{V_0(\nu)}{V_0(W)}, \nonumber \\
578: v_1(\nu)&=&\frac{V_1(\nu)}{V_0(W)}-v_0(\nu)\frac{V_1(W)}{V_0(W)},\nonumber \\
579: v_2(\nu)&=&\frac{V_2(\nu)}{V_0(W)}-v_0(\nu)\frac{V_2(W)}{V_0(W)}
580: -\frac{3\pi}{4}v_1(\nu)\frac{V_1(W)}{V_0(W)},\nonumber \\
581: v_3(\nu)&=&\frac{V_3(\nu)}{V_0(W)}-v_0(\nu)\frac{V_3(W)}{V_0(W)}
582: -\frac92 v_1(\nu)\frac{V_2(W)}{V_0(W)} \nonumber \\
583: & &
584: -\frac92 v_2(\nu)\frac{V_1(W)}{V_0(W)}.\nonumber\\
585: \end{eqnarray}
586: Here $V_i(\nu)$ denote the observed (raw) values of Minkowski
587: functionals, and $v_i(\nu)$ denote the corrected densities.
588: \begin{figure*}
589: \centering
590: \resizebox{.32\textwidth}{!}{\includegraphics*{Ngsv1.eps}}
591: \resizebox{.32\textwidth}{!}{\includegraphics*{Ngsv2.eps}}
592: \resizebox{.32\textwidth}{!}{\includegraphics*{Ngsv3.eps}}
593: \caption{Demonstration of border corrections for complex
594: borders. The raw densities
595: of Minkowski functionals for the 2dFGRS NGP sample volume $252^3$,
596: cut from a periodic $256^2\times64$ realization of a Gaussian
597: random field (dotted lines)
598: are shown together with the border-corrected estimates (dashed lines)
599: and the estimates for the original brick (full lines).
600: Upper panels show the densities, and smaller lower panels
601: show the differences between the densities for the sample volume
602: and those for the brick (for both the raw and corrected cases,
603: and the same line types are used as in the upper
604: panels).
605: The densities of the second MF $v_1$ are shown in the left panels, the
606: densities of the third MF $v_2$ -- in the middle panels, and the
607: densities of the fourth MF $v_3$ -- in the right panels.
608: \label{fig:Nmaskv123}}
609: \end{figure*}
610:
611: We tested these corrections with our original Gaussian realization,
612: masked at all faces. The correction for the second Minkowski
613: functional $v_1$ is practically perfect. The corrected version of
614: $v_2$ is also close to the original for all the argument range, and
615: only a little higher than the original. The higher the order of the
616: functional, the more difficult it is to correct for the borders, as
617: small errors from the lower orders accumulate. The discrepancy with
618: the corrected version of $v_3$ and the original estimates is the
619: largest amongst the three densities, but it balances well the
620: amplitudes of the maxima, and is only a little lower than the
621: original for low densities ($\nu\approx1.5$). There are practically
622: no differences from the original at the high-density end.
623:
624: A note on the use of masks in practice: we ensure that there is at
625: least one-vertex thick mask layer around our data brick. This allows
626: us to assume periodic borders for the brick itself.
627: And there is also another way to use the mask, ignoring the vertices in
628: the mask, not building any elements from the vertices in the data
629: region to the mask vertices. Then we do not have to apply the
630: correction chain (\ref{corrchain}) and do not have to build the
631: basic sets in the mask region. The latter fact makes the
632: algorithm about twice as fast for the 2dF data (the data region
633: occupies only a fraction of the encompassing brick). We compared this
634: version with the border-corrected algorithm described above
635: (see Appendix~B)
636: and found that it gives slightly worse results for Gaussian
637: realizations, so we dropped it. The present algorithm is fast
638: enough, taking 12 seconds for a iso-level for a $256^3$ grid
639: on a laptop with the Intel Celeron 1500 MHz processor.
640:
641: As the data masks are complex (see Fig.~\ref{fig:masks}), we should
642: test the border corrections for real masks, too.
643: Fig.~\ref{fig:Nmaskv123} shows the effect of border corrections as
644: used with the data mask for the 2dF NGC sample volume (see
645: section~\ref{sec:data}). As the corrections give densities of
646: functionals, we will show the densities from this point on. The
647: density field in this volume was generated by simulating a Gaussian
648: random field for a $256\times256\times64$ periodic brick, combining
649: these bricks to cover the sample extent, and masking this
650: realization with the Northern data mask. Smaller bricks were
651: combined because the spatial extent of the data was too large for
652: the available core memory to generate a single FFT brick to cover
653: it; as the brick is periodic, the realization remains Gaussian. We
654: show the raw densities of the Minkowski functionals as dashed lines,
655: the corrected versions with full lines, and the densities for the
656: original brick with dotted lines.
657:
658:
659: We see that the density $v_1$ of the $V_1$ functional is restored
660: well, apart from a slight deviation near $\nu=0$. The density $v_2$
661: is also corrected well, only its maximum amplitude is slightly
662: smaller than that for the original brick. The restoration is almost
663: perfect for $v_3$, with the same small amplitude problem than for
664: the two other densities. The fact that the restoration works so well
665: is really surprising -- first, our mask is extremely complex, and,
666: secondly, our realization of the Gaussian random field is certainly
667: not exactly homogeneous and isotropic inside the sample volume. We
668: generate realizations of random fields in this work by the FFT
669: technique; these fields are homogeneous, but isotropic only for
670: small scales, not for scales comparable to the brick size.
671:
672: The importance of the good restoration of the Minkowski functional
673: means that when checking theoretical predictions, we can
674: directly compare observational results with the predictions,
675: and do not have to use costly Monte-Carlo simulations.
676:
677: \subsection{Multiscale Minkowski Functionals}
678:
679: A natural way to study cosmological fields is the multi-scale
680: approach, scale by scale. According to the present paradigm,
681: evolution of structure should proceed with different pace at
682: different scales. At smaller scales signatures of gravitational
683: dynamics should be seen, and traces of initial conditions could be
684: discovered at larger scales.
685:
686: The matter density in the universe is formed by perturbations at all
687: scales. In the beginning they all grow at a similar rate, but soon
688: this rate becomes scale-dependent, the smaller the scale, the faster
689: it will go nonlinear and non-Gaussian. Thus it is interesting to
690: decompose the density (and gravitational potential, and velocity)
691: field into different scales and check their Gaussianity (and other
692: interesting characteristics).
693:
694: Using the wavelet transform as it is described in Appendix~A, we
695: obtain a set of wavelet scales ${\cal W} = {w_0, ..., w_J,
696: c_{J+1}}$, and each scale $w_j(x,y,z)$ corresponds to the
697: convolution product of the observed galaxies with a wavelet function
698: $\psi_j$, where $\psi_j(x,y,z) = \psi(
699: \frac{x}{2^j},\frac{y}{2^j},\frac{z}{2^j})$ and $\psi$ is the
700: analyzing wavelet function described in Appendix~A. Now, we can
701: apply the MF calculation at each scale independently, and we get
702: four MF values per scale using Eq.~\ref{crofton}. The set $(V_{j,0},
703: V_{j,1}, V_{j,2}, V_{j,3})$ will denote the MF at scale $j$. Note
704: that in this framework, we do not have to convolve the data anymore
705: with a Gaussian kernel, avoiding the delicate choice of the size of
706: the kernel bandwidth.
707:
708:
709: \section{The data}
710: \label{sec:data} There exist two large-volume galaxy redshift
711: surveys at the moment, the Two Degree Field Galaxy Redshift Survey
712: (2dFGRS) \citep{2df} and the Sloan Digital Sky Survey (SDSS)
713: \citep{sdss}. The 2dFGRS is completed; although the SDSS is not, its
714: data volume has already surpassed that of the 2dFGRS. We shall use
715: in our paper the 2dFGRS dataset; it is easier to handle, and we make
716: use of the mock catalogues created to estimate the cosmic variance
717: of the data.
718:
719: The galaxies of the 2dFGRS have been selected from an earlier
720: photometric APM survey \citep{maddox96} and its extensions. The
721: survey covers about 2000 deg$^2$ in the sky and consists of two
722: separate regions, one in the North Galactic Cap (NGC) and the other
723: in the South Galactic Cap (SGC), plus a number of small randomly
724: located fields; we do not use the latter. The total number of
725: galaxies in the survey is about 250,000. The depth of the survey is
726: determined by its limiting apparent magnitude, which was chosen to
727: be \mbox{$b_J=19.45$}. Caused by varying observing conditions,
728: however, this limit depends on the sky coordinates, varying almost
729: the full magnitude. Another cause of non-uniformity of the catalogue
730: is its spectroscopic incompleteness -- as the fibres used to direct
731: the light from a galaxy image in the focal plane to the spectrograph
732: have a finite size, a number of galaxies in close pairs were not
733: observed for redshifts. However, these corrections can be estimated;
734: the 2dFGRS team has made public the programs that calculate the
735: completeness factors and magnitude limit, given a line-of-sight
736: direction.
737:
738: The 2dFGRS survey, as all redshift surveys, is magnitude-limited.
739: This means that the density of observed galaxies decreases with
740: distance; at large distances only intrinsically brighter galaxies
741: can be seen. For certain statistical studies (luminosity functions,
742: correlation functions, power spectra) this decrease can be corrected
743: for. For texture studies there are yet no appropriate correction
744: methods, and maybe, these do not exist, as the scales of the details
745: that can be resolved are inevitably different, and small-scale
746: information is certainly lost at large distances. The usual approach
747: is to use volume-limited subsamples extracted from the survey. In
748: order to create such a sample, one chooses absolute magnitude
749: limits, and retains only the galaxies with absolute magnitudes
750: between these limits. This discards most of the data, but assures
751: that the spatial resolution is the same throughout the survey volume
752: (taking also account of the possible luminosity evolution and
753: K-correction).
754:
755: The 2dFGRS team has created such catalogues and used them to
756: study higher-order correlation functions
757: \citep{norberg02,croton1,croton2}. They kindly made these catalogues
758: available to us, and these constitute our main data. These
759: catalogues span one magnitude each, from $M= -17+5\log10(h)$ until
760: $M= -21+5\log10(h)$. The catalogues for least bright galaxies span
761: small spatial volumes, and those for the brightest galaxies are
762: sparsely populated. Thus we chose for our work the catalogue
763: spanning the magnitude range $-20\le M-5\log10(h)\le-19$, this is
764: the most informative. This has been also the conclusion of
765: \citet{croton2}. We shall call this sample 2dF19. This sample, as
766: all 2dF volume-limited samples, consists of two spatially distinct
767: subsamples, one in the North Galactic Cap region (2df19N) and
768: another in the South Galactic Cap region (2df19S). The sample lies
769: between 61.1~Mpc/$h$ and 375.6~Mpc/$h$; the general features of the
770: subsamples are listed in Table~\ref{tab:cats}.
771:
772: \begin{table*}
773: \centering \caption{The 2dF volume-limited catalogues used.
774: \label{tab:cats}}
775: \medskip
776:
777: \begin{tabular}{|c|r|rr|rr|r|r|}\hline
778: sample&galaxies&\multicolumn{2}{c|}{ra limits (deg)}&
779: \multicolumn{2}{c|}{dec limits}& \multicolumn{1}{c|}{Vol ($10^6$}&
780: \multicolumn{1}{c|}{dmean}\\
781: &&\multicolumn{2}{c|}{(deg)}& \multicolumn{2}{c|}{(deg)}&
782: \multicolumn{1}{c|}{Mpc$^3h^{-3}$)}&
783: \multicolumn{1}{c|}{(Mpc/$h$)}\\
784: \hline
785: 2dF19N&19080&147.0&223.0&$-$6.4&2.6&2.75&5.24\\
786: 2dF19S&25633&$-$35.5&55.2&$-$37.6&$-$22.4&4.43&5.57\\
787: \hline
788: \end{tabular}
789: \end{table*}
790:
791: In a previous paper on the morphology of the 2dFGRS \citep{mart05}
792: we extracted bricks from the data to avoid the influence of border
793: effects. This forced us to use only a fraction of the volume-limited
794: samples. This time we tried to use all the available data, and
795: succeeded with that.
796: \begin{figure}
797: \centering
798: \resizebox{.48\textwidth}{!}{\includegraphics*{uno.ps}}\\
799: \resizebox{.48\textwidth}{!}{\includegraphics*{dos.ps}}\\
800: \caption{Survey masks for the 2dFGRS volume-limited sample 2dF19.
801: Upper panel -- the Northern subsample, lower panel -- the Southern
802: subsample. Spatial orientations are chosen to better visualise the
803: volumes.\label{fig:masks}}
804: \end{figure}
805:
806: The 2dFGRS catalogue is composed of measurements in a large number
807: of circular patches in the sky, and its footprint in the sky is
808: relatively complex (see the survey web-page
809: \texttt{http://www.mso.anu.au/2dFGRS}). Furthermore, due to the
810: variations in the final magnitude limit used, the catalogue depth is
811: also a function of the direction. In order to use all the data for
812: wavelet and morphological analysis, we had to create a spatial mask,
813: separating the sample volume from regions outside. We did it by
814: creating first a spatial grid for a brick that surrounded the
815: observed catalogue volume, calculated the sky coordinates for all
816: vertices of the brick, and used then the software provided by the
817: 2dFGRS team to find the correction factors for these directions. The
818: completeness factor told us if the direction was inside the sample
819: footpath. If it was, we found the apparent magnitude of the
820: brightest galaxies (with $M=M_{\mbox{\scriptsize min}}$) of our
821: sample for that distance and checked, if it was lower than the
822: sample limit. If it was, the grid point was included in the mask.
823: For the last comparison we had to change from the comoving grid
824: distance to the luminosity distance. We assumed the $\Omega_M=0.3$,
825: $\Omega_\Lambda=0.7$ cosmology for that and interpolated a tabulated
826: relation between the two distances. As explained in
827: \citet{norberg02}, the 2dF volume-limited samples were built using
828: the $k+e$-correction as dependent on the spectral type of a galaxy.
829: We do not have such a quantity for the mask, so we tuned a little
830: the bright absolute magnitude limit, checking that the mask should
831: extend as far as the galaxy sample. For that, we had to increase the
832: effective bright absolute magnitude limit by $\Delta M= 0.25$. The
833: nearby regions of the mask were cut off at the nearest distance
834: limit for the observational sample.
835:
836: The original survey mask in the sky includes holes around bright
837: stars; these holes generate narrow tunnels through our spatial mask.
838: As the masks have a complex geometry, we did not want to add
839: discreteness effects due to resolving such tunnels. We filled them
840: in by counting the number of neighbours in a $3^3$ cube around
841: non-mask points and, if the neighbour number was larger than a
842: chosen limit, assigning the points to the mask. We chose the
843: required number $n$ of neighbours to avoid filling in at flat mask
844: borders ($n=9$ is enough for that) and iterated the procedure until
845: the tunnels disappeared. This was determined by visual checks (using
846: the 'ds9' fits file viewer \citep{ds9}).
847:
848: We show the 3-D views of our masks in Fig.~\ref{fig:masks}. The
849: mask volumes are, in general, relatively thin curved slices with
850: heavily corrugated outer walls. These corrugations are caused by the
851: unobserved survey fields. Also, the outer edges of the mask are
852: uneven, due to the variations in the survey magnitude limit. One 3-D
853: view does not give a good impression of the mask; the slices we
854: shall show below will complement these.
855:
856:
857: As mentioned in Appendix~A, in order to apply the wavelet
858: convolution cascade, the initial density on the grid should be
859: extirpolated by the chosen scaling function. We chose our initial
860: grid step as 1~Mpc/$h$, and used the $B^{(3)}_3$ kernel for
861: extirpolation. In order to have a better scale coverage, we repeated
862: the analysis, using the grid step $\sqrt2$~Mpc/$h$. The smoothing
863: scale (the spatial extent of the kernel) is 4 grid units, the
864: smoothing radius corresponds to 2 units. When calculating the
865: densities, we used the spectroscopic completeness corrections
866: $c_{\mbox{i\scriptsize sp}}$, included in the 2dFGRS volume-limited
867: catalogues, and weighted the galaxies by the factor
868: $w=1/c_{\mbox{i\scriptsize sp}}$. Most of the weights are close to
869: unity, but a few of them are large. In order not to 'overweight'
870: these galaxies, we fixed the maximum weight level as $w=2$. The same
871: procedure was chosen by \citet{croton1}.
872:
873:
874:
875:
876: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
877:
878:
879: \section{Gaussian fields}
880: \label{sec:gauss}
881:
882: The filters used to perform wavelet expansion
883: are linear, and thus should keep the morphological
884: structure of Gaussian fields; the Minkowski functionals
885: should be Gaussian for any wavelet order.
886: This is certainly true for periodic densities,
887: but for densities restricted to finite volumes
888: the boundary conditions can introduce correlations.
889: The most popular boundary condition -- reflection at
890: the boundary -- will keep the density field mostly
891: Gaussian for brick masks. Our adopted zero boundary condition
892: will certainly work destroying Gaussianity, as the
893: random field which is zero outside a given
894: volume and has finite values inside is certainly not Gaussian.
895:
896: \begin{figure*}
897: \centering
898: \resizebox{0.32\textwidth}{!}{\includegraphics*{mf1comp.eps}}
899: \resizebox{0.32\textwidth}{!}{\includegraphics*{mf2comp.eps}}
900: \resizebox{0.32\textwidth}{!}{\includegraphics*{mf3comp.eps}}
901: \caption{Densities of the Minkowski functionals (from the second
902: at the left until the fourth at the right)
903: for the same Gaussian realization for
904: the wavelet orders 3 and 4 (lower wavelet orders give higher
905: amplitudes). The case for the wavelets generated using the mirror
906: boundary conditions is shown by full lines, for zero boundary
907: conditions dotted lines are used. The amplitudes of the functionals
908: for the wavelet order 4 are rescaled by 2 for $v_2$ and by 4 for $v_3$
909: to show more details.
910: \label{fig:mfcomp}}
911: \end{figure*}
912:
913:
914:
915: \begin{figure*}
916: \centering
917: \fbox{\resizebox{0.25\textwidth}{!}{\includegraphics*{MNG2.ps}}}~
918: \fbox{\resizebox{0.25\textwidth}{!}{\includegraphics*{MNG1.ps}}}~
919: \fbox{\resizebox{0.25\textwidth}{!}{\includegraphics*{NG.ps}}}\\[4pt]
920: \fbox{\resizebox{0.25\textwidth}{!}{\includegraphics*{WNG2.ps}}}~
921: \fbox{\resizebox{0.25\textwidth}{!}{\includegraphics*{WNG1.ps}}}~
922: \fbox{\resizebox{0.25\textwidth}{!}{\includegraphics*{WNG0.ps}}}\\
923: \caption{Multi-scale decomposition of a realization of a Gaussian
924: random field in the 2df19N volume mask, for the $z=34$~Mpc/$h$
925: slice (the data and the first orders). The upper panel column shows
926: the scaling orders, with the original density at the right.
927: The lower panel shows the wavelet orders, with the lowest
928: order at the right.
929: \label{fig:Ngsbunch1}}
930: \end{figure*}
931:
932: \begin{figure*}
933: \centering
934: \newlength{\pushlen}
935: \setlength{\pushlen}{0.25\textwidth}
936: \addtolength{\pushlen}{2\fboxsep}
937: \addtolength{\pushlen}{2\fboxrule}
938: \fbox{\resizebox{0.25\textwidth}{!}{\includegraphics*{MNG5.ps}}}~
939: \fbox{\resizebox{0.25\textwidth}{!}{\includegraphics*{MNG4.ps}}}~
940: \fbox{\resizebox{0.25\textwidth}{!}{\includegraphics*{MNG3.ps}}}\\[4pt]
941: \parbox{\pushlen}{\hfill}~
942: \fbox{\resizebox{0.25\textwidth}{!}{\includegraphics*{WNG4.ps}}}~
943: \fbox{\resizebox{0.25\textwidth}{!}{\includegraphics*{WNG3.ps}}}\\
944: %\fbox{\resizebox{0.32\textwidth}{!}{\includegraphics*{WNG2.ps}}}\\
945: \caption{Multi-scale decomposition of of a Gaussian
946: random field in the 2df19N volume mask (continued).
947: Upper panel -- scaling orders, lower panel -- wavelet orders,
948: the highest orders at the left.
949: The last scaling solution shows
950: already strong effects of the boundary conditions.
951: \label{fig:Ngsbunch2}}
952: \end{figure*}
953:
954: We compare the effect of the mirror
955: and zero boundary conditions in Fig.~\ref{fig:mfcomp},
956: for high wavelet orders that are yet not dominated by noise,
957: calculated for a Gaussian realization in a $256^3$ cube
958: with the $\sigma=3$ Gaussian smoothing described above, and
959: masked at all borders by a layer two vertices thick
960: (thus the effective volume of the data cube is $252^3$)
961:
962: Fig.~\ref{fig:mfcomp} shows that the second Minkowski functional
963: is certainly better restored by the wavelets obtained by using
964: mirror boundary conditions.
965: These conditions lead to the functionals that
966: are symmetric about $\nu=0$, as it should be, while the functionals
967: obtained by applying zero boundary conditions display a shift towards smaller
968: values of $\nu$.
969: However, although our brick data
970: should prefer mirror boundary conditions, it is not easy to say which
971: boundary conditions give better estimates for the remaining two functionals.
972: Both cases have comparable errors (look at the amplitudes at extrema,
973: which should be equal). Hence, other considerations have to be used.
974: As our data mask is complex, with corrugated planes and sharp corners,
975: the mirror boundary conditions will amplify the corner densities and
976: propagate them inside, while the zero boundary conditions will
977: gradually remove the influence of the boundary data. Thus we have
978: selected zero boundary conditions for the present study; these are
979: probably natural for all observational samples.
980:
981:
982: \begin{figure}
983: \centering
984: \resizebox{0.48\textwidth}{!}{\includegraphics*{mf3wng.eps}}
985: \caption{The density $v_3$ of the fourth MF
986: for the wavelet decomposition of a realization of a Gaussian
987: random field in the 2dfN19 mask. The legends in the figure show
988: the wavelet order and the scaling factor (2W1 denotes the
989: functional for the wavelet order 1, multiplied by 2). The
990: legend NG denotes the original realization.
991: \label{fig:mf3wng}}
992: \end{figure}
993:
994: \begin{figure*}
995: \centering
996: \fbox{\resizebox{0.25\textwidth}{!}{\includegraphics*{smWN2.ps}}}~
997: \fbox{\resizebox{0.25\textwidth}{!}{\includegraphics*{smWN1.ps}}}~
998: \fbox{\resizebox{0.25\textwidth}{!}{\includegraphics*{2dfNsm.ps}}}\\[4pt]
999: \fbox{\resizebox{0.25\textwidth}{!}{\includegraphics*{smMN4.ps}}}~
1000: \fbox{\resizebox{0.25\textwidth}{!}{\includegraphics*{smWN4.ps}}}~
1001: \fbox{\resizebox{0.25\textwidth}{!}{\includegraphics*{smWN3.ps}}}\\
1002: \caption{Multi-scale decomposition of the 2df19N volume-limited sample,
1003: for the $z=39$~Mpc/$h$ slice, and for the grid step $\sqrt2$~Mpc/$h$.
1004: The original density is shown at top right,
1005: the last scaling order at bottom left, the wavelet orders in between.
1006: The wavelet orders increase from right to left and from top to bottom.
1007: The weakest gray level shows the sample mask.
1008: \label{fig:2df19Nbunch2}}
1009: \end{figure*}
1010:
1011:
1012: \begin{figure*}
1013: \centering
1014: \fbox{\resizebox{0.25\textwidth}{!}{\includegraphics*{smWS2.ps}}}~
1015: \fbox{\resizebox{0.25\textwidth}{!}{\includegraphics*{smWS1.ps}}}~
1016: \fbox{\resizebox{0.25\textwidth}{!}{\includegraphics*{2dfSsm.ps}}}\\[4pt]
1017: \fbox{\resizebox{0.25\textwidth}{!}{\includegraphics*{smMS4.ps}}}~
1018: \fbox{\resizebox{0.25\textwidth}{!}{\includegraphics*{smWS4.ps}}}~
1019: \fbox{\resizebox{0.25\textwidth}{!}{\includegraphics*{smWS3.ps}}}\\
1020: \caption{Multi-scale decomposition of the 2df19S volume-limited sample,
1021: for the $z=97$~Mpc/$h$ slice, and for the grid step $\sqrt2$~Mpc/$h$.
1022: The original density is shown at top right,
1023: the last scaling order at bottom left, the wavelet orders in between.
1024: The wavelet orders increase from right to left and from top to bottom.
1025: The weakest gray level shows the sample mask (in most cases; sometimes
1026: the mask is missing and sometimes the gray level is wider than the
1027: mask).
1028: \label{fig:2df19Sbunch2}}
1029: \end{figure*}
1030:
1031:
1032: We have seen that the wavelet components of the multi-scale
1033: decomposition of a realization of a Gaussian random field remain
1034: practically Gaussian for simple sample boundaries. This means that
1035: testing for Gaussianity is straightforward. However, we have to
1036: assess the boundary effect for our application, where the boundaries
1037: are extremely complex. We shall demonstrate it on the example of
1038: the Northern mask. For that, we generated a realization of a
1039: Gaussian random field for a volume encompassing the mask, as
1040: described in section~\ref{sec:data}, and masked out the region outside
1041: the NGP data volume. As we want to see the effects that could show
1042: up in the data, we used the standard dark matter power spectrum for
1043: the $\Lambda$CDM cosmology \citep{klypin97}, for the cosmological
1044: parameters $\Omega_0=0.3, \Omega_{\Lambda}=0.7,
1045: \Omega_{\mbox{\scriptsize bar}}=0.026, h=0.7$ (this is pretty close
1046: to the standard 'concordance' power spectrum), generated the
1047: realization on a grid with the step of 1~Mpc/$h$, and smoothed the
1048: field with a Gaussian of $\sigma=2$~Mpc/$h$. The original density
1049: distribution, the scaling distributions and the wavelets are shown
1050: in Figs.~\ref{fig:Ngsbunch1} and \ref{fig:Ngsbunch2} for a slice at
1051: $z=34$~Mpc/$h$, at about the middle of the sample volume.
1052:
1053:
1054: The Minkowski functional $V_3$ for the wavelets is shown in
1055: Fig.~\ref{fig:mf3wng}. In order not to overcrowd the figure, we do
1056: not show the theoretical predictions. The functional has been
1057: rescaled to show all functionals together in a single diagram and
1058: scaling factors are shown in the labels.
1059:
1060: As we see, the lower the wavelet order, the larger values of $V_3$
1061: are obtained (this is also true for the other functionals);
1062: this is expected, as higher orders represent
1063: increasingly smoother details of the field.
1064: The values of the functionals for the zero-order wavelet are
1065: always higher than those for the full field, as it includes
1066: only the high-resolution details that the iso-levels have
1067: to follow.
1068: Also, we have seen that the lower the order of the functional,
1069: the smaller are the distortions from Gaussianity.
1070:
1071: The distortions of the third Minkowski functional are the largest.
1072: The functional for the zeroth-order wavelet (curve W0 in
1073: Fig.~\ref{fig:mf3wng}) shows argument compression, the result of
1074: insufficient spatial resolution of the smallest details of the
1075: field. The functional for the first-order wavelet is close to
1076: Gaussian, as is the functional for the total realization, but the
1077: second-order curve 8W2 shows strong distortions, due to a small
1078: number of independent resolution elements in the volume, and to the
1079: small height of the slice. The characteristic volume of these
1080: elements is $(4 \times 2^2)^3=4096$~Mpc$^3/h^3$, and their number is
1081: about 670. The functional for the third wavelet order is already
1082: completely dominated by noise.
1083:
1084: So, we can estimate Minkowski functionals with a high precision, and
1085: the border corrections work well. The most difficult part at the
1086: moment is the scale separation in observed samples. The sample
1087: geometries are yet slice-like, limiting the range of useful scales
1088: by the mean thickness of the slice. Complex sample borders are also
1089: a nuisance when applying the wavelet cascade. In order to take
1090: account of these difficulties, we have yet to resort to running
1091: Monte-Carlo simulations of Gaussian realizations of a right power
1092: spectrum, and to compare the obtained distributions of the
1093: functionals with the functionals for the galaxy data. We hope that
1094: for the future surveys (e.g., the full SDSS), the data volume will
1095: be large enough to do without Monte-Carlo runs.
1096:
1097:
1098:
1099: \section{Morphology of the 2dF19 sample}
1100: \label{sec:datamorph}
1101:
1102: Having developed all necessary tools, we apply them to the 2dF19
1103: volume-limited sample, separately for the NGC and SGC regions.
1104:
1105: We show selected slices for the two subsamples, first
1106: (Figs.~\ref{fig:2df19Nbunch2}--\ref{fig:2df19Sbunch2}). The Northern
1107: slice was chosen to show the richest super-cluster in the 2dfGRS
1108: NGC, super-cluster 126 (middle and low, \citep{maret97}). The
1109: Southern slice has the maximum area in the $z=\mbox{const}$ slices
1110: of this sample. We choose the gray levels to show also the mask
1111: area.
1112:
1113: \begin{figure*}
1114: \centering
1115: \resizebox{.45\textwidth}{!}{\includegraphics*{2dfNmf3.eps}}
1116: \resizebox{.45\textwidth}{!}{\includegraphics*{2dfSmf3.eps}}
1117: \caption{Summary of the densities of the fourth MF $v_3$ for the
1118: data and all wavelet orders for the 2dFN19 sample (left) and
1119: 2dFS19 sample (right), in the logn
1120: mapping -- the higher the wavelet order (indicated by labels), the lower the density
1121: amplitude. Thick lines show reference Gaussian predictions. Thin
1122: lines stand for the $\sqrt2$~Mpc/$h$ grid. \label{fig:2dfNmf3}}
1123: \end{figure*}
1124: Figs.~\ref{fig:2dfNmf3}--\ref{fig:NSw1mf3} show the
1125: results -- the Minkowski functionals for
1126: the wavelet decompositions of the 2dFN19 and 2dFS19
1127: subsamples.
1128:
1129: First, we show the summary figures: the density $v_3$ for the
1130: original data and for wavelet orders from 0 to 3, and for grid unit
1131: of $\sqrt{2}$~Mpc/$h$. Higher wavelet orders are not usable -- there
1132: the boundary rules used for the wavelet cascade influence strongly
1133: the results, and the number of independent resolution elements
1134: becomes very small, letting the functionals to be dominated by
1135: noise. The wavelet orders used span the scale range 2--22.6~Mpc/$h$.
1136: In order to show all the densities in the same plot, we use the
1137: mapping
1138: \[
1139: \mbox{logn}(v_3)=\mbox{sgn}(v_3)\log(1+|v_3|).
1140: \]
1141: It is almost linear for $|v_3|<1$,
1142: logarithmic for $|v_3|>1$, and can be applied to negative
1143: arguments, also. The density in Fig.~\ref{fig:2dfNmf3} is shown with
1144: dotted lines. For reference, the two thick lines show the Gaussian
1145: predictions for small and large functional density amplitudes for
1146: the logn($v_3$) mapping. The first glance at the figures of the
1147: fourth Minkowski functional reveals that none of the wavelet scales
1148: shows Gaussian behavior.
1149:
1150:
1151:
1152:
1153: In order to estimate the spread in the values of the functionals, we
1154: ran about 100 Gaussian realizations for every sample and grid,
1155: generated wavelets and found the Minkowski functionals. The power
1156: spectrum for these realizations was chosen as described in
1157: Sec.~\ref{sec:gauss} above (see \citet{klypin97}), and smoothed by a
1158: Gaussian of $\sigma=1$ (in grid units). This is practically
1159: equivalent to the $B_3$ extirpolation used to generate the observed
1160: density on the grid. Now, if the observational MF-s lie outside the
1161: limiting values of these realizations, we can say that the Gaussian
1162: hypothesis is rejected with the $p$-value less than 1\%. We also
1163: calculated the multiscale functionals for a set of 22 mock samples,
1164: specially created for the 2dFGRS \cite{norberg02}. The mock catalogs
1165: were extracted from the Virgo Consortium $\Lambda$CDM Hubble volume
1166: simulation, and a biasing scheme described in \citet{cole98} was
1167: used to populate the dark matter distribution with galaxies.
1168:
1169: Fig.~\ref{fig:Nw2mf3} and ~\ref{fig:Sqw2mf3} show respectively the
1170: densities of the fourth Minkowski functional for the Northern and the Southern
1171: Galactic caps. The MF density $v_3$ corresponding to the data is
1172: plotted with a continuous line while error bars correspond to the
1173: total variation for mocks. The minimum and maximum limits for
1174: Gaussian realizations are plotted with dotted lines.
1175:
1176: The Gaussian realizations show very small spread, and are clearly
1177: different from the $v_3$ Minkowski Functional of the observational
1178: samples. Results from mocks are much closer to the data.
1179:
1180: Fig.~\ref{fig:Nw2mf3} shows clear non-Gaussianity for the north
1181: galactic cap at a high confidence level. Gaussian realizations are
1182: not much deformed by the combination of boundary conditions
1183: (wavelets) and border corrections (functionals) effects. In this
1184: figure we can appreciate that with respect to this MF, mocks follow
1185: data well for smaller density iso-levels, but deviate around
1186: $\nu=1$; we have seen similar effects before \citep{mart05}. An
1187: interesting detail is the knee around $\nu=0.5$, seen both in the
1188: data and in the mocks, but not for Gaussian realizations. The latter
1189: fact tells us that it is not caused by the specific geometry of the
1190: data sample. The grid step was 1Mpc/$h$.
1191:
1192:
1193: \begin{figure}
1194: \centering
1195: %\resizebox{.4\textwidth}{!}{\includegraphics*{mf/Nw2mf3.eps}}
1196: \resizebox{.4\textwidth}{!}{\includegraphics*{Nw2mf3.eps}}
1197: \caption{The density of the fourth MF $v_3$ for
1198: the wavelet order 2 for the 2dFN19 sample (full line).
1199: Dotted lines show the minima and maxima of 102
1200: Gaussian realizations, and bars show the full variation in a
1201: sample of 22 mock catalogues.
1202: \label{fig:Nw2mf3}}
1203: \end{figure}
1204:
1205: \begin{figure}
1206: \centering
1207: %\resizebox{.4\textwidth}{!}{\includegraphics*{mf/Sqw2mf3.eps}}
1208: \resizebox{.4\textwidth}{!}{\includegraphics*{Sqw2mf3.eps}}
1209: \caption{The density of the fourth MF $v_3$ for
1210: the wavelet order 2 (grid $\sqrt2$) for the 2dFS19 sample (full line).
1211: Dotted lines show the minima and maxima of 108
1212: Gaussian realizations, and bars show the full variation in a
1213: sample of 22 mock catalogues.
1214: \label{fig:Sqw2mf3}}
1215: \end{figure}
1216:
1217:
1218: \begin{figure}
1219: \centering
1220: %\resizebox{.4\textwidth}{!}{\includegraphics*{mf/NSw1mf3.eps}}
1221: \resizebox{.4\textwidth}{!}{\includegraphics*{NSw1mf3.eps}}
1222: \caption{The densities of the fourth MF $v_3$ for
1223: the wavelet order 1 for the 2dFS19 and 2dFN19 samples (full lines).
1224: Bars show the full variation in two
1225: samples of 22 mock catalogues.
1226: \label{fig:NSw1mf3}}
1227: \end{figure}
1228:
1229: Fig.~\ref{fig:Sqw2mf3} for the Southern data shows also clear
1230: non-Gaussianity, but no strong features like the northern one (it
1231: describes larger scales, as this wavelet is based on the $\sqrt{2}$
1232: Mpc/$h$ step grid). An interesting point is that mocks follow the
1233: data curve almost perfectly here, much better than for the Northern
1234: sample.
1235:
1236: Fig.~\ref{fig:NSw1mf3} shows the $v_3$ MF density at the first
1237: wavelet scale for both north and south slices. Again, it is clearly
1238: non-Gaussian. However, it is remarkable how well the functionals for
1239: both volumes coincide. This shows that the border and boundary
1240: effects are small, and we are seeing real features in the density
1241: distribution that we have to explain. Also, the mocks follow the
1242: data rather well. This means that the structure (and galaxy) formation
1243: recipes used to build the mocks already implicitly
1244: include mechanisms responsible for these features.
1245:
1246: It is useful to compare our results with the recent careful analysis of
1247: the topology of the SDSS galaxy distribution by \citet{park05}.
1248: They used Gaussian kernels to find the density distribution, and as a result,
1249: their genus curves (see, e.g., their Figs.~6 and 8) are close to those of
1250: Gaussian fields. They describe deviations from Gaussianity by moments of
1251: the genus curve, taken in carefully chosen $\nu$ intervals, and normalized
1252: by corresponding Gaussian values. Our Minkowski functionals differ so much
1253: from Gaussian templates (see Fig.~\ref{fig:2dfNmf3}) that we cannot fit a reference
1254: Gaussian curve. The only analogue we can find is the shift of the genus curve $\Delta\nu$,
1255: that we estimate by fitting an expression
1256: $v_3(\nu)=A\left((\nu-\Delta\nu)^2-1\right)\exp\left(-(\nu-\Delta\nu)^2/2\right)$
1257: to our results. The values of the shift corroborate the visual impression of
1258: strong differences between our results and those of \cite{park05}. While they find
1259: that $\Delta\nu$ lies in the interval $[-0.1,0.26]$ (for the scale range
1260: $R_G\in[4.5,11.0]$~Mpc/$h$, their table 2), our $\Delta\nu$ assumes values between
1261: $-1.3$ and $-0.2$, for approximately the same scale interval
1262: ($\lambda\in[4,22.6]$~Mpc/$h$).
1263: As the morphology of the 2dFGRS and the SDSS should not differ much, the difference
1264: is clearly caused by different kernels, Gaussians compared to compact wavelets.
1265: Dropping the conventional Gaussian
1266: kernels makes the discriminative force of the morphological tests
1267: considerably stronger.
1268:
1269: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1270:
1271: \section{Conclusions}
1272: \label{sec:concl}
1273:
1274: The main results of this paper are:
1275: \begin{enumerate}
1276: \item We have shown how to compute the MF, taking into account
1277: both the biases due to the discrete
1278: grid from the Crofton method and the border effects related to
1279: complex observational sample volumes.
1280:
1281: \item Our experiments have shown that the multiscale MF functionals of
1282: a Gaussian Random Field have always a Gaussian behavior, even in
1283: case where the field lies within complex boundaries.
1284: Therefore, we have established a solid base for calculating the Minkowski functionals
1285: for real data sets and their multi-scale decompositions.
1286:
1287: \item We found that both the observed galaxy density fields and mocks
1288: show clear non-Gaussian features of the morphological descriptors
1289: over the whole scale range we have considered. For smaller scales,
1290: this non-Gaussianity
1291: of the present cosmological fields should be expected,
1292: but it has been an elusive quality,
1293: not detected in most of previous papers (see, e.g., \citet{hoyle02},
1294: \citet{park05}). But even for the largest scales that the data allows us to study
1295: (about 20 Mpc/$h$),
1296: the density fields are yet not Gaussian. We believe that
1297: the Gaussianity reported in the papers cited above could be just a consequence of
1298: oversmoothing the data. This effect was clearly described in
1299: \citet{mart05}.
1300:
1301: \item The mocks that are generated from initial
1302: Gaussian density perturbations by gravitational evolution and by
1303: applying semi-analytic galaxy formation recipes, are pretty
1304: close to the data. However, as in a previous study \citep{mart05},
1305: we confirm a discrepancy around $\nu=1$ between the mocks and the
1306: data. This analysis clearly shows that there are more
1307: faint structures in the data than in the mocks,
1308: and clusters in the mocks have a larger intensity
1309: than in the real data.
1310:
1311: \end{enumerate}
1312:
1313: \section*{Acknowledgments}
1314:
1315: We thank Darren Croton for providing us with the 2dF volume-limited
1316: data and explanations and suggestions on the manuscript, and Peter
1317: Coles for discussions. We thank the anonymous referee for constructive
1318: comments. This work has been supported by the University of Valencia
1319: through a visiting professorship for Enn Saar, by the Spanish MCyT
1320: project AYA2003-08739-C02-01 (including FEDER), by the National
1321: Science Foundation grant DMS-01-40587 (FRG), and by the Estonian
1322: Science Foundation grant 6104.
1323:
1324: %\bibliography{multires}
1325:
1326: \begin{thebibliography}{}
1327: \bibitem[Adler(1981)]{adler} Adler, R.~J. 1981,
1328: The Geometry of Random Fields, (New York: John Wiley \& Sons)
1329: \bibitem[Ascasibar \& Binney(2005)]{ascasibar05}
1330: Ascasibar, Y., \& Binney, J., 2005, \mnras, 356, 872
1331: \bibitem[Cole et al.(1998)]{cole98}
1332: Cole, S., Hatton, S., Weinberg, D. H., \& Frenk, C. S.,
1333: 1998, \mnras, 300, 945.
1334: \bibitem[Coles et al.(1996)]{coles96}
1335: Coles, P., Davies, A. G., \& Pearson, R. C. 1996, \mnras, 281, 1375.
1336: \bibitem[Colless et al.(2003)]{2df} Colless, M., et al. 2003,
1337: astro-ph/0306581.
1338: \bibitem[Croton et al.(2004a)]{croton1} Croton, D.~J., et al.\
1339: 2004, \mnras, 352, 828
1340: \bibitem[Croton et al.(2004b)]{croton2} Croton, D.~J., et al.\
1341: 2004, \mnras, 352, 1232
1342: \bibitem[Donoho(1988)]{donoho88} Donoho, D. L., 1988,
1343: Ann. Statist., 16, 1390
1344: \bibitem[Einasto et al.(1997)]{maret97} Einasto, M., et al.\
1345: 1997, \aaps, 123, 119
1346: \bibitem[Gott et al.(1986)]{gott86}
1347: Gott, J.~R., Dickinson, M., \& Melott, A.~L. 1986, \apj, 306, 341
1348: \bibitem[Hamilton et al.(1986)]{ham86}
1349: Hamilton, A.~J.~S., Gott, J.~R.~I., Weinberg, D. 1996, \apj, 309, 1
1350: \bibitem[Hoyle et al. (2002)]{hoyle02}
1351: Hoyle, F., et al. 2002, \apj, 580, 663
1352: \bibitem[Joye \& Mandel(2003)]{ds9} Joye, W. A., \& Mandel, E.,\
1353: 2003, ADASS XII ASP Conf. Ser.. 295, 489
1354: \bibitem[Kerscher et al.(1997)]{ker2}
1355: Kerscher, M., et al. 1997, \mnras, 284, 73
1356: \bibitem[Klypin \& Holtzman(1997)]{klypin97}
1357: Klypin, A., \& Holtzman J. 1997,
1358: astro-ph/9712217
1359: \bibitem[Maddox et al.(1996)]{maddox96} Maddox, S.~J., Efstathiou, G.,
1360: \& Sutherland, W.~J. 1996, \mnras, 283, 1227
1361: \bibitem[Mallat(1999)]{mallat:book99}Mallat, S., 1999, A wavelet tour
1362: of signal processing (San Diego: Academic Press)
1363: \bibitem[Mart\'{\i}nez et al.(1993)]{martinez93}
1364: Mart\'{\i}nez, V.~J., Paredes, S., \& Saar, E. 1993, \mnras, 260,
1365: 365
1366: \bibitem[Mart{\'{\i}}nez \& Saar(2002)]{martinezsaar}
1367: Mart{\'{\i}}nez, V.~J., \& Saar, E. 2002,
1368: Statistics of the Galaxy Distribution, (Boca-Raton:
1369: Chapman \& Hall/CRC press)
1370: \bibitem[Mart{\'\i}nez et al.(2005)]{mart05} Mart{\'\i}nez, V.~J., et al.\
1371: 2005, \apj, 634, 744
1372: \bibitem[Mecke et al.(1994)]{cf:mecke94}
1373: Mecke K.~R., Buchert, T., \& Wagner, H. 1994, \aap, 288, 697
1374: \bibitem[Minkowski(1903)]{minkowski}
1375: Minkowski, H., 1903, Mathematische Annalen 57, 447
1376: \bibitem[Norberg et al.(2002)]{norberg02} Norberg, P., et al.\
1377: 2002, \mnras, 336, 907
1378: \bibitem[Park et al.(2005)]{park05}
1379: Park, C., et al. 2005, \apj, 633, 11
1380: \bibitem[Pelupessy et al.(2003)]{pelupessy03} Pelupessy, F.~I.,
1381: Schaap, W.~E., \& van de Weygaert, R.\ 2003, \aap, 403, 389
1382: \bibitem[Sahni et al.(1998)]{sahni98}
1383: Sahni, V., Sathyaprakash, B. S., \& Shandarin, S. F.,\
1384: 1998, \apj, 495, L5
1385: \bibitem[Schaap \& van de Weygaert(2000)]{schaap00}
1386: Schaap, W. E., \& van de Weygaert, R., 2000, \aap, 363, L29
1387: \bibitem[Schmalzing \& Buchert(1997)]{jens97}
1388: Schmalzing J., \& Buchert, T. 1997, \apj, 482, L1
1389: \bibitem[Schmalzing et al.(1996)]{ker96var}
1390: Schmalzing J., Kerscher M. and Buchert T. (1996),
1391: in Dark Matter in the Universe, eds. S.~Bonometto, J.R.
1392: Primack \& A.~Provenzale, (Amsterdam: IOS Press), 281
1393: \bibitem[Shandarin et al.(2004)]{shandarin04}
1394: Shandarin, S. F., Sheth, J. V., \& Sahni, V., 2004, \mnras,
1395: 353, 162
1396: \bibitem[Sheth \& Sahni(2005)]{sheth05} Sheth, J.~V., \& Sahni,
1397: V.\ 2005, Current Science, 88, 1101
1398: \bibitem[Sheth et al.(2003)]{surfgen}
1399: Sheth, J. V., Sahni, V., Shandarin, S. F., \& Sathyaprakash, B. S.,
1400: 2003, \mnras, 343, 22
1401: \bibitem[Shensa(1992)]{shensa92} Shensa,M.~J., 1992, IEEE Trans. Signal
1402: Proc., 40, 2464
1403: \bibitem[Starck \& Murtagh(2002)]{starck:book02}
1404: Starck, J.-L., \& Murtagh, F. 2002,
1405: Astronomical Image and Data Analysis, (Berlin: Springer-Verlag)
1406: \bibitem[Starck et al.(1998)]{starck:book98}
1407: Starck, J.-L., Murtagh, F. \& Bijaoui, A. 1998,
1408: Image Processing and Data Analysis: The Multiscale Approach,
1409: (Cambridge: Cambridge University Press)
1410: \bibitem[Starck \& Pierre(1998)]{stapie} Starck, J.-L.,
1411: Pierre, M.\ 1998, \aaps, 128, 397
1412: \bibitem[Starck et al.(1995)]{sta95_1}
1413: Starck J.-L., Bijaoui, A., Murtagh, F., 1995,
1414: CVGIP: Graphical Models and Image Processing, 57, 420.
1415: \bibitem[Starck et al.(2005)]{sta05_3}
1416: Starck, J.-L., Mart\'{\i}nez, V.~J., Donoho, D.~L., Levi, O.,
1417: Querre, P. \& Saar, E. 2005, Eurasip Journal of Signal Processing,
1418: 2005-15, 2455.
1419: \bibitem[Tomita(1990)]{tomita}
1420: Tomita, H., 1990, in Formation, Dynamics and Statistics of Patterns
1421: eds. K. Kawasaki, et al., Vol. 1, (World Scientific), 113
1422: \bibitem[York et al.(2000)]{sdss} York, D.~G., et al.\ 2000, \aj,
1423: 120, 1579
1424: \end{thebibliography}
1425:
1426:
1427:
1428: \section*{Appendix A: The \`a trous algorithm and galaxy catalogues}
1429: \label{sec:atrous}
1430:
1431: A good description of the \`a trous algorithm and of
1432: its applications to image processing in astronomy can
1433: be found in \citet{starck:book02}. Readers interested in the mathematical
1434: basis of the algorithm can consult \citet{mallat:book99} and
1435: \citet{shensa92}. We give below a short summary of the algorithm
1436: and describe the additional intricacies that arise when the
1437: algorithm is applied to galaxy catalogues (point data).
1438:
1439: We start with forming the initial density distribution $d_0$ on a
1440: grid. In order to form the discrete distribution, we have to
1441: weight the point data (extirpolate), using the scaling kernel
1442: for the wavelet \citep{mallat:book99}.
1443: As we shall use the $B_3$ box spline as the
1444: scaling kernel, the extirpolation step is:
1445: \beq
1446: \label{eq:extirp}
1447: d^{(0)}(\mathbf{n}_i)=\int\rho(\mathbf{x})B_3^{(3)}(\mathbf{x-n}_i)d^3x,
1448: \eeq
1449: where $\mathbf{n}_i\equiv(n)_i=(x_i,y_i,z_i)$ is a grid vertex,
1450: $\rho(\mathbf{x})$ is the original density, delta-valued at
1451: galaxy positions, and $B_3^{(3)}(\mathbf{x})$ is the direct
1452: product of three $B_3$ splines:
1453: \[
1454: B_3^{(3)}(\mathbf{x})=B_3(x)B_3(y)B_3(z),
1455: \]
1456: where $\mathbf(x)=(x,y,z)$.
1457: The $B_3$ spline is given by
1458: \[
1459: B_3(x)=\frac1{12}\left[|x-2|^3-4|x-1|^3+6|x|^3-4|x+1|^3+|x+2|^3\right].
1460: \]
1461: As this function is zero outside the cube $[-2,2]^3$, every data
1462: point contributes only to its immediate grid neighbourhood, and
1463: extirpolation is fast.
1464:
1465: The main computation cycle starts now by convoluting the data
1466: $d$ with a specially chosen discrete filter $h_{(k)}$:
1467: \beq
1468: \label{conv}
1469: d^{(I+1)}_{(n)}=\sum_{(k)}h_{(k)}d^{(I)}_{(n)+2^I(k)}.
1470: \eeq
1471: Here $I$ stands for the convolution order (octave),
1472: the 3-dimensional filter $h_{(k)}=h_lh_mh_n$,
1473: $(k)=(l,m,n)$
1474: is the direct product
1475: of three one-dimensional filters $h_i=\{1/16, 1/4,$
1476: $3/8, 1/4, 1/16\}$,
1477: for $i\in [-2,2]$. Two points should be noted:
1478: \begin{itemize}
1479: \item As the filter is the direct product of the one-dimensional
1480: filters, the convolution can be applied consecutively for each
1481: coordinate, and can be done in place, with extra
1482: memory only for a data line.
1483: \item The data index $(n)+2^I(k)$ in the convolution shows that the
1484: data is assessed from consecutively larger regions for further
1485: octaves, leaving intermediate grid vertices unused. This is
1486: equivalent to inserting zeroes in the filter for these points,
1487: and this is where the name of the method comes from (\`a trous is
1488: ``with holes'' in French). This makes the convolution very fast,
1489: as the number of operations does not increase when the filter
1490: width increases.
1491: \end{itemize}
1492:
1493: The filter $h_i$ is satisfies the dilation equation
1494: \[
1495: \frac12B_3\left(\frac{x}2\right)=\sum_k h_k B_3(x-k).
1496: \]
1497:
1498: After we have performed the convolution (\ref{conv}), we
1499: find the wavelet coefficients $w^{(J)}$ for the octave $J$ by
1500: simple substraction:
1501: \beq
1502: \label{wave}
1503: w^{(J)}_{(n)}=d^{(J)}_{(n)}-d^{(J+1)}_{(n)}.
1504: \eeq
1505: The combination of steps (\ref{conv},\ref{wave}) is
1506: equivalent to convolution of the data with the associated
1507: wavelet $\psi^{(3)}(x)$, where
1508: \beq
1509: \psi(x)=2B_3(2x)-B_3(x).
1510: \eeq
1511: Repeating the sequence (\ref{conv},\ref{wave}) we find
1512: wavelet coefficients for a sequence of octaves. The number
1513: of octaves is, evidently, limited by the grid size, and in real
1514: applications by the geometry of the sample.
1515:
1516: We have illustrated the wavelet cascade in the main text by application to
1517: real galaxy samples. As our wavelet amplitudes were
1518: obtained by subtraction, we can easily reconstruct the initial density:
1519: \beq
1520: \label{reconst}
1521: d^{(0)}_{(n)}=d^{(J+1)}_{(n)}+\sum_{j=0}^{j=J} w^{(j)}_{(n)}.
1522: \eeq
1523: Here the upper indices show the octave, the lower indices denote grid
1524: vertices, and $d^{(J+1)}$ is the result of the last convolution.
1525: This formula can also be interpreted as the decomposition of
1526: the original data (density field) into contributions from
1527: different scales -- the wavelet octaves describe contributions
1528: from a limited (dyadic) range of scales.
1529:
1530: Here we have to note that while the scaling kernel
1531: \[
1532: \Phi(x,y,z)=B_3(x)B_3(y)B_3(z)
1533: \]
1534: is a direct
1535: product of three one-dimensional functions, it is surprisingly
1536: almost isotropic. Its innermost iso-levels are slightly
1537: concave, and outer iso-levels tend to be cubic, but this
1538: happens at very low function values.
1539: In order to characterise the deviation from anisotropy, let us
1540: first define the angle-averaged scaling kernel
1541: \[
1542: \bar{\Phi}(r)=\frac1{4\pi}\int_S(r)\Phi(r,\theta,\phi)dS,
1543: \]
1544: where $S(r)$ is a spherical surface of a radius $r$.
1545: The anisotropy can now be calculated as the
1546: integral of the absolute value of the difference between the
1547: kernel and its angle-averaged value:
1548: \[
1549: \int_2^2\int_2^2\int_2^2\left|\Phi(x,y,z)
1550: -\bar{\Phi}(\sqrt{x^2+y^2+z^2})\right|\,dx\,dy\,dz=0.030.
1551: \]
1552: As the integral of the kernel itself is unity, the deviation is
1553: only a couple of per cent.
1554:
1555:
1556:
1557: A similar integral over the wavelet profile gives the value
1558: 0.052. Here the natural scale, the integral of the square of the wavelet
1559: profile, is also unity, so the deviation from isotropy
1560: is small.
1561:
1562: Isotropy of the wavelet is essential, if we want to be sure that our results
1563: do not depend on the orientation of the grid. This is usually
1564: assumed, but with a different choice of the scaling kernel
1565: this could easily happen.
1566:
1567: As our wavelet transform is not orthogonal, there remain correlations
1568: between wavelet amplitudes of different octaves. The Fourier transform
1569: of the $B_3$ scaling function is
1570: \[
1571: \hat{B}_3(\omega)=\left(\frac{\sin(\omega/2)}{\omega/2}\right)^4
1572: \]
1573: The Fourier transform of the associated wavelet (\ref{wave}) is
1574: \[
1575: \hat{w}(\omega)=\hat{B}_3(\omega/2)-\hat{B}_3(\omega).
1576: \]
1577:
1578:
1579: \begin{figure}
1580: \centering
1581: \resizebox{0.4\textwidth}{!}{\includegraphics*{b3four.eps}}
1582: \caption{The square of the Fourier transform of the wavelet for
1583: two neighbouring octaves.
1584: \label{fig:wavefour}}
1585: \end{figure}
1586: We show the square of the Fourier transform of the wavelet for
1587: two neighbouring octaves in Fig.~\ref{fig:wavefour}. As we see,
1588: the overlap between the octaves is not large, but substantial.
1589: This is the price we pay for keeping the wavelet transform shift
1590: invariant. We can now compare wavelet amplitudes for different
1591: octaves (scale ranges) at any grid vertex, but we have to
1592: keep in mind that the separation of scales is not complete.
1593: It may seem an unpleasant restriction that the wavelet scales have
1594: to increase in dyadic steps. It is not, in fact, as one can choose
1595: the starting scale (the step of the grid) at will.
1596:
1597: Before applying the wavelet transform to the data, we have to
1598: decide how to calculate the convolution (\ref{conv}) near the
1599: spatial boundaries of the sample. Exact convolution can be carried
1600: out only for periodic test data, and spatially limited data need
1601: special consideration. For density estimation, a useful method to
1602: deal with boundaries is to renormalise the kernel. This cannot be
1603: done here, as renormalisation would destroy the wavelet nature
1604: of our convolution cascade. The only assumptions that can be used
1605: are those about the behaviour of the density outside the boundaries
1606: of the sample. Let us consider, for example, the one-dimensional case
1607: and the data $d(i)$ known only for the grid indices $i>0$. The possible
1608: boundary conditions are, then:
1609: \[
1610: d(i;i<0)=\left\{ \begin{array}{r@{\quad:\quad}l}
1611: 0&\mbox{zero boundary}\\
1612: d(0)&\mbox{constant boundary}\\
1613: d(-i)&\mbox{reflecting boundary}.
1614: \end{array}\right.
1615: \]
1616: The constant boundary condition is rarely used; the most popular
1617: case seems to be the reflecting boundary. For brick-type sample
1618: geometries, where the coordinate lines are perpendicular to the
1619: sample boundary, this condition gives good results. However, in our
1620: case the sample boundary has a complex geometry, and reflections from
1621: nearby boundary surface details would soon interfere with each other.
1622:
1623:
1624:
1625: \section*{Appendix B. Comparing border corrections}
1626: \label{app:boundcorr}
1627: \begin{figure}
1628: \centering
1629: \resizebox{0.4\textwidth}{!}{\includegraphics*{cmdiffv1.eps}}
1630: \caption{Relative errors of border-corrected densities of the
1631: second MF $v_1$ for a realization of a Gaussian
1632: random field in the 2df19N sample mask. The case of the border
1633: correction chain is shown by solid line, the 'raw' correction case --
1634: by dotted line.
1635: \label{fig:cmdiffv1}}
1636: \end{figure}
1637:
1638:
1639: We noted above (Sec.~\ref{subsec:bound}) that alongside with
1640: the border correction chain (\ref{corrchain}) there is
1641: another possibility to correct for borders.
1642: In this case we ignore the vertices in
1643: the mask and do not build any basic elements if one of the vertices
1644: belong to the mask. This also means that we do not have to
1645: build the basic elements in the mask region. The latter fact makes the
1646: algorithm faster (about twice faster for the 2dF data), as
1647: the data region occupies usually only a fraction of the encompassing brick).
1648:
1649: \begin{figure}
1650: \centering
1651: \resizebox{0.4\textwidth}{!}{\includegraphics*{cmdiffv2.eps}}
1652: \caption{Relative errors of border-corrected densities of the
1653: third MF $v_2$ for a realization of a Gaussian
1654: random field in the 2df19N sample mask. The case of the border
1655: correction chain is shown by solid line, the 'raw' correction case --
1656: by dotted line.
1657: \label{fig:cmdiffv2}}
1658: \end{figure}
1659:
1660: We call this method the 'raw' border correction and compare
1661: it with the border-correction algorithm (\ref{corrchain})
1662: on the example of the realization of a Gaussian random field
1663: for the 2dF NGC region, smoothed by a Gaussian of $\sigma=3$~Mpc/$h$
1664: to ensure that we resolve the density distribution. (We used the same
1665: realization to compare the border-corrected and uncorrected case,
1666: in Sec.~\ref{subsec:bound}.) We combine the encompassing Gaussian brick
1667: with the 2dFN19 mask, calculate the Minkowski functionals for both
1668: border correction methods, and compare them with the functionals
1669: found for the periodic brick. We show below the relative errors
1670: of the functionals, defined as
1671: \[
1672: \varepsilon(v_i(\nu))=\frac{v_i(\nu)-v^b_i(\nu)}{\max_\nu|v^b_i(\nu)|},
1673: \]
1674: where $v^b_i(\nu)$ are the densities of the functionals for the brick.
1675: We cannot use $v^b_i(\nu)$ themselves to normalise the errors, as
1676: their values pass through zero, so we use their maximum absolute
1677: values.
1678:
1679: \begin{figure}
1680: \centering
1681: \resizebox{0.4\textwidth}{!}{\includegraphics*{cmdiffv3.eps}}
1682: \caption{Relative errors of border-corrected densities of the
1683: fourth MF $v_3$ for a realization of a Gaussian
1684: random field in the 2df19N sample mask. The case of the border
1685: correction chain is shown by solid line, the 'raw' correction case --
1686: by dotted line.
1687: \label{fig:cmdiffv3}}
1688: \end{figure}
1689:
1690: The relative differences are shown in
1691: Figs.~\ref{fig:cmdiffv1}-\ref{fig:cmdiffv3}. We see that both
1692: correcting methods give the results that do not differ from the true
1693: densities more than 10\% (12\% for $v_3$). We see also that the
1694: border correction chain (\ref{corrchain}) gives always better estimates of the
1695: functionals; the maximum error is 3--4\%, and the error is about three times
1696: smaller than that for the 'raw' border correction. Thus we use
1697: this chain throughout the paper.
1698:
1699: It is useful to recall, though, that the border corrections
1700: (\ref{corrchain}) are based on the assumption of homogeneity and
1701: isotropy of the data, which may not always be the case. The 'raw'
1702: border corrections do not rely on any assumptions, and are therefore
1703: useful for verifying the results obtained by the correction chain.
1704: \label{lastpage}
1705:
1706: \end{document}
1707: \endinput
1708: